245
Efecto de la pectina sobre la actividad de algunas enzimas digestivas y la digestión de lípidos Mauricio Espinal Ruiz Universidad Nacional de Colombia Facultad de Ciencias Departamento de Química Bogotá DC, Colombia 2016

Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

  • Upload
    dobao

  • View
    220

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Efecto de la pectina sobre la actividad de algunas enzimas

digestivas y la digestión de lípidos

Mauricio Espinal Ruiz

Universidad Nacional de Colombia

Facultad de Ciencias

Departamento de Química

Bogotá DC, Colombia

2016

Page 2: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Efecto de la pectina sobre la actividad de algunas enzimas

digestivas y la digestión de lípidos

Mauricio Espinal Ruiz

Tesis presentada como requisito para optar al título de:

Doctor en Ciencias Química

Director:

Carlos Eduardo Narváez Cuenca, PhD

Grupo de Investigación:

Estudio de los cambios químicos y bioquímicos de alimentos frescos y procesados

Universidad Nacional de Colombia

Facultad de Ciencias

Departamento de Química

Bogotá DC, Colombia

2016

Page 3: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Effect of pectin on the activity of some digestive enzymes

and the digestion of lipids

Mauricio Espinal Ruiz

Universidad Nacional de Colombia

Facultad de Ciencias

Departamento de Química

Bogotá DC, Colombia

2016

Page 4: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Effect of pectin on the activity of some digestive enzymes

and the digestion of lipids

Mauricio Espinal Ruiz

Thesis submitted in fulfillment of the requirements for the degree of:

Doctor of Science - Chemistry

Director:

Carlos Eduardo Narváez Cuenca, PhD

Research Group:

Study of the chemical and biochemical changes of fresh and processed foods

Universidad Nacional de Colombia

Facultad de Ciencias

Departamento de Química

Bogotá DC, Colombia

2016

Page 5: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Resumen

Aunque se ha demostrado que la pectina tiene varias funciones fisiológicas benéficas para la

salud humana, la información disponible sobre los mecanismos mediante los cuales la pectina es

capaz de ejercer dichas funciones es limitada. En esta tesis se estudiaron los mecanismos

mediante los cuales la pectina es capaz de ejercer sus funciones fisiológicas. Se evaluó el efecto

de la pectina sobre la actividad de las principales enzimas digestivas [lipasa pancreática, -

amilasa, fosfatasa alcalina y proteasa (quimotripsina)]. Entre las enzimas estudiadas, la lipasa

pancreática fue la enzima más susceptible de ser inhibida por la pectina. Se hizo especial énfasis

en el efecto de la pectina sobre el destino gastrointestinal de los lípidos, haciendo uso de una

emulsión como modelo experimental. Se utilizó un modelo de digestión in vitro que consistió en

la simulación de las fases oral, gástrica e intestinal con el objetivo de evaluar el efecto de la

pectina sobre la velocidad y la magnitud del proceso de digestión de lípidos. Las digestiones in

vitro mostraron que la velocidad y la magnitud del proceso de digestión fueron inhibidos cuando

se aumentó la concentración de pectina. Además, se encontró que la pectina de alto grado de

metoxilación (HMP) tuvo mayor capacidad de inhibir el proceso de digestión que la pectina de

bajo grado de metoxilación (LMP). Esto permitió sugerir que los mecanismos fisicoquímicos que

pueden explicar la influencia de la pectina sobre la digestión de lípidos son el incremento de la

viscosidad de los fluidos gastrointestinales, la floculación de lípidos y las interacciones existentes

entre la pectina y los lípidos, así como con los componentes gastrointestinales que participan en

el proceso de digestión (lipasa pancreática, sales biliares, CaCl2, y NaCl). Finalmente, se

encontró que el aumento de la viscosidad de los fluidos gastrointestinales por efecto de la pectina

y las interacciones electrostáticas (atractivas y repulsivas) inhibió la velocidad y la magnitud del

proceso de transferencia de masa de los compuestos nutricionales más importantes

(monosacáridos, aminoácidos y lípidos emulsificados). Los resultados obtenidos en esta tesis

podrían conllevar a la comprensión de la funcionalidad fisiológica de la pectina y cómo esta

funcionalidad puede verse afectada por las características estructurales de la pectina.

Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión, digestión in vitro.

Page 6: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Abstract

Although it has been demonstrated that pectin has several physiological functions beneficial to

human health, limited information is available concerning the mechanisms by which pectin is

able to exert such functions. In this thesis, the mechanisms by which pectin is able to exert its

physiological functions were studied. The effect of pectin on the activities of the major digestive

enzymes [pancreatic lipase, -amylase, alkaline phosphatase and protease (chymotrypsin)] was

evaluated. Among the studied enzymes, pancreatic lipase was the most likely to be inhibited by

pectin. Special emphasis was made on the effect of pectin on the gastrointestinal fate of lipids

(and pancreatic lipase) by using a corn oil-in-water emulsion as the experimental model. A

simulated in vitro digestion model consisting of oral, gastric, and small intestinal phases was used

to elucidate the impact of pectin on the rate and extent of the digestion process of emulsified

lipids. The simulated in vitro digestions revealed that both the rate and extent of the digestion

process of emulsified lipids were inhibited with increasing concentration of pectin. In addition,

high methoxylated pectin (HMP) was found to be more effective in inhibiting the digestion

process of emulsified lipids as compared to low methoxylated pectin (LMP). We suggest that the

physicochemical mechanisms that may account for the influence of pectin on the digestion of

emulsified lipids are the modification of the viscosity of the gastrointestinal (GI) fluids, the

flocculation of emulsified lipids, and the interactions between pectin and both the emulsified

lipids and the GI components involved in the lipid digestion process (e.g., pancreatic lipase, bile

salts, CaCl2, and NaCl). Finally, the increase of the GI fluids viscosity by means of pectin and

electrostatic interactions (repulsive and attractive) were found to affect the rate and extent of the

mass transfer process of the most important nutritional compounds (monosaccharides, amino

acids, and emulsified lipids). The results obtained in this thesis might lead to the comprehension

of the physiological functionality of pectin and how this functionality can be affected by the

structural characteristics of pectin.

Keywords: Pectin, methoxylation degree, digestive enzymes, emulsion, in vitro digestion.

Page 7: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Objetivos

Objetivo General

Estudiar los efectos que ejerce la pectina sobre la actividad de algunas enzimas digestivas y sobre

la digestión de lípidos en un sistema gastrointestinal simulado.

Objetivos específicos

1. Evaluar el efecto de la pectina sobre la actividad de las principales enzimas digestivas (lipasa,

proteasa y -amilasa) en soluciones modelo.

2. Evaluar el efecto de la pectina sobre la transferencia de masa de nutrientes (azúcares, lípidos

y aminoácidos) en un sistema de difusión controlada.

3. Evaluar el efecto de pectinas de alto y bajo grado de metoxilación sobre la digestión de

lípidos en un sistema gastrointestinal simulado y compararlo con diferentes fuentes de fibra

dietaria (quitosano y metil celulosa).

4. Evaluar las posibles interacciones moleculares de la pectina con cada uno de los principales

componentes del sistema gastrointestinal (lipasa pancreática, sales biliares y electrolitos).

Page 8: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Objectives

General objective

Study the effects exerted by pectin on both the activity of some digestive enzymes and lipid

digestion in a simulated gastrointestinal system.

Specific objectives

1. Evaluate the effect of pectin on the activity of the main digestive enzymes (lipase, protease,

and -amylase) in model solutions.

2. Evaluate the effect of pectin on the mass transfer of nutrients (monosaccharides, lipids, and

amino acids) in a diffusion controlled system.

3. Evaluate the effect of both low and high methoxylated pectins on lipid digestion in a

simulated gastrointestinal system, and compare them with different sources of dietary fiber

(chitosan and methyl cellulose).

4. Evaluate the possible molecular interactions of pectin with each of the main components of

the gastrointestinal system (pancreatic lipase, bile salts, and electrolytes).

Page 9: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Table of contents

Resumen – Abstract

Objetivos – Objectives

Chapter 1 General introduction 1

Chapter 2

Optimization of the reaction conditions affecting the activity of

digestive enzymes

34

Chapter 3 Inhibition of digestive enzyme activities by pectins in model solutions 56

Chapter 4

Impact of dietary fibers [methyl cellulose, chitosan, and pectin] on

digestion of lipids under simulated gastrointestinal conditions

86

Chapter 5

Impact of pectin properties on lipid digestion under simulated

gastrointestinal conditions: Comparison of citrus and banana passion

fruit (Passiflora tripartita var. mollisima) pectins

117

Chapter 6

Interaction of a dietary fiber (pectin) with gastrointestinal components

(bile salts, calcium, and lipase): A calorimetry, electrophoresis, and

turbidity study

154

Chapter 7

Effect of pectin on the mass transfer kinetics of monosaccharides,

amino acids, and a corn oil-in-water emulsion in a Franz diffusion cell

187

Chapter 8 General discussion 214

Concluding remarks 233

Academic production 234

Acknowledgements 236

Page 10: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

General introduction

Page 11: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

2

1.1. Background

In recent years, an increase in the incidence of cardiovascular disease (CVD) and coronary heart

disease (CHD) has increased among some European countries, United States, and also in

Colombia (Threapleton, Greenwood, Evans, Cleghorn, Nykjaer, Woodhead, et al., 2013). CVD

and CHD account for almost half (48 and 52%) of all deaths in Europe and Colombia,

respectively, and a third (33%) of all deaths in the United States (Bolívar-Mejía & Vesga-

Angarita, 2013). Epidemiologic studies have shown that consumption of grains, cereals, fruits,

and vegetables may lower the risk of suffering both CVD and CHD (Brownlee, 2011). Dietary

fiber is one of the components that may be responsible for the beneficial effect of these foods

(Hollmann, Themeier, Neese, & Lindhauer, 2013). The protective connection between the

consumption of dietary fiber and both CVD and CHD was proposed in the 1970s (Phillips & Cui,

2011). Many experimental studies have examined the relationship between different sources of

dietary fiber and controlling both CVD and CHD risk factors. Results from experiments have

shown that dietary fiber may lower the concentrations of glucose and cholesterol in blood

(Threapleton, et al., 2013), reduce blood pressure (Streppel, Arends, & van’tVeer, 2005),

promote body-weight loss (Howarth, Saltzman, & Roberts, 2001), and improve insulin sensitivity

(de Leeuw, Jongbloed, & Verstegen, 2004), thereby reducing the risk of CVD and CHD

mortalities. Studies regarding the effect of dietary fiber on the CVD and CHD risk factors have

been focused on the major representative sources of soluble dietary fiber: legumes, fruits, seeds,

and grains (Galisteo, Duarte, & Zarzuelo, 2008). Nevertheless, information of purified soluble

dietary fiber (e.g., pectin, chitosan, and methyl cellulose) is scarce. While information on the

physiological effects of soluble dietary fiber has been provided, no information on the molecular

mechanisms by which soluble dietary fiber may act on CVD and CHD risk factors, is available.

In this chapter, we present the theoretical background required to address this thesis. First, we

make a brief description of the definition, classification, and functional properties of dietary fiber.

Then, we focus on the structural characteristics and classification of pectin. Next, we make a

description of the structure and properties of emulsions. Finally, we emphasize on the use of an in

vitro digestion model for evaluating the gastrointestinal fate of emulsified lipids.

Page 12: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

3

1.2. Dietary fiber

1.2.1. Definition of dietary fiber

Currently, the most accepted definition of dietary fiber is that of the European Commission of

Nutrition and Health, which has defined dietary fiber as the carbohydrate polymers with three or

more monomeric units, which are neither digested nor absorbed in the upper gastrointestinal tract

(GIT) of humans (mouth, stomach, and small intestine), and belong to the following categories: i)

edible carbohydrate polymers which have been obtained from food raw material by physical,

enzymatic or chemical means and that have a beneficial physiological effect demonstrated by

accepted scientific evidence; and ii) edible synthetic carbohydrate polymers which also have a

beneficial physiological effect demonstrated by accepted scientific evidence (Cummings, Mann,

Nishida, & Vorster, 2009; DeVries, 2003; Harris & Pijls, 2013). Although dietary fiber is defined

as carbohydrate polymers, it has been stated that lignin and other compounds can be classified as

dietary fiber, as long as they are closely associated with the carbohydrate polymers in plant cell

walls and behave chemically as dietary fiber (Mann & Cummings, 2009; Phillips & Cui, 2011;

Turner & Lupton, 2011).

1.2.2. Classification of dietary fiber

Dietary fiber includes a large variety of carbohydrates with various constituent monosaccharide

compositions, molecular weights, physical properties, and physiological effects. Dietary fiber is

usually classified based on their physical properties, especially based on their solubility in water,

viscosity, and fermentability. Dietary fiber can be classified as soluble (viscous and fermentable)

and insoluble (non-viscous and slowly fermentable) (Ramulu & Udayasekhara Rao, 2003).

Solubility has proven to be a very convenient parameter in the understanding of the

physicochemical properties of dietary fiber, allowing a simple division into those that principally

have effects on glucose and lipid absorption from the small intestine (soluble), and those which

are slowly and incompletely fermented and have pronounced effects in the large intestine

(insoluble) (Cummings & Stephen, 2007).

Page 13: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

4

Figure 1.1. Classification of dietary fiber based on their digestibility in the upper gastrointestinal tract

(GIT) and their molecular weight. Grey box indicates the carbohydrates that can be classified as dietary

fiber (Cummings & Stephen, 2007). FOS, fructo-oligosaccharides; GOS, galacto-oligosaccharides.

It has been mentioned, nevertheless, that the classification of dietary fiber based on the water

solubility is less convenient, because solubility of dietary fiber can also be influenced by several

environmental factors, such as pH, ionic strength, and temperature (Cummings & Stephen, 2007).

Therefore, it was suggested to classify dietary fiber based either on their chemical structure (e.g.,

molecular weight) or on recognized physiological properties (e.g., digestibility in the upper GIT)

(Cummings & Stephen, 2007).

Figure 1.1 illustrates the classification of dietary fiber based on their digestibility in the upper

GIT (mouth, stomach, and small intestine) and their molecular weight. Figure 1.1 shows that

dietary fiber can be classified as non-digestible oligosaccharides (NDOs), resistant starch (RS),

and non-starch polysaccharides (NSPs).

Digestible

starch

Resistant

starch (RS) Glucans

Chitosan

Mannans

Fructans

Xylans

Pectins

Maltodextrins

Polysaccharides

Non-starch polysaccharides (NSPs)

Non-digestible

oligosaccharides

(NDOs)

FOS and GOS

Oligosaccharides

Non-digestible polysaccharides

Dietary fiber

Digestible Non-digestible

Monosaccharides

Disaccharides

Non-digestible

disaccharides

Lactulose

Digestibility in the upper GIT

Mo

lecu

lar

wei

gh

t

Cellulose

Methyl cellulose

Hemicellulose

Page 14: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

5

1.2.2.1. Non-digestible oligosaccharides (NDOs)

Oligosaccharides (from 3 to 10 monomeric units) which are not digested in the upper GIT can be

included as a dietary fiber source according to the definition of the European Commission of

Nutrition and Health (Harris & Pijls, 2013). Considering the extent of the definition, NDOs are

very diverse, including degradation products of NSPs and some synthetized oligosaccharides,

excluding mono and disaccharides. Therefore, non-digestible disaccharides such as lactulose

cannot considered as dietary fiber (Phillips & Cui, 2011). The most recognized NDOs are fructo-

oligosaccharides (FOS) and galacto-oligosaccharides (GOS), which have been demonstrated to

present prebiotic activity by promoting the growth of beneficial bacteria in the colon (Mussatto &

Mancilha, 2007). In addition, NDOs are suitable ingredients to be used in the fabrication of low-

caloric foods for diabetic patients (Mussatto & Mancilha, 2007).

1.2.2.2. Resistant starch (RS)

Chemically, all starch molecules are similar among them, since they have glucose residues linked

by -(1,4) and -(1,6) linkages (amylose and amylopectin fractions). Although human -amylase

can hydrolyze the -(1,4) linkages, some fractions of starch are not available to be digested in the

upper GIT. The portion of starch that is not digested in the upper GIT is known as RS. RS can be

divided into four groups (Fuentes-Zaragoza, Riquelme-Navarrete, Sánchez-Zapata, & Pérez-

Álvarez, 2010): i) RS1, which is physically inaccessible because of entrapment in a non-

digestible matrix; ii) RS2, which is physically inaccessible because its compacted structure; iii)

RS3, retrograded starch (crystalline structures of amylopectin); and iv) RS4, modified starches

obtained by chemical treatments, e.g., di-starch phosphate ester. Sources and properties of RS

have been reviewed in detail elsewhere (Fuentes-Zaragoza, Riquelme-Navarrete, Sánchez-

Zapata, & Pérez-Álvarez, 2010; Haralampu, 2000; Raigond, Ezekiel, & Raigond, 2015; Sajilata,

Singhal, & Kulkarni, 2006). Among the four types of RS, it is important to stress that RS3 is of

particular interest, because of its thermal stability. This allows RS3 to be stable in normal

processing conditions, and allows its use as a functional ingredient in a wide variety of food

systems (Haralampu, 2000).

Page 15: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

6

1.2.2.3. Non-starch polysaccharides (NSPs)

All non-digestible polysaccharides besides RS belong to this group. NSPs, therefore, include a

wide diversity of polysaccharides with several constituent monosaccharide composition and

linkage types, as described in Table 1.1. The majority of dietary fibers are NSPs. The diversity in

the chemical structures of NSPs leads to diverse physical and chemical properties, as well as

diversity in their physiological effects (Kumar, Sinha, Makkar, de Boeck, & Becker, 2012). NSPs

are usually categorized based on their water solubilities (Sasaki, Kohyama, & Yasui, 2004).

However, physical properties of NSPs depend on their chemical structures (e.g., molecular

weight and monosaccharide composition) as well as environmental factors, including how the

NSPs can be isolated and processed (Izydorczyk, Macri, & MacGregor, 1998; You, Xie, Liu, &

Gu, 2010). Some of the most representative NSPs are insoluble NSPs such as cellulose,

hemicellulose, and chitin that can be found in cereals, legumes, and crustaceans, respectively; and

soluble NSPs such as pectin, chitosan, and alginate that can be found in fruits, crustaceans, and

algae, respectively (Table 1.1).

1.2.3. Health effects of dietary fiber

Dietary fiber is recognized for having a positive role in regulating body weight (Slavin, 2005),

alleviating diabetes (Post, Mainous, King, & Simpson, 2012), preventing cardiovascular diseases

(Threapleton, et al., 2013), maintaining colon health (Brownlee, 2011), and preventing various

types of cancer (Kritchevsky & Story, 1982). In the description that follows, the mechanisms by

which dietary fiber can influence the health of the consumers at different sites in the GIT are

described.

1.2.3.1. Dietary fiber in the upper GIT

The presence of dietary fiber in food potentially reduce energy uptake (Brownlee, 2011). Dietary

fiber provides bulk volume with less energy as compared to other nutrients, thus reducing the

energy density of foods (Howarth, Saltzman, & Roberts, 2001; Slavin, 2005).

Page 16: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

7

Table 1.1. Examples of the diversity of non-starch polysaccharides (NSPs) (Cummings & Stephen, 2007).

Dietary Fiber Fiber Variability Main Constituents Sources

Cellulose Cellulose

Methyl Cellulose

Glucose

Methyl-glucose

Plants

Synthetic

-Glucans -(1,3)-(1,4)-glucan

-(1,3)-(1,6)-glucan Glucose Cereals

Hemicelluloses

Arabinoxylans

Xyloglucans

Glucomannan

Galactoglucomannan

Arabinose, xylose

Glucose, xylose

Mannose,glucose

Mannose, glucose, galactose

Cereals

Dicots and conifers

Amorphophallus konjac

Gymnospermae

Pectins

Homogalacturonan

Rhamnogalacturonan I

Rhamnogalacturonan II

Galacturonic acid

Galacturonic acid, rhamnose,

arabinose, galactose

Galacturonic acid, rhamnose,

arabinose, galactose

Fruits and vegetables

Plant exudate gums

Arabic gum

Ghatti gum

Karaya gum

Tragacanth gum

Arabinose, galactose,

rhamnose, glucuronic acid

Arabinose, galactose,

mannose, xylose, rhamnose

Galactose, rhamnose,

galacturonic acid

Arabinose, galactose, xylose,

rhamnose, fucose

Acacia senegal

Anodeissus latifolia

Sterculia spp.

Astragalus spp.

Mucilages Ispaghula husk

Xylose, arabinose, rhamnose,

galacturonic acid

Plantago ovate

Galactomannans Guar gum

Locust bean gum

Mannose, galactose

Mannose, galactose

Guar plant

Carob tree

Seaweed polysaccharides

Alginate

Carrageenan

Agar

Guluronic acid,

mannuronic acid

Galactose, anhydro galactose

Galactose, anhydro galactose

Brown seaweed

Red seaweed

Redseaweed

Chitin and chitosan Chitin

Chitosan

N-acetyl glucosamine

Glucosamine Crustaceans, insects

Microbial polysaccharides

Xanthan gum

Reuteran

Gellan gum

Glucose, mannose

Glucose, rhamnose

Glucose, rhamnose

Xanthomonas campestris

Lactobacillus reuteri

Pseudomonas elodea

Page 17: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

8

The effects of dietary fiber after consumption of foods start in the oral phase. Foods that are rich

in dietary fiber are usually recognized as less palatable (Yuan, Smeele, Harington, van Loon,

Wanders, & Venn, 2014), and they usually require a lot of mastication. The effects of dietary

fiber in the gastric and small intestinal phases are usually related to their physicochemical

features, mainly by increasing the viscosity of digesta (Wanders, Jonathan, van den Borne, Mars,

Schols, Feskens, et al., 2013). It has been shown that a high viscosity digesta may delay gastric

emptying and induce satiety (Burton-Freeman, 2000). Gastric emptying is one of the many

factors that have a role in satiation and regulation of food intake (Janssen, Vanden Berghe,

Verschueren, Lehmann, Depoortere, & Tack, 2011). In the small intestine, increased viscosity of

the digesta upon consumption of dietary fiber has been related to a decreased rate of nutrient

digestion and absorption (Torsdottir, Alpsten, Holm, Sandberg, & Tölli, 1991). Besides having a

role in the regulation of food intake, the reduction of the glycemic index can prevent the

incidence of type II diabetes (Post, Mainous, King, & Simpson, 2012). The effects of viscous

dietary fiber on glycemic response (Yuan, Smeele, Harington, van Loon, Wanders, & Venn,

2014), blood lipids (Jenkins, Wolever, Rao, Hegele, Mitchell, Ransom, et al., 1993), intestinal

enzymatic activity (Dunaif & Schneeman, 1981), and nutrient digestibility (Zhang, Li, Liu, Zang,

Duan, Yang, et al., 2013) have been documented. The effects of dietary fiber aforementioned

might be due to the increased viscosity of the GIT fluids. Dietary fiber without viscous properties

(such as insoluble dietary fiber), therefore, do not have considerable effect in the upper GIT

(Dikeman & Fahey, 2006; Wanders, et al., 2013). In addition, it has been demonstrated that

soluble dietary fiber with a high water holding capacity (e.g., chitosan, pectin, and methyl

cellulose) can reduce appreciably the free water content and increase the viscosity of the digesta

(Dikeman & Fahey, 2006).

1.2.3.2. Dietary fiber in the large intestine

After passing the upper GIT tract, dietary fiber reaches the large intestine. In the large intestine

there is a large community of microorganisms that utilizes the dietary fiber as energy source for

their growth (prebiotic activity). It has been proposed that the fermentation of dietary fiber by

these microorganisms is one of the main mechanism by which dietary fiber can influence health

Page 18: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

9

(Hamaker & Tuncil, 2014). Dietary fiber that can be fermented by these microorganisms are

capable to modify the composition and diversity of the microbiota (Hamaker & Tuncil, 2014).

For example, FOS and GOS are reported to stimulate the growth of Bifidobacteria and

Lactobacilli (Sims, Ryan, & Kim, 2014), which are considered to be beneficial for health

(Guarner & Malagelada, 2003). The final products of dietary fiber fermentation in the large

intestine are gasses (e.g., H2 and CH4) and short chain fatty acids (SCFAs). The SCFAs are

mainly acetic, propionic, and butyric acids. An increased amount of SCFAs in the digesta is

associated with a lower pH, which contributes to the inhibition of the growth of pathogenic

bacteria (Sun & O’Riordan, 2013). As mentioned above, dietary fiber has an important number of

physiological functions. However, the molecular mechanisms by which the dietary fiber is able to

exert their physiological functions remain unknown for most of the dietary fiber sources

available. Among the different sources of dietary fiber, pectin stands out because of its

recognized functional properties and its versatility to be used in several industrial applications.

Below, we provide an overview of the structural characteristics of pectin.

1.3. Pectin

1.3.1. Structure of pectin

Pectin is a polysaccharide with several applications in the pharmaceutical, food, and

biotechnology industries (Munarin, Tanzi, & Petrini, 2012). It has been used successfully for

many years in the food and beverages industries as a thickening and gelling agent, as well as a

colloidal stabilizer (Yapo, 2011). Pectin is a complex mixture of polysaccharides that comprises

about the third part of the cell wall of higher plants. The highest concentration of pectin is found

in the middle lamella of cell walls, with a gradual decrease toward the plasma membrane (Fruk,

Cmelik, Jemric, Hribar, & Vidrih, 2014). Different types of pectin fractions can be isolated from

the cell wall, including homogalacturonan (HG), rhamnogalacturonan I (RG-I), and

rhamnogalacturonan II (RG-II). Typically, HG is the most abundant polysaccharide, constituting

about 65% (mol/mol) of the pectin, whereas RG-I and RG-II constitute about 25 and 10%

(mol/mol), respectively (Ridley, O'Neill, & Mohnen, 2001).

Page 19: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

10

1.3.1.1. Homogalacturonan (HG)

HG is a linear chain of 1,4-linked -D-galactopyranosyl uronic acid (GalA) units in which some

of the carboxyl groups can be methyl esterified (methoxylated) in the C6 position, as shown in

Figure 1.2. HG can be partially O-acetylated at C2 and C3 positions, depending on the plant

source (Ridley, O'Neill, & Mohnen, 2001). The methoxylation degree (MD) is defined as the

percentage of carboxyl groups which have been methoxylated. If more than 50% of the carboxyl

groups are methoxylated, the pectin is called high methoxylated pectin (HMP), and less than that

methoxylation degree is called low methoxylated pectin (LMP). The MD is variable and affects

the overall physicochemical properties of pectin, especially, its electrical properties and capacity

to form calcium-mediated interactions between HG chains (Mohnen, 2008). In addition, the

physicochemical properties of HG are significantly affected by the spontaneous ionization of the

free carboxyl groups in water (–COOH + H2O –COO⊝ + H3O⊕) and by the binding capacity

of metal ions (divalent cations such as Ca2⊕ and Fe

2⊕), especially in LMP (Yapo, 2011).

Figure 1.2. Structure of homogalacturonan (HG). The pKa value of the carboxyl group is 3.5. In high-

acidic conditions (pH<3.5) the carboxyl groups will be protonated (neutral), whereas in low-acidic,

neutral, or alkaline conditions (pH>3.5), the carboxyl groups will be dissociated, forming the carboxylate

group (anionic). Each number represents a carbon atom of the ring (Ridley, O´Neill, & Mohnen 2001).

-(1,4) linkage

O-acetylation

O-acetylation

Methoxylation

CarboxylatepH>3.5

CarboxylpH<3.5

23

6

Page 20: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

11

The experimental pKa value of the free carboxyl group of pectin is 3.5 (Caffall & Mohnen, 2009).

Thus, in high-acidic conditions (pH<3.5) a large fraction of the carboxyl groups will be

protonated and unavailable for ionic cross-linking by cations, whereas in low-acidic, neutral, or

alkaline conditions (pH>3.5), the carboxyl groups of pectin will be partially or completely

ionized (depending on the pH), and available for ionic cross-linking by cations, especially

divalent cations such as Ca2⊕, forming the egg-box structures as shown in Figure 1.3 (Pérez,

Mazeau, & Hervé du Penhoat, 2000; Ridley, O'Neill, & Mohnen, 2001).

It has been suggested that HG are relatively rigid chains (Pérez, Mazeau, & Hervé du Penhoat,

2000). The binding of calcium ions by two HG chains facilitates their alignment, which

consequently allows the binding of another calcium ion, and so on along the HG chains, as

observed in the Figure 1.3. Subsequent additions of calcium ions promotes the formation of

trimers, tetramers and hexamers of HG chains, promoting the growing of the HG chains and the

formation of complex three-dimensional structures that are able to trap water and compounds of

nutritional interest (Munarin, Tanzi, & Petrini, 2012; Yang, Zhang, Hong, Gu, & Fang, 2013;

Yapo, 2011).

Figure 1.3. The egg-box structures in homogalacturonan. Adapted from http://genialab.de.

Egg-box structures

(LMP chain)

Rhamnosyl fold

HMP

chain

Ca2+

Page 21: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

12

1.3.1.2. Rhamnogalacturonan I (RG-I)

RG-I is a branched polymer with a backbone of the repeated disaccharide [4)--D-GalA-

(12)--L-Rha-(1]n, where Rha corresponds to rhamnose (Silva, Jers, Meyer, & Mikkelsen,

2016). The Rha units in the backbone can be substituted with -(1,4) galactan, branched

arabinan, and arabinogalactan branches (Figure 1.4). The predominant branches of RG-I contain

both linear and branched -L-arabinofuranosyl, and -D-galactopyranosyl residues (Figure 1.4).

In addition, the glycosyl residues -L-fucosyl, -D-glucuronosyl, and 4-O-methyl--D-

glucuronosyl may also be present (Khodaei & Karboune, 2013), as may be polyphenolics such as

ferulic and coumaric acids (Oosterveld, Pol, Beldman, & Voragen, 2001). As the structure of

RG-I is very heterogeneous, there is no knowledge of how its structure relates to functionality

(Silva, Jers, Meyer, & Mikkelsen, 2016). It has been suggested, nevertheless, that RG-I is

responsible for preventing the linear backbone of HG to form multivalent associations (Pérez,

Rodríguez-Carvajal, & Doco, 2003). In addition, RG-I may control the interaction of HG with

other cell wall components such as proteins and cellulose (Khodaei & Karboune, 2013).

Figure 1.4. Schematic structure of pectin. Pectin consists of three different types of polysaccharides,

including homogalacturonan (HG), rhamnogalacturonan I (RG-I), and rhamnogalacturonan II (RG-II).

Kdo, 3-deoxy-D-manno-2-octulosonic acid; D-Dha, 3-deoxy-D-lyxo-2-heptulosaric acid. Adapted from

http://plantphysiol.org (Plant Physiology. 153 (2010): 384-395).

RG-II RG-IHG

D-Galacturonic acid

L-Rhamnose

D-Glucoronic acid

Kdo

L-Arabinose

D-Galactose

L-Aceric acid

D-Dha

L-Apiose

L-Fucose

D-Xylose

L-Galactose

O-Acetyl

O-Methyl

Borate

Page 22: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

13

1.3.1.3. Rhamnogalacturonan II (RG-II)

RG-II is present in the primary cell wall of higher plants (Pérez, Rodríguez-Carvajal, & Doco,

2003). RG-II is quite different to RG-I since its backbone is composed of 1,4-linked -D-GalA

units rather than the repeated disaccharide [4)--D-GalA-(12)--L-Rha-(1]n (Ishii &

Matsunaga, 2001). Complex branches are attached to C2 and C3 positions in the backbone to

form RG-II (Figure 1.4). These branches are composed of up to 12 types of glycosyl units,

linked together by at least 22 different glycosidic bonds (Ridley, O'Neill, & Mohnen, 2001).

Some of the monosaccharide units and glycosidic linkages found in RG-II branches are

considered unique in plant polysaccharides, such as L-aceric acid, L-fucose, L-apiose, 3-deoxy-

D-manno-2-octulosonic acid, and 3-deoxy-D-lyxo-2-heptulosaric acid (Pérez, Rodríguez-

Carvajal, & Doco, 2003). Frequently, RG-II forms a dimer mediated by borate ion (BO33⊝)

attached to L-apiose units (Whitcombe, O'Neill, Steffan, Albersheim, & Darvill, 1995). RG-II

dimerization is a very important process for ensuring the integrity of the cell wall (Kaneko, Ishii,

& Matsunaga, 1997). Therefore, although RG-II is a minor component of pectin, it has an

essential role in the pectin structure stability (Whitcombe, O'Neill, Steffan, Albersheim, &

Darvill, 1995).

1.3.2. Classification of pectin

During the early work with pectins, different names were reported in literature, leading to

confusion. Because of this, a committee was organized by the American Chemical Society (ACS,

1944), giving a rigid classification of pectin. Pectins are mainly classified as function of both

molecular weight and MD since they are the main parameters influencing their physicochemical

and functional properties (Yapo, 2011) (Figure 1.5). Pectins may be grouped into three

categories: protopectins, pectinic acids, and pectic acids (Worth, 1967). Protopectin can be found

in immature fruits, whereas pectinic and pectic acids can be found in ripe and overripe fruits,

respectively. In addition, pectinic and pectic acids can be obtained by enzymatic hydrolysis of

protopectin and they represent the wide variety of naturally processed pectins (Worth, 1967).

Page 23: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

14

1.3.2.1. Protopectin

Protopectin is a water-insoluble pectic substance which upon hydrolysis (either enzymatic or

chemical), yields pectic and pectinic acids. Protopectins are characterized by their high molecular

weights (MW; n>2000, where n is the number of GalA units per molecule of pectin) as well as

their high MD (100% mol/mol).

1.3.2.2. Pectic acids

Pectic acids are pectic substances characterized by their intermediate MW (n ranging from 50 to

2000) and for having all of the carboxyl groups completely free [0% (mol/mol) MD, LMP]. The

salts of pectic acids are referred as pectates.

1.3.3.3. Pectinic acids

Pectinic acids are pectic substances characterized by their intermediate MW (n ranging from 50

to 2000) and for having variable MD. Pectinic acids can be classified as LMP and HMP

depending on their MD [LMP if MD<50% (mol/mol); and HMP if MD>50% (mol/mol)]. The

salts of pectinic acids are referred as pectinates.

As aforementioned, pectins are a group of polysaccharides with complex three-dimensional

structures which are responsible for their functional properties. It has been postulated that the

presence of pectins in the upper GIT will result in a decreased rate of intestinal digestion of a

range of nutrients, including carbohydrates, lipids and proteins (Brownlee, 2011). Furthermore,

the molecular mechanisms involved in the interaction of pectin with gastrointestinal components

as well as the control of lipid digestion by means of pectin are complex and remain under

investigation. As part of the experimental chapters we use emulsified lipids as a model for

evaluating the effect of the physicochemical properties of pectins on the rate and extent of the

lipid digestion process. Below, we provide an overview of the emulsion characteristics,

preparation, and the main factors affecting their long-term stability.

Page 24: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

15

Figure 1.5. Classification of pectic substances according to the American Chemical Society (ACS, 1944). MD, methoxylation degree; MW,

molecular weight; n, number of galacturonic acid units per molecules of pectin; LMP, low methoxylated pectin; HMP, high methoxylated pectin.

Page 25: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

16

1.4. Emulsions

1.4.1. General characteristics of emulsions

An emulsion is formed by two immiscible liquids (usually oil and water), with one of the liquids

dispersed as small spherical droplets in the other one (Figure 1.6). Emulsions can be classified

according to the spatial distribution of the oil and aqueous phases (Rodríguez-Abreu & Lazzari,

2008). An emulsion consisting of oil droplets dispersed into an aqueous phase is called an oil-in-

water emulsion (o/w emulsion), whereas an emulsion consisting of water droplets dispersed into

an oil phase is called water-in-oil emulsion (w/o emulsion). The compound forming the droplets

in an emulsion is referred to as the dispersed phase, whereas the compound forming the bulk

liquid is referred to as the continuous phase (Leal-Calderon, Thivilliers, & Schmitt, 2007). The

process of converting two separate bulk immiscible liquids into an emulsion and of reducing the

size of the droplets in a coarse emulsion is known as homogenization (Yang, Marshall-Breton,

Leser, Sher, & McClements, 2012). This process is often performed by using mechanical devices

known as homogenizers, which subject the bulk liquids to strong mechanical mixing (Kizling,

Kronberg, & Eriksson, 2006).

Figure 1.6. An example of an emulsion, consisting of oil droplets dispersed in an aqueous phase (o/w

emulsion). Picture obtained from McClements (2010).

Page 26: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

17

The production of an emulsion directly from two separate bulk liquid phases can be defined as

primary homogenization, whereas the reduction in size of the droplets in an coarse emulsion can

be defined as secondary homogenization (Figure 1.7) (Elwell, Roberts, & Coupland, 2004). It is

feasible to form an emulsion by homogenizing oil and water together. However, the two phases

rapidly separate into an oil layer on top and a water layer at the bottom. The driving force of this

separation process is the fact that the contact between the oil and water phases is

thermodynamically unfavorable (Kabalnov, 1998). However, it is possible to produce long-term

kinetically stable (metastable) emulsions by adding substances known as emulsifiers (Dickinson,

2015). An emulsifier is a compound (e.g., proteins, polysaccharides, surfactants, and

phospholipids) that can be used to improve the stability of an emulsion (Dickinson, 2009).

Emulsifiers are surface-active compounds that absorb onto the surface of the lipid droplets during

the homogenization process, forming a layer which prevents them to aggregate (Ozturk &

McClements, 2016). All in all, the chemical nature of the ingredients composing the emulsion

(oil phase, aqueous phase, and emulsifier), as well as the homogenization process used for

fabricating the emulsion, are determining factors defining the physicochemical properties of the

emulsion and its further stability (Dickinson, 2009; Kabalnov, 1998; Leal-Calderon, Thivilliers,

& Schmitt, 2007).

Figure 1.7. Homogenization can be conveniently classified into two categories: primary and secondary

homogenization. Primary homogenization is the conversion of two bulk liquids into a coarse emulsion,

whereas secondary homogenization is the reduction of the droplet size in an existing coarse emulsion to

form a fine emulsion. Picture obtained from McClements (2010).

Water

Oil

Primary

homogenization

Secondary

homogenization

Bulk

liquids

Coarse

emulsion

Fine

emulsion

Page 27: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

18

1.4.2. Physicochemical properties of emulsions

1.4.2.1. Particle size distribution

The most important properties of emulsions such as shelf life, appearance, texture, and flavor are

influenced by the size of the lipid droplets (Huang, Kakuda, & Cui, 2001). For example, the

stability of an emulsion decreases as the droplet size increases because larger droplets are highly

susceptible to aggregation (Saiki, Horn, & Prestidge, 2008). Lipid droplets are formed by

homogenization and its particle size depends on the chemical nature of the components of the

emulsion, the environmental conditions, and the homogenization method (Guzey & McClements,

2006). If the lipid droplets of an emulsion are of the same particle size, it is referred to as a

monodisperse emulsion, whereas if there is a range of lipid droplet sizes in the emulsion, it is

referred to as a polydisperse emulsion (Figure 1.8). The droplet size of a monodisperse emulsion

can be characterized by a single number (e.g., droplet diameter or radius). Monodisperse

emulsions are often produced for fundamental studies since the interpretation of the experimental

measurements is much easier than for polydisperse emulsions. Emulsions in food systems,

nevertheless, contain a wide distribution of lipid droplet sizes, and therefore, the characterization

of their droplet sizes is more complicated than that for monodisperse emulsions. Therefore, the

whole particle size distribution in polydisperse emulsions should be reported (Guzey &

McClements, 2006).

Figure 1.8. Schematic representation of monodisperse and polydisperse emulsions (McClements, 2010).

Monodisperse

emulsion

Polydisperse

emulsion

Page 28: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

19

1.4.2.2. Surface electrical charge

The physicochemical and sensory properties of emulsions are influenced by both the magnitude

and sign of the electrical charge on the lipid droplets (Guzey & McClements, 2006). The origin

of this charge is usually the absorption of emulsifier molecules possessing an electrical charge

(Ichikawa, Dohda, & Nakajima, 2006). Emulsifiers often have ionizable groups that may be

neutral, positively charged, or negatively charged, depending on the pH. Therefore, an emulsion

droplet may have an electrical charge depending on the surfactant type and the pH of the aqueous

phase. The electrical charge of a droplet can be characterized by several ways (Snarski & Dunn,

1991), including the surface charge density (), the electrical surface potential (0), and the zeta-

potential (). The surface charge density is the electrical charge per unit surface area, whereas the

electrical surface potential is the free energy required to increase the surface charge density from

0 to . The zeta-potential is the net surface potential of a droplet and it takes into account the

electrical charge of the species in the surrounding medium, as well as the environmental

conditions such as the ionic strength and the pH (McClements, 2004). Zeta-Potential influences

the interaction between emulsion droplets and other charged species such as biopolymers,

surfactants, vitamins, antioxidant, flavors, and minerals (Leunissen, van Blaaderen,

Hollingsworth, Sullivan, & Chaikin, 2007). These interactions usually have significant

implications for the overall stability and quality of an emulsion. For example, the volatility of a

flavor can be reduced if the flavor is electrostatically attracted to the surface of the lipid droplet,

altering the flavor profile of the emulsion (Given, 2009), or the susceptibility of lipid droplets to

oxidation depends on the attraction ability of the metal ion to the lipid droplet surface (Coupland

& McClements, 1996).

1.4.3. Stability of emulsions

The term emulsion stability is widely used to describe the capacity of an emulsion to resist

changes in its properties over time (Kabalnov, 1998; Kizling, Kronberg, & Eriksson, 2006). The

more stable the emulsion, the more slowly its properties change. Nevertheless, there are a wide

variety of physicochemical mechanisms that may account for the alteration in emulsion

Page 29: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

20

physicochemical properties. The most common physical mechanisms that are responsible for the

instability of emulsions are shown schematically in Figure 1.9. A brief description of the main

physical mechanisms of emulsion instability is given below: Gravitational separation, droplet

aggregation, and phase inversion.

1.4.3.1. Gravitational separation

Lipid droplets in an emulsion possess a different density to that of the continuous phase

surrounding them, and therefore a gravitational and a buoyant force acts on them (Robins, 2000).

Creaming and sedimentation are mechanisms of gravitational separation. Creaming implies the

upward movement of the lipid droplets because of their lower density as compared to the

continuous phase (Robins, 2000), whereas sedimentation implies the downward movement of the

lipid droplets because of the higher density as compared to the continuous phase (Yan &

Masliyah, 1993). The densities of the most of edible oils are lower than that of water. Therefore,

there is a trend for lipid droplets to accumulate at the top of an emulsion and water at the bottom.

Thus, the lipid droplets in an o/w emulsion tend to cream, whereas those in a w/o emulsion tend

to sediment (Basaran, Demetriades, & McClements, 1998). Gravitational separation is often

considered as having an adverse effect on the quality of emulsions. The separation of an emulsion

into an opaque cream layer on top and a clear aqueous layer at the bottom is undesirable (Robins,

2000). In addition, the textural attributes of an emulsion are also negatively affected by

gravitational separation, since the cream layer on top tends to be more viscous than the aqueous

layer on bottom (Robins, Watson, & Wilde, 2002).

1.4.3.2. Droplet aggregation

The lipid droplets in an emulsion are in continual movements because of the effects of the

thermal energy, gravity, or mechanical forces (e.g., agitation), and therefore, they can collide to

each other. After a collision takes place, the emulsion droplets may either move away or

aggregate, depending on the magnitude of both the attractive and repulsive interactions between

them (Saiki, Horn, & Prestidge, 2008). Flocculation and coalescence are both types of droplet

Page 30: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

21

aggregation. Flocculation takes place when two or more lipid droplets collide and form an

aggregate in which lipid droplets maintain their individual integrity (Starov & Zhdanov, 2003).

Flocculation increases the rate of the gravitational separation in o/w emulsions, which is often

undesirable since flocculation reduces their shelf life (Dickinson, 2010). Flocculation also

induces a pronounced increase of the emulsion viscosity, which may lead to the formation of a

gel network structure (Starov & Zhdanov, 2003).

Coalescence is the process by which two or more droplets collide to form a single large droplet

(Tcholakova, Denkov, Ivanov, & Campbell, 2006). Coalescence is the main mechanism by which

an emulsion evolves toward its most thermodynamically stable state since coalescence involves

an appreciable decrease in the contact area between the oil and water phases (Krebs, Schroën, &

Boom, 2013). Extensive lipid droplet coalescence eventually leads to the phases separation,

forming a layer of oil on top, which is known as the oiling-off process (Degner, Chung, Schlegel,

Hutkins, & McClements, 2014).

Figure 1.9. Emulsions may become unstable through a variety of physical mechanisms, including

creaming, sedimentation, flocculation, coalescence, and phase inversion (McClements, 2010).

Creaming Sedementation Flocculation Coalescence

Phase

inversionKinetically stable

emulsion

Page 31: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

22

1.4.3.3. Phase inversion

Phase inversion is the process by which an o/w emulsion is converted into a w/o emulsion or vice

versa (Perazzo, Preziosi, & Guido, 2015). Phase inversion is usually produced by alterations in

the composition or environmental conditions of an emulsion, for example, modification of the

disperse phase volume fraction (emulsion concentration), emulsifier type and concentration,

solvent presence, temperature, or mechanical agitation (Rao & McClements, 2010). Phase

inversion is an important step in the production of a wide number of food systems, including

butter and margarine. However, in other food systems, phase inversion is undesirable because it

has a negative effect on their appearance, texture, stability, and taste.

Emulsions are often used as a model to evaluate the stability and the gastrointestinal fate of lipids

upon digestion (Hur, Lim, Decker, & McClements, 2011). In vitro digestion models are therefore

needed to test the efficacy of different approaches of controlling lipid digestion (e.g., by addition

of pectins or other sources of dietary fiber) under conditions that simulate the human GIT. In the

next section, we provide an overview of the major physicochemical and physiological events that

occur in each region of the human GIT for simulating the in vitro digestion of emulsified lipids.

Emphasis will be placed in the composition, structure, and dynamics of the different regions of

the GIT and their further influence on the gastrointestinal fate of emulsified lipids. This approach

utilizes a number of sequential steps (oral, gastric, and small intestine phases) to more accurately

mimic the entire digestion.

1.5. In vitro digestion model of emulsified lipids

After ingestion, emulsified lipids are subjected to a complex series of chemical and physical

changes as they pass through the mouth, stomach, small intestine, and large intestine phases,

which affect their ability to be digested and absorbed. A schematic diagram of the

physicochemical conditions in the different regions of the human GIT is shown in Figure 1.10

(Hur, Lim, Decker, & McClements, 2011; McClements & Li, 2010).

Page 32: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

23

Figure 1.10. Schematic diagram of the physicochemical conditions in the different regions of the human

gastrointestinal tract. These conditions are used for simulating the in vitro digestion of emulsified lipids

and were obtained from previous reports (Hur, Lim, Decker, & McClements, 2011; McClements & Li,

2010).

By using this GIT model, the emulsified lipids are prepared and then subjected to three sequential

steps designed to mimic particular regions of the human GIT (e.g., mouth, stomach, and small

intestine). It is important to clarify that although the colonic phase will be mentioned, this phase

was not simulated in this thesis because there is evidence suggesting that emulsified lipids are

fully digested and adsorbed within the small intestine, therefore this step can be omitted (Hur,

Lim, Decker, & McClements, 2011; Singh, Ye, & Horne, 2009).

1.5.1. Oral phase

The main function of the mouth is to ingest the foods and to convert them into a form appropriate

for swallowing. The composition, structure, and properties of lipid droplets change appreciably

Oral phase:

• pH 5 - 7

• -Amylase

• Salts

• Biopolymers (mucin)

• 5 - 60 s

Gastric phase:

• pH 1 - 3

• Pepsin

• Salts

• Biopolymers (mucin)

• Agitation

• 30 min – 4 h

Intestinal phase:

• pH 6 - 8

• Pancreatin

• Bile salts

• Agitation

• 1 - 2 h

Colonic phase:

• pH 5 - 7

• Bacterial enzymes

• Anaerobic conditions

• Agitation

• 12 – 24 h

Page 33: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

24

during mastication due to the complex physicochemical and physiological processes that occur

within the mouth (Singh, Ye, & Ferrua, 2015). An ingested food is subjected to a wide number of

physicochemical processes: mixing with saliva; changing the pH, ionic strength, and temperature;

attacking by various digestive enzymes (e.g., -amylase); interacting with biopolymers from

saliva (e.g., mucin); and breaking down after mastication (Singh, Ye, & Horne, 2009;

Vingerhoeds, Blijdenstein, Zoet, & van Aken, 2005). The most important factor affecting the

behavior of emulsified lipids in the mouth is the interaction with saliva (Vingerhoeds,

Blijdenstein, Zoet, & van Aken, 2005). Human saliva is usually around pH 5.5 to 6.0 during

fasting and around 6.5 to 7.0 after food ingestion (Brandão, Soares, Mateus, & de Freitas, 2014).

Saliva contains water (99%), minerals (<1%), and proteins (0.1 to 0.5%). The protein fraction is

complex and contains digestive enzymes (e.g., -amylase), immunoglobulins, and glycosylated

proteins (e.g., mucins) (Brandão, Soares, Mateus, & de Freitas, 2014). The mucins are proteins

capable of inducing flocculation of the lipid droplets (Vingerhoeds, Blijdenstein, Zoet, & van

Aken, 2005). Therefore, limited lipid digestion may occur during the mastication process because

of the flocculation of lipid droplets and to the lack of lingual lipases secreted within the mouth in

adults (Neyraud, Palicki, Schwartz, Nicklaus, & Feron, 2012). The material that is swallowed

after mastication is referred to as bolus.

1.5.2. Gastric phase

After the bolus is swallowed it rapidly passes through the esophagus and reaches the stomach,

where it is mixed with the gastric fluids containing gastric enzymes (e.g., gastric lipase and

pepsin), minerals, and surface-active materials, and is also subjected to mechanical forces due to

stomach motility (Pal, Abrahamsson, Schwizer, Hebbard, & Brasseur, 2003). The pH of the

human stomach has been reported to be between 1.0 to 3.0 during fasting and around 2.0 to 2.7

after food ingestion (Singh, Ye, & Horne, 2009). The high acidity of stomach plays a wide

variety of physiological roles, including activating enzymes (e.g., zymogens), hydrolysis of food

components (both enzymatic and chemical hydrolysis), and the inactivation of microorganisms

(Guerra, Etienne-Mesmin, Livrelli, Denis, Blanquet-Diot, & Alric, 2012). The digestion of

emulsified lipids begins in the stomach because of the presence of gastric lipase. Gastric lipase

Page 34: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

25

adsorbs onto the lipid droplet surfaces, where it is able to convert triacylglycerols (TAG) into

diacylglycerols (DAG), monoacylglycerols (MAG), and free fatty acids (FFA). It has been

established, however, that the gastric digestion of emulsified lipids is negligible as compared to

that occurring in the small intestine, because gastric lipases are considerably less active than

intestinal lipases (Canaan, Roussel, Verger, & Cambillau, 1999; Reis, Holmberg, Miller, Leser,

Raab, & Watzke, 2009; Reis, Holmberg, Watzke, Leser, & Miller, 2009). The FFA released in

the stomach after gastric digestion of TAG, nevertheless, may play an important role in the

subsequent digestion and absorption process of nutrients since FFA are capable to i) promote

lipid digestion by enhancing droplet disruption, ii) increase solubilization of lipid digestion

products, iii) stimulate hormone release, iv) increase the binding capacity of colipase protein, and

v) increase the activity of pancreatic lipase in the small intestine (Singh, Ye, & Ferrua, 2015;

Singh, Ye, & Horne, 2009). The partially digested and disrupted food that leaves the stomach and

reaches the small intestine is usually known as chyme.

1.5.3. Small intestine phase

The small intestine is the region in the GIT where most of the lipid digestion and absorption

processes normally occur. After reaching the small intestine phase, the chyme is mixed with

bicarbonate (HCO3⊝), bile salts, phospholipids, and enzymes secreted by the liver and the

pancreas (McClements & Li, 2010). The bicarbonate secreted into the small intestine increases

the pH from highly acidic (around pH 1.0 to 3.0) to neutral (around pH 5.8 to 6.5), where the

pancreatic enzymes work more efficiently (Wilde & Chu, 2011). Lipid hydrolysis is carried out

in the small intestine by the enzymatic action of lipases secreted by the pancreas (Singh, Ye, &

Ferrua, 2015). Pancreatic lipase plays an important role in the lipid digestion process because it is

the digestive enzyme responsible for hydrolyzing TAGs into FFAs, DAGs, and MAGs. To

catalyze this reaction, pancreatic lipase needs to adsorb onto the lipid droplet surfaces to be in

proximity to TAG (Reis, Holmberg, Miller, Leser, Raab, & Watzke, 2009). Pancreatic lipase

often does this as part of a complex with a protein known as colipase, and also with bile salts

(Erlanson-Albertsson, 1983). Usually, colipase binds to the C-terminal domain of pancreatic

lipase, increasing its hydrophobicity and facilitating its adsorption onto the lipid droplet surfaces

Page 35: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

26

(Erlanson-Albertsson, 1983). Therefore, both pancreatic lipase and colipase come together in

commercial enzymatic preparations so then, further addition of colipase is not necessary (Singh,

Ye, & Horne, 2009).

Bile salts are capable of stabilizing lipid droplets against aggregation, and forming micelles that

solubilize and transport hydrophobic molecules such as TAGs and the lipid digestion products

(Maldonado-Valderrama, Wilde, Macierzanka, & Mackie, 2011). Once the digestion process of

emulsified lipids has been carried out by pancreatic lipase, the undigested lipids, lipid digestion

products, and organic-soluble compounds (e.g., FFAs, DAGs, MAGs, cholesterol, phospholipids,

and fat-soluble vitamins) are solubilized within micelles formed by bile salts and phospholipids,

and then, they are transported to the epithelium cells for further absorption (Golding & Wooster,

2010; Singh, Ye, & Horne, 2009; Wilde & Chu, 2011).

1.5.4. Large intestine phase

The material that is not digested and absorbed within the small intestine reaches the large

intestine. The main physiological function of the large intestine is the absorption of water and

electrolytes, the fermentation of polysaccharides and proteins, the re-absorption of bile salts, and

the formation, storage, and elimination of fecal matter (Emmanuel & Butt, 2015). Any material

that is undigested in the upper GIT eventually reaches the colon. Usually, lipids are fully digested

in the stomach and small intestine. However, undigested lipids may pass through the GIT and

reach the colon (Guerra, Etienne-Mesmin, Livrelli, Denis, Blanquet-Diot, & Alric, 2012;

McClements & Li, 2010; Singh, Ye, & Horne, 2009). For example, if lipid droplets are

surrounded by an indigestible layer or trapped within an indigestible matrix (e.g., a dietary fiber

matrix), then they are not capable to be fully digested in the upper GIT (McClements, 2010). The

large intestine contains different kinds of anaerobic microorganisms, which are capable to

ferment some food components that were not digested in the upper GIT (Gibson, Probert, Loo,

Rastall, & Roberfroid, 2004). Therefore, lipids trapped within dietary fiber matrices may only be

released after they reach the large intestine by means of bacterial fermentation where they cannot

longer be digested or absorbed (McClements, 2010, 2015; Yao, Xiao, & McClements, 2014).

Page 36: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

27

1.6. Aim and outline of the thesis

In recent years, the effect of pectin consumption on a number of physiological responses such as

the regulation of body-weight and blood pressure, as well as the control of the glucose and lipid

levels in blood have been identified. However, the mechanisms by which pectin is able to exert

its physiological functions have not yet been studied. For example, little information is available

about the inhibitory capacity of pectin on the activity of the major digestive enzymes (pancreatic

lipase, -amylase, alkaline phosphatase, and protease), or about the inhibitory capacity of pectin

on the rate and extent of the mass transfer process of nutrients (e.g., monosaccharides, amino

acids, and lipids). Furthermore, few attempts on the evaluation of the effect of pectin properties

(e.g., molecular weight and methoxylation degree) on the gastrointestinal fate of emulsified lipids

have been done. Finally, little is known about the molecular interactions between pectin and the

major gastrointestinal components governing the digestion process of emulsified lipids.

Therefore, with the objective to evaluate the mechanisms by which pectin is able to exert its

physiological functions, the aim of this thesis is to obtain fundamental understanding of the

influence of pectin properties (methoxylation degree) on the activity of the major digestive

enzymes (pancreatic lipase, -amylase, alkaline phosphatase, and protease), and the rate and

extent of the mass transfer process of nutrients (monosaccharides, amino acids, and lipids). This

thesis is also aimed at evaluating the incidence of pectin properties on the gastrointestinal fate of

emulsified lipids by using a simulated in vitro GIT model, and to evaluate both the nature and

magnitude of the molecular interactions between pectin and the compounds related to the

digestion of emulsified lipids.

In chapter 2, the evaluation of the factors affecting the activity of the main digestive enzymes

[pancreatic lipase, -amylase, alkaline phosphatase, and protease (chymotrypsin)] is described.

The results obtained in this chapter are used in chapter 3, where the influence of pectin

properties (methoxylation degree) on the enzymatic activity of these enzymes is evaluated.

Because pancreatic lipase was found to be the enzyme affected the most by pectin, subsequent

chapters are focused on evaluating the effect of pectin on the gastrointestinal fate of lipids. In

Page 37: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

28

chapter 4, the inhibitory effect of pectin on the rate and extent of the digestion of emulsified

lipids is evaluated by using a simulated in vitro GIT model, and the results are compared with

those obtained with other sources of dietary fiber (chitosan, and methyl cellulose). In addition,

chapter 5 describes the influence of pectin properties (molecular weight and methoxylation

degree) on the gastrointestinal fate of emulsified lipids by using a simulated in vitro GIT model.

To give a deeper insight on the mechanism governing the effect of pectin on the digestion of

emulsified lipids, chapter 6 describes both the nature and magnitude of the molecular

interactions between pectin and the major components of the GIT. Considering that pectin is

known for increasing the viscosity of the digesta and it is able to provide a restriction of the

nutrient mobility, chapter 7 describes the effect of pectin properties (methoxylation degree) on

the rate and extent of the diffusion process of the most important nutrients (monosaccharides,

amino acids, and an emulsion). Finally, in chapter 8, a general discussion is presented and

implications of the findings are elaborated.

References

ACS. (1944). Report of the Committee for the Revision of the Nomenclature of Pectic Substances.

Chemical & Engineering News Archive, 22(2), 105-106.

Basaran, T. K., Demetriades, K., & McClements, D. J. (1998). Ultrasonic imaging of gravitational

separation in emulsions. Colloids and Surfaces A: Physicochemical and Engineering Aspects,

136(1–2), 169-181.

Bolívar-Mejía, A., & Vesga-Angarita, B. E. (2013). Burden of Cardiovascular Disease in Colombia.

Brandão, E., Soares, S., Mateus, N., & de Freitas, V. (2014). Human saliva protein profile: Influence of

food ingestion. Food Research International, 64, 508-513.

Brownlee, I. A. (2011). The physiological roles of dietary fibre. Food Hydrocolloids, 25(2), 238-250.

Burton-Freeman, B. (2000). Dietary Fiber and Energy Regulation. The Journal of Nutrition, 130(2), 272.

Caffall, K. H., & Mohnen, D. (2009). The structure, function, and biosynthesis of plant cell wall pectic

polysaccharides. Carbohydrate Research, 344(14), 1879-1900.

Canaan, S., Roussel, A., Verger, R., & Cambillau, C. (1999). Gastric lipase: crystal structure and activity.

Biochimica et Biophysica Acta (BBA) - Molecular and Cell Biology of Lipids, 1441(2–3), 197-

204.

Coupland, J. N., & McClements, D. J. (1996). Lipid oxidation in food emulsions. Trends in Food Science

& Technology, 7(3), 83-91.

Cummings, J. H., Mann, J. I., Nishida, C., & Vorster, H. H. (2009). Dietary fibre: an agreed definition.

The Lancet, 373(9661), 365-366.

Cummings, J. H., & Stephen, A. M. (2007). Carbohydrate terminology and classification. Eur J Clin Nutr,

61(S1), S5-S18.

Page 38: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

29

de Leeuw, J. A., Jongbloed, A. W., & Verstegen, M. W. A. (2004). Dietary Fiber Stabilizes Blood

Glucose and Insulin Levels and Reduces Physical Activity in Sows (Sus scrofa). The Journal of

Nutrition, 134(6), 1481-1486.

Degner, B. M., Chung, C., Schlegel, V., Hutkins, R., & McClements, D. J. (2014). Factors Influencing the

Freeze-Thaw Stability of Emulsion-Based Foods. Comprehensive Reviews in Food Science and

Food Safety, 13(2), 98-113.

DeVries, J. W. (2003). On defining dietary fibre. Proceedings of the Nutrition Society, 62(01), 37-43.

Dickinson, E. (2009). Hydrocolloids as emulsifiers and emulsion stabilizers. Food Hydrocolloids, 23(6),

1473-1482.

Dickinson, E. (2010). Flocculation of protein-stabilized oil-in-water emulsions. Colloids and Surfaces B:

Biointerfaces, 81(1), 130-140.

Dickinson, E. (2015). Structuring of colloidal particles at interfaces and the relationship to food emulsion

and foam stability. Journal of Colloid and Interface Science, 449, 38-45.

Dikeman, C. L., & Fahey, G. C. (2006). Viscosity as Related to Dietary Fiber: A Review. Critical Reviews

in Food Science and Nutrition, 46(8), 649-663.

Dunaif, G., & Schneeman, B. O. (1981). The effect of dietary fiber on human pancreatic enzyme activity

in vitro. The American Journal of Clinical Nutrition, 34(6), 1034-1035.

Elwell, M. W., Roberts, R. F., & Coupland, J. N. (2004). Effect of homogenization and surfactant type on

the exchange of oil between emulsion droplets. Food Hydrocolloids, 18(3), 413-418.

Emmanuel, A., & Butt, S. (2015). Small intestine and colon motility. Medicine, 43(5), 271-275.

Erlanson-Albertsson, C. (1983). The interaction between pancreatic lipase and colipase: a protein-protein

interaction regulated by a lipid. FEBS Letters, 162(2), 225-229.

Fruk, G., Cmelik, Z., Jemric, T., Hribar, J., & Vidrih, R. (2014). Pectin role in woolliness development in

peaches and nectarines: A review. Scientia Horticulturae, 180, 1-5.

Fuentes-Zaragoza, E., Riquelme-Navarrete, M. J., Sánchez-Zapata, E., & Pérez-Álvarez, J. A. (2010).

Resistant starch as functional ingredient: A review. Food Research International, 43(4), 931-942.

Galisteo, M., Duarte, J., & Zarzuelo, A. (2008). Effects of dietary fibers on disturbances clustered in the

metabolic syndrome. The Journal of Nutritional Biochemistry, 19(2), 71-84.

Gibson, G. R., Probert, H. M., Loo, J. V., Rastall, R. A., & Roberfroid, M. B. (2004). Dietary modulation

of the human colonic microbiota: updating the concept of prebiotics. Nutrition Research Reviews,

17(02), 259-275.

Given, P. S. (2009). Encapsulation of Flavors in Emulsions for Beverages. Current Opinion in Colloid &

Interface Science, 14(1), 43-47.

Golding, M., & Wooster, T. J. (2010). The influence of emulsion structure and stability on lipid digestion.

Current Opinion in Colloid & Interface Science, 15(1–2), 90-101.

Guarner, F., & Malagelada, J.-R. (2003). Gut flora in health and disease. The Lancet, 361(9356), 512-519.

Guerra, A., Etienne-Mesmin, L., Livrelli, V., Denis, S., Blanquet-Diot, S., & Alric, M. (2012). Relevance

and challenges in modeling human gastric and small intestinal digestion. Trends in Biotechnology,

30(11), 591-600.

Guzey, D., & McClements, D. J. (2006). Formation, stability and properties of multilayer emulsions for

application in the food industry. Advances in Colloid and Interface Science, 128–130, 227-248.

Hamaker, B. R., & Tuncil, Y. E. (2014). A Perspective on the Complexity of Dietary Fiber Structures and

Their Potential Effect on the Gut Microbiota. Journal of Molecular Biology, 426(23), 3838-3850.

Haralampu, S. G. (2000). Resistant starch—a review of the physical properties and biological impact of

RS3. Carbohydrate Polymers, 41(3), 285-292.

Harris, S. S., & Pijls, L. (2013). Dietary fibre: refining a definition. The Lancet, 374(9683), 28.

Hollmann, J., Themeier, H., Neese, U., & Lindhauer, M. G. (2013). Dietary fibre fractions in cereal foods

measured by a new integrated AOAC method. Food Chemistry, 140(3), 586-589.

Page 39: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

30

Howarth, N. C., Saltzman, E., & Roberts, S. B. (2001). Dietary Fiber and Weight Regulation. Nutrition

Reviews, 59(5), 129-139.

Huang, X., Kakuda, Y., & Cui, W. (2001). Hydrocolloids in emulsions: particle size distribution and

interfacial activity. Food Hydrocolloids, 15(4–6), 533-542.

Hur, S. J., Lim, B. O., Decker, E. A., & McClements, D. J. (2011). In vitro human digestion models for

food applications. Food Chemistry, 125(1), 1-12.

Ichikawa, T., Dohda, T., & Nakajima, Y. (2006). Stability of oil-in-water emulsion with mobile surface

charge. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 279(1–3), 128-141.

Ishii, T., & Matsunaga, T. (2001). Pectic polysaccharide rhamnogalacturonan II is covalently linked to

homogalacturonan. Phytochemistry, 57(6), 969-974.

Izydorczyk, M. S., Macri, L. J., & MacGregor, A. W. (1998). Structure and physicochemical properties of

barley non-starch polysaccharides — I. Water-extractable β-glucans and arabinoxylans.

Carbohydrate Polymers, 35(3–4), 249-258.

Janssen, P., Vanden Berghe, P., Verschueren, S., Lehmann, A., Depoortere, I., & Tack, J. (2011). Review

article: the role of gastric motility in the control of food intake. Alimentary Pharmacology &

Therapeutics, 33(8), 880-894.

Jenkins, D., Wolever, T., Rao, A. V., Hegele, R. A., Mitchell, S. J., Ransom, T., Boctor, D. L., Spadafora,

P. J., Jenkins, A. L., Mehling, C., Relle, L. K., Connelly, P. W., Story, J. A., Furumoto, E. J.,

Corey, P., & Wursch, P. (1993). Effect on Blood Lipids of Very High Intakes of Fiber in Diets

Low in Saturated Fat and Cholesterol. New England Journal of Medicine, 329(1), 21-26.

Kabalnov, A. (1998). Thermodynamic and theoretical aspects of emulsions and their stability. Current

Opinion in Colloid & Interface Science, 3(3), 270-275.

Kaneko, S., Ishii, T., & Matsunaga, T. (1997). A boron-rhamnogalacturonan-II complex from bamboo

shoot cell walls. Phytochemistry, 44(2), 243-248.

Khodaei, N., & Karboune, S. (2013). Extraction and structural characterisation of rhamnogalacturonan I-

type pectic polysaccharides from potato cell wall. Food Chemistry, 139(1–4), 617-623.

Kizling, J., Kronberg, B., & Eriksson, J. C. (2006). On the formation and stability of high internal phase

O/W emulsions. Advances in Colloid and Interface Science, 123–126, 295-302.

Krebs, T., Schroën, C. G. P. H., & Boom, R. M. (2013). Coalescence kinetics of oil-in-water emulsions

studied with microfluidics. Fuel, 106, 327-334.

Kritchevsky, D., & Story, J. (1982). Dietary Fiber and Cancer. In J. Vitale & S. Broitman (Eds.),

Advances in Human Clinical Nutrition, (pp. 175-188): Springer Netherlands.

Kumar, V., Sinha, A. K., Makkar, H. P. S., de Boeck, G., & Becker, K. (2012). Dietary Roles of Non-

Starch Polysachharides in Human Nutrition: A Review. Critical Reviews in Food Science and

Nutrition, 52(10), 899-935.

Leal-Calderon, F., Thivilliers, F., & Schmitt, V. (2007). Structured emulsions. Current Opinion in Colloid

& Interface Science, 12(4–5), 206-212.

Leunissen, M. E., van Blaaderen, A., Hollingsworth, A. D., Sullivan, M. T., & Chaikin, P. M. (2007).

Electrostatics at the oil–water interface, stability, and order in emulsions and colloids.

Proceedings of the National Academy of Sciences, 104(8), 2585-2590.

Maldonado-Valderrama, J., Wilde, P., Macierzanka, A., & Mackie, A. (2011). The role of bile salts in

digestion. Advances in Colloid and Interface Science, 165(1), 36-46.

Mann, J. I., & Cummings, J. H. (2009). Possible implications for health of the different definitions of

dietary fibre. Nutrition, Metabolism and Cardiovascular Diseases, 19(3), 226-229.

McClements, D. J. (2004). Protein-stabilized emulsions. Current Opinion in Colloid & Interface Science,

9(5), 305-313.

McClements, D. J. (2010). Emulsion Design to Improve the Delivery of Functional Lipophilic

Components. Annual Review of Food Science and Technology, 1(1), 241-269.

Page 40: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

31

McClements, D. J. (2015). Encapsulation, protection, and release of hydrophilic active components:

Potential and limitations of colloidal delivery systems. Advances in Colloid and Interface Science,

219, 27-53.

McClements, D. J., & Li, Y. (2010). Review of in vitro digestion models for rapid screening of emulsion-

based systems. Food & Function, 1(1), 32-59.

Mohnen, D. (2008). Pectin structure and biosynthesis. Current Opinion in Plant Biology, 11(3), 266-277.

Munarin, F., Tanzi, M. C., & Petrini, P. (2012). Advances in biomedical applications of pectin gels.

International Journal of Biological Macromolecules, 51(4), 681-689.

Mussatto, S. I., & Mancilha, I. M. (2007). Non-digestible oligosaccharides: A review. Carbohydrate

Polymers, 68(3), 587-597.

Neyraud, E., Palicki, O., Schwartz, C., Nicklaus, S., & Feron, G. (2012). Variability of human saliva

composition: Possible relationships with fat perception and liking. Archives of Oral Biology,

57(5), 556-566.

Oosterveld, A., Pol, I. E., Beldman, G., & Voragen, A. G. J. (2001). Isolation of feruloylated arabinans

and rhamnogalacturonans from sugar beet pulp and their gel forming ability by oxidative cross-

linking. Carbohydrate Polymers, 44(1), 9-17.

Ozturk, B., & McClements, D. J. (2016). Progress in natural emulsifiers for utilization in food emulsions.

Current Opinion in Food Science, 7, 1-6.

Pal, A., Abrahamsson, B., Schwizer, W., Hebbard, G. S., & Brasseur, J. G. (2003). Application of a virtual

stomach to evaluate gastric mixing and breakdown of solid food. Gastroenterology, 124(4,

Supplement 1), A673-A674.

Perazzo, A., Preziosi, V., & Guido, S. (2015). Phase inversion emulsification: Current understanding and

applications. Advances in Colloid and Interface Science, 222, 581-599.

Pérez, S., Mazeau, K., & Hervé du Penhoat, C. (2000). The three-dimensional structures of the pectic

polysaccharides. Plant Physiology and Biochemistry, 38(1–2), 37-55.

Pérez, S., Rodríguez-Carvajal, M. A., & Doco, T. (2003). A complex plant cell wall polysaccharide:

rhamnogalacturonan II. A structure in quest of a function. Biochimie, 85(1–2), 109-121.

Phillips, G. O., & Cui, S. W. (2011). An introduction: Evolution and finalisation of the regulatory

definition of dietary fibre. Food Hydrocolloids, 25(2), 139-143.

Post, R. E., Mainous, A. G., King, D. E., & Simpson, K. N. (2012). Dietary Fiber for the Treatment of

Type 2 Diabetes Mellitus: A Meta-Analysis. The Journal of the American Board of Family

Medicine, 25(1), 16-23.

Raigond, P., Ezekiel, R., & Raigond, B. (2015). Resistant starch in food: a review. Journal of the Science

of Food and Agriculture, 95(10), 1968-1978.

Ramulu, P., & Udayasekhara Rao, P. (2003). Total, insoluble and soluble dietary fiber contents of Indian

fruits. Journal of Food Composition and Analysis, 16(6), 677-685.

Rao, J., & McClements, D. J. (2010). Stabilization of Phase Inversion Temperature Nanoemulsions by

Surfactant Displacement. Journal of Agricultural and Food Chemistry, 58(11), 7059-7066.

Reis, P., Holmberg, K., Miller, R., Leser, M. E., Raab, T., & Watzke, H. J. (2009). Lipase reaction at

interfaces as self-limiting processes. Comptes Rendus Chimie, 12(1–2), 163-170.

Reis, P., Holmberg, K., Watzke, H., Leser, M. E., & Miller, R. (2009). Lipases at interfaces: A review.

Advances in Colloid and Interface Science, 147–148, 237-250.

Ridley, B. L., O'Neill, M. A., & Mohnen, D. (2001). Pectins: structure, biosynthesis, and

oligogalacturonide-related signaling. Phytochemistry, 57(6), 929-967.

Robins, M. M. (2000). Emulsions — creaming phenomena. Current Opinion in Colloid & Interface

Science, 5(5–6), 265-272.

Robins, M. M., Watson, A. D., & Wilde, P. J. (2002). Emulsions—creaming and rheology. Current

Opinion in Colloid & Interface Science, 7(5–6), 419-425.

Page 41: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

32

Rodríguez-Abreu, C., & Lazzari, M. (2008). Emulsions with structured continuous phases. Current

Opinion in Colloid & Interface Science, 13(4), 198-205.

Saiki, Y., Horn, R. G., & Prestidge, C. A. (2008). Droplet structure instability in concentrated emulsions.

Journal of Colloid and Interface Science, 320(2), 569-574.

Sajilata, M. G., Singhal, R. S., & Kulkarni, P. R. (2006). Resistant Starch–A Review. Comprehensive

Reviews in Food Science and Food Safety, 5(1), 1-17.

Sasaki, T., Kohyama, K., & Yasui, T. (2004). Effect of water-soluble and insoluble non-starch

polysaccharides isolated from wheat flour on the rheological properties of wheat starch gel.

Carbohydrate Polymers, 57(4), 451-458.

Silva, I. R., Jers, C., Meyer, A. S., & Mikkelsen, J. D. (2016). Rhamnogalacturonan I modifying enzymes:

an update. New Biotechnology, 33(1), 41-54.

Sims, I. M., Ryan, J. L. J., & Kim, S. H. (2014). In vitro fermentation of prebiotic oligosaccharides by

Bifidobacterium lactis HN019 and Lactobacillus spp. Anaerobe, 25, 11-17.

Singh, H., Ye, A., & Ferrua, M. J. (2015). Aspects of food structures in the digestive tract. Current

Opinion in Food Science, 3, 85-93.

Singh, H., Ye, A., & Horne, D. (2009). Structuring food emulsions in the gastrointestinal tract to modify

lipid digestion. Progress in Lipid Research, 48(2), 92-100.

Slavin, J. L. (2005). Dietary fiber and body weight. Nutrition, 21(3), 411-418.

Snarski, S. R., & Dunn, P. F. (1991). Experiments characterizing the interaction between two sprays of

electrically charged liquid droplets. Experiments in Fluids, 11(4), 268-278.

Starov, V. M., & Zhdanov, V. G. (2003). Viscosity of emulsions: influence of flocculation. Journal of

Colloid and Interface Science, 258(2), 404-414.

Streppel, M. T., Arends, L. R., & van’tVeer, P. (2005). Dietary fiber and blood pressure. ACC Current

Journal Review, 14(4), 19.

Sun, Y., & O’Riordan, M. X. D. (2013). Regulation of Bacterial Pathogenesis by Intestinal Short-Chain

Fatty Acids. In S. Sima & M. G. Geoffrey (Eds.), Advances in Applied Microbiology, vol.

Volume 85 (pp. 93-118): Academic Press.

Tcholakova, S., Denkov, N. D., Ivanov, I. B., & Campbell, B. (2006). Coalescence stability of emulsions

containing globular milk proteins. Advances in Colloid and Interface Science, 123–126, 259-293.

Threapleton, D. E., Greenwood, D. C., Evans, C. E. L., Cleghorn, C. L., Nykjaer, C., Woodhead, C., Cade,

J. E., Gale, C. P., & Burley, V. J. (2013). Dietary fibre intake and risk of cardiovascular disease:

systematic review and meta-analysis. BMJ, 347.

Torsdottir, I., Alpsten, M., Holm, G., Sandberg, A. S., & Tölli, J. (1991). A small dose of soluble alginate-

fiber affects postprandial glycemia and gastric emptying in humans with diabetes. The Journal of

Nutrition, 121(6), 795-799.

Turner, N. D., & Lupton, J. R. (2011). Dietary Fiber. Advances in Nutrition: An International Review

Journal, 2(2), 151-152.

Vingerhoeds, M. H., Blijdenstein, T. B. J., Zoet, F. D., & van Aken, G. A. (2005). Emulsion flocculation

induced by saliva and mucin. Food Hydrocolloids, 19(5), 915-922.

Wanders, A. J., Jonathan, M. C., van den Borne, J. J. G. C., Mars, M., Schols, H. A., Feskens, E. J. M., &

de Graaf, C. (2013). The effects of bulking, viscous and gel-forming dietary fibres on satiation.

British Journal of Nutrition, 109(07), 1330-1337.

Whitcombe, A. J., O'Neill, M. A., Steffan, W., Albersheim, P., & Darvill, A. G. (1995). Structural

characterization of the pectic polysaccharide, rhamnogalacturonan-II. Carbohydrate Research,

271(1), 15-29.

Wilde, P. J., & Chu, B. S. (2011). Interfacial and colloidal aspects of lipid digestion. Advances in Colloid

and Interface Science, 165(1), 14-22.

Worth, H. G. J. (1967). The Chemistry and Biochemistry of Pectic Substances. Chemical Reviews, 67(4),

465-473.

Page 42: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 1

33

Yan, Y., & Masliyah, J. H. (1993). Sedimentation of solid particles in oil-in-water emulsions.

International Journal of Multiphase Flow, 19(5), 875-886.

Yang, Y., Marshall-Breton, C., Leser, M. E., Sher, A. A., & McClements, D. J. (2012). Fabrication of

ultrafine edible emulsions: Comparison of high-energy and low-energy homogenization methods.

Food Hydrocolloids, 29(2), 398-406.

Yang, Y., Zhang, G., Hong, Y., Gu, Z., & Fang, F. (2013). Calcium cation triggers and accelerates the

gelation of high methoxy pectin. Food Hydrocolloids, 32(2), 228-234.

Yao, M., Xiao, H., & McClements, D. J. (2014). Delivery of Lipophilic Bioactives: Assembly,

Disassembly, and Reassembly of Lipid Nanoparticles. Annual Review of Food Science and

Technology, 5(1), 53-81.

Yapo, B. M. (2011). Pectic substances: From simple pectic polysaccharides to complex pectins—A new

hypothetical model. Carbohydrate Polymers, 86(2), 373-385.

You, X., Xie, C., Liu, K., & Gu, Z. (2010). Isolation of non-starch polysaccharides from bulb of tiger lily

(Lilium lancifolium Thunb.) with fermentation of Saccharomyces cerevisiae. Carbohydrate

Polymers, 81(1), 35-40.

Yuan, J. Y., Smeele, R. J., Harington, K., van Loon, F., Wanders, A., & Venn, B. (2014). The effects of

functional fiber on postprandial glycemia, energy intake, satiety, palatability and gastrointestinal

wellbeing: a randomized crossover trial. Nutrition Journal, 13(1), 76.

Zhang, W., Li, D., Liu, L., Zang, J., Duan, Q., Yang, W., & Zhang, L. (2013). The effects of dietary fiber

level on nutrient digestibility in growing pigs. Journal of Animal Science and Biotechnology, 4(1),

17-17.

Page 43: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

Optimization of the reaction conditions affecting

the activity of digestive enzymes

Page 44: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

35

Abstract

Enzymatic activities were evaluated spectrophotometrically by using the artificial chromogenic

substrates p-nitrophenyl palmitate, 2-chloro-p-nitrophenyl--D-maltotrioside, p-nitrophenyl

phosphate, and p-nitrophenyl acetate for pancreatic lipase, -amylase, alkaline phosphatase, and

protease, respectively. Artificial chromogenic substrates were used because they allow obtaining

fast, sensitive, and reproducible results. Both temperature (37 °C) and pH (7.0) of each of the

reaction mixtures were fixed to ensure the optimum performance of digestive enzymes. Buffer

and cofactor types and buffer, cofactor, and enzyme concentrations were evaluated as the reaction

conditions affecting the activity of the digestive enzymes. Pancreatic lipase and alkaline

phosphatase were more active in Tris-HCl buffer (at concentrations of 50 and 20 mM,

respectively) as compared to phosphate buffer. Conversely, -amylase and protease were more

active in phosphate buffer (at concentrations of 20 and 50 mM, respectively) as compared to Tris-

HCl buffer. -Amylase, alkaline phosphatase, and protease required the presence of NaCl as

cofactor (at concentrations of 10, 15, and 20 mM, respectively), whereas pancreatic lipase

required the presence of both NaCl and CaCl2 (at concentrations of 150 and 1.0 mM,

respectively). The optimum enzyme concentrations were found to be 900, 10, 0.8, and 3 U mL-1

for pancreatic lipase, -amylase, alkaline phosphatase, and protease, respectively. Finally, the

optimization process of the reaction conditions allowed improving the activities of all of the

digestive enzymes during each step. However, the optimization of the reaction conditions had no

a significant effect on the reproducibility of the measurements.

Keywords: Digestive enzymes, buffer, cofactor, specific activity, physiological conditions.

Page 45: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

36

2.1. Introduction

The chemical reactions occurring in living organisms are controlled by enzymes

(Martínez Cuesta, Rahman, Furnham, & Thornton, 2015). Enzymes are globular proteins with

complex three-dimensional structures, responsible for their unique catalytic functions (Jacobson,

Kalyanaraman, Zhao, & Tian, 2014). Regarding the complex structures of enzymes, it is

reasonable to expect that many environmental parameters are capable to affect their three-

dimensional conformation and their subsequent catalytic activities (Daniel & Danson, 2013). The

activity of an enzyme can be largely affected by several environmental conditions such as

temperature, pH, types of buffer and cofactor, and concentrations of buffer, cofactor, substrate,

and enzyme (Iyer & Ananthanarayan, 2008). Enzymes display their highest activities at their

optimum conditions; deviations from these conditions may cause a reduction of the enzymatic

activity, depending on the degree of the deviation (Vendruscolo, 2002). Moderate deviations

produce small activity decrease. However, severe deviations may lead to a complete loss of the

enzymatic activity. Consequently, it is important to establish the optimum parameters for

measuring the activity of enzymes to ensure their optimal performance.

The enzymatic reactions can be carried out by using either native or artificial substrates. The use

of native substrates is preferred because of the biological relevance of the results. However, the

reproducibility obtained with native substrates is sometimes unacceptable, especially, when the

substrates have a high structural complexity (Svendsen, Blombäck, Blombäck, & Olsson, 1972).

Instead, it is possible to use artificial substrates to obtain faster and reliable results as compared

to native substrates (Bisswanger, 2014). A wide variety of artificial chromogenic substrates are

available for the rapid and reliable in vitro assaying of the activity of several enzymes (Goddard

& Reymond, 2004). Artificial substrates are structurally similar to their analogous native

substrates, but they are conjugated to a chromogenic leaving group. After enzymatic reaction

takes place, the chromogenic leaving group is liberated and it can be conveniently monitored by

changes in absorbance.

Page 46: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

37

It has been established that enzymes are less specific (ability of the enzyme to select the

substrate) but more selective (ability of the substrate to bind to the enzyme) to artificial substrates

(Svendsen, Blombäck, Blombäck, & Olsson, 1972). Therefore, the use of artificial substrates has

some advantages and disadvantages. In the one hand, the high selectivity existing between

enzymes and artificial substrates allows the enzymatic assay to be more sensitive, faster, and

reproducible as compared to native substrates. In the other hand, the low specificity of enzymes

by artificial substrates may lead to collateral enzymatic reactions if other enzymes are present

(Wenger, Sattler, Clark, & Wharton, 1976). However, because the enzymes used in this study are

commercial and relatively pure, interferences can be considered as negligible. Furthermore, extra

development efforts for adjusting the experimental conditions when using artificial substrates are

required because the experimental conditions reported in literature for native substrates may

differ considerably to those required for artificial substrates (Nagaki, Kimura, Kimura, Maki,

Goto, Nishino, et al., 2001).

In this chapter the reaction conditions affecting the activity of some hydrolytic digestive enzymes

were studied. Pancreatic lipase, -amylase, and protease were selected because they are the main

enzymes involved in the digestion of lipids, carbohydrates, and proteins, respectively. Although

alkaline phosphatase is an enzyme not directly involved in the digestive processes, it was also

selected because it participates in the dephosphorylation of nutrients, which is a necessary

process for their further digestion (Lallès, 2014). Enzymatic activities were evaluated

spectrophotometrically by using the artificial chromogenic substrates p-nitrophenyl palmitate

(pNPPA), 2-chloro-p-nitrophenyl--D-maltotrioside (G3CNP), p-nitrophenyl phosphate (pNPP),

and p-nitrophenyl acetate (pNPA) for pancreatic lipase, -amylase, alkaline phosphatase, and

protease, respectively (Figure 2.1). Buffer and cofactor types and buffer, cofactor, and enzyme

concentrations were evaluated as the reaction conditions affecting the activities of the

aforementioned digestive enzymes. Because of the human physiological conditions were taken

into account, both temperature (37 °C) and pH (7.0) were fixed to ensure the optimum

performance of digestive enzymes. The best reaction conditions obtained in this study for the

measurement of the activity of digestive enzymes will be used in further studies (Chapter 3) to

evaluate the effect of pectin on the activity of these enzymes.

Page 47: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

38

Figure 2.1. Chemical structures of the artificial chromogenic substrates 2-chloro-p-nitrophenyl--D-

maltotrioside (G3CNP), p-nitrophenyl acetate (pNPA), p-nitrophenyl palmitate (pNPPA), and p-

nitrophenyl phosphate (pNPP) for -amylase, protease, pancreatic lipase, and alkaline phosphatase,

respectively. Arrows indicate the sites of bond cleavage by the enzymes. Chromogenic leaving groups are

represented in red: 2-chloro-p-nitrophenol (CNP) and p-nitrophenol (pNP).

2.2. Materials and methods

2.2.1. Chemicals

The hydrolytic enzymes porcine pancreatic -amylase, from Sus scrofa (16 U mg-1

type VI-B,

E.C. 3.2.1.1); porcine pancreatic lipase, from Sus scrofa (100 U mg-1

type II, E.C. 3.1.1.3);

bovine pancreatic protease (chymotrypsin), from Bos taurus (5 U mg-1

type I, E.C. 3.4.21.1); and

bovine intestinal mucosa alkaline phosphatase, from Bos taurus (10 U mg-1

type I, E.C. 3.1.3.1);

the artificial substrates G3CNP, pNPPA, pNPA, and pNPP; the reaction products p-nitrophenol

(pNP) and 2-chloro-p-nitrophenol (CNP); and the protein determination reagents brilliant blue G-

250 (Coomassie Blue) and bovine serum albumin (BSA) were purchased from Sigma-Aldrich

Chemical Company (St. Louis, MO, USA). Ox bile extract with cholic acid content higher than

55% (w/w) was purchased from MP Biomedicals (Solon, OH, USA). Other chemicals were

purchased from Merck KGaA (Darmstadt, Germany).

( )14

G3CNP pNPA pNPPA pNPP

NO2

O

O

CH3

NO2

O

O

CH3

NO2

O

P

O

OH OH

O

OO

OH

OH

OH

OH

OO

OH

OH

OH

OH

O

OH

OH

OH

OH

NO2

Cl

NO2

O

O

CH3

NO2

O

O

CH3

NO2

O

P

O

OH OH

O

OO

OH

OH

OH

OH

OO

OH

OH

OH

OH

O

OH

OH

OH

OH

NO2

Cl

pNPCNP

O

OOH

OH

OH

OH

O

OOH

OH

OH

OH

O

OOH

OH

OH

OH

NO2

Cl

NO2

O

O

CH3

pNP pNP

Page 48: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

39

2.2.2. Reaction conditions affecting the activity of digestive enzymes

The reaction conditions affecting the enzymatic activities for each of the tested enzymes were

adjusted by using a stepwise design: the best result obtained for one parameter was fixed and then

used to evaluate the next one. Because of the variability in the reaction conditions that can be

found for each enzyme, depending on the nature of the substrate and the source of the enzyme, a

range of experimental conditions typically found in literature were selected to be evaluated for

each enzyme. The reaction mixtures for all of the tested enzymes were prepared in the proper

buffer solution at pH 7.0 (depending on the buffer that was evaluated). All buffer solutions

evaluated for each of the tested enzymes contained 3.5 mg mL-1

bile acid extract (equivalent to

5.0 mM cholic acid in the total mixture). It is known that pancreatic lipase is the only enzyme

which is dependent on bile salts (Reis, Holmberg, Watzke, Leser, & Miller, 2009); however, bile

salts were also added to the reaction mixtures of -amylase, alkaline phosphatase, and protease

because bile salts are capable to modify the chemical environment of these enzymes and affect

their further stabilities (Robic, Linscott, Aseem, Humphreys, & McCartha, 2011). All reaction

mixtures for each of the tested enzymes were incubated at 37 °C and the absorbance of the

product was measured every 5 s throughout 120 s. The product obtained for pancreatic lipase,

alkaline phosphatase, and protease was pNP and the absorbance was recorded at 415 nm, whereas

the product obtained for -amylase was CNP and the absorbance was recorded at 405 nm

(Figure 2.1).

2.2.2.1. Pancreatic lipase

The conditions for measuring the activity of pancreatic lipase are based on those reported by

(Tsujita, Takaichi, Takaku, Sawai, Yoshida, & Hiraki, 2007). Buffer type [Tris-HCl and

phosphate buffer solution (PBS)], buffer concentration (Tris-HCl at concentrations ranging from

20 to 100 mM), cofactor type (NaCl, CaCl2, and a mixture of both NaCl and CaCl2), cofactor

concentration (NaCl at concentrations ranging from 30 to 270 mM, and CaCl2 at concentrations

ranging from 0.3 to 2.7 mM), and enzyme concentration (pancreatic lipase at concentrations

ranging from 300 to 1500 U mL-1

) were evaluated in the mentioned order. For each of the

Page 49: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

40

enzymatic assays, the substrate was kept constant at a concentration of 4.0 mM pNPPA. All

reaction mixtures contained 100 g mL-1

Tween-20 to improve the solubility of the substrate.

2.2.2.2. -Amylase

The conditions for measuring the activity of -amylase are based on those reported by

(Morishita, Iinuma, Nakashima, Majima, Mizuguchi, & Kawamura, 2000). Buffer type (Tris-HCl

and PBS), buffer concentration (PBS at concentrations ranging from 5 to 45 mM), cofactor type

(NaCl, CaCl2, and a mixture of both NaCl and CaCl2), cofactor concentration (NaCl at

concentrations ranging from 3 to 23 mM), and enzyme concentration (-amylase at

concentrations ranging from 3 to 15 U mL-1

) were evaluated in the mentioned order. For each of

the enzymatic assays, the substrate was kept constant at a concentration of 0.4 mM G3CNP. All

reaction mixtures contained 15% (v/v) glycerol required to improve the stability of the enzyme.

2.2.2.3. Alkaline phosphatase

The conditions for measuring the activity of alkaline phosphatase are based on those reported by

(Chaudhuri, Chatterjee, Venu-Babu, Ramasamy, & Thilagaraj, 2013). Buffer type (Tris-HCl and

PBS), buffer concentration (Tris-HCl at concentrations ranging from 5 to 45 mM), cofactor type

(NaCl, CaCl2, and a mixture of both NaCl and CaCl2), cofactor concentration (NaCl at

concentrations ranging from 3 to 23 mM), and enzyme concentration (alkaline phosphatase at

concentrations ranging from 0.3 to 1.3 U mL-1

) were evaluated in the mentioned order. For each

of the enzymatic assays, the substrate was kept constant at a concentration of 0.4 mM pNPP.

2.2.2.4. Protease

The conditions for measuring the activity of protease are based on those reported by (Verma &

Ghosh, 2010). Buffer type (Tris-HCl and PBS), buffer concentration (PBS at concentrations

ranging from 10 to 60 mM), cofactor type (NaCl, CaCl2, and a mixture of both NaCl and CaCl2),

cofactor concentration (NaCl at concentrations ranging from 5 to 30 mM), and enzyme

Page 50: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

41

concentration (protease at concentrations ranging from 1 to 5 U mL-1

) were evaluated in the

mentioned order. For each of the enzymatic assays, the substrate was kept constant at a

concentration of 1.0 mM pNPA.

2.2.2.5. Determination of the specific enzymatic activities

To determine the specific enzymatic activities, the absorbance obtained from the enzymatic

activity experiments was plotted versus time. From the straight lines, the slope (expressed as

absorbance units min-1

) was calculated and then interpolated in the calibration lines for each

reaction product to express it as μmol product min-1

. The calibration lines were obtained using

pNP at concentrations ranging from 5 to 50 M (7 data points, r2=0.999) for lipase, alkaline

phosphatase, and protease; and using CNP at concentrations ranging from 5 to 60 M (7 data

points, r2=0.998) for -amylase. The protein concentration of each working enzyme solution was

determined by using the modified Bradford method (Zor & Selinger, 1996), with BSA as the

standard. The specific enzymatic activities of pancreatic lipase, alkaline phosphatase, and

protease were then expressed as mol pNP min-1

mg-1

protein, while the specific enzyme activity

of -amylase was expressed as mol CNP min-1

mg-1

protein.

2.2.3. Data analysis

All measurements were conducted with three analytical and three technical (instrumental)

replicates, for a total of nine replicates (n=9). The mean values and their standard deviations were

reported. Statistical descriptive analyses were performed by using STATGRAPHICS Centurion

XVI version 16.1.11 for Windows (Statpoint Technologies Inc., Warrenton, VA, United States).

Page 51: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

42

2.3. Results and discussion

2.3.1. Effect of the buffer type and concentration

Figure 2.2 presents the specific enzymatic activities of pancreatic lipase, -amylase, alkaline

phosphatase, and protease in the presence of either Tris-HCl or PBS (pH 7.0). The specific

enzymatic activities of pancreatic lipase (Figure 2.2a) and alkaline phosphatase (Figure 2.2c) in

the presence of Tris-HCl were found to be 2.4 and 6.1 times higher, respectively, than those

obtained in the presence of PBS. Conversely, the specific enzymatic activities of -amylase

(Figure 2.2b) and protease (Figure 2.2d) in the presence of PBS were found to be 4.7 and 1.6

times higher, respectively, than those obtained in the presence of Tris-HCl. After defining the

appropriate buffer type (Tris-HCl for pancreatic lipase and alkaline phosphatase, and PBS for -

amylase and protease), the effect of buffer concentrations was then evaluated. The buffer

concentration dependence on the activity of each of the tested digestive enzymes is shown in

Figure 2.3. The specific enzymatic activities increased from the low buffer concentration region

up to a maximum value (optimum buffer concentration), and then decreased to the high buffer

concentration region. For pancreatic lipase, the maximum specific activity was reached at a

concentration of 50 mM Tris-HCl (Figure 2.3a). This value was higher than that determined for

alkaline phosphatase, in which the optimum Tris-HCl concentration was about 20 mM (Figure

2.3c). For -amylase the maximum specific activity was reached at a concentration of 20 mM

PBS (Figure 2.3b). This value was lower than that determined for protease, in which the

optimum PBS concentration was about 50 mM (Figure 2.3d). The effect of buffer on the activity

of the digestive enzymes can be addressed by considering two criteria: i) buffer capacity and ii)

buffer concentration (Bisswanger, 2014).

i) Buffer serves to adjust and stabilize the pH during the enzymatic reaction. The pKa value

indicates the pH where the buffer possesses its highest buffer capacity (Chiriac & Balea, 1997). It

is accepted that the maximum capacity of a buffer comprises a range from one pH unit below to

one pH unit above the pKa value (pH=pKa 1) (Chiriac & Balea, 1997).

Page 52: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

43

Figure 2.2. Effect of the buffer type on the enzymatic activity of pancreatic lipase (a), -amylase (b),

alkaline phosphatase (c), and protease (d). The reaction conditions are presented below. Pancreatic lipase:

Either Tris-HCl or phosphate (PBS) buffer solutions (50 mM, pH 7.0), 0.6 mM CaCl2, 600 U mL-1

enzyme concentration; and 4.0 mM pNPPA. -Amylase: Either Tris-HCl or phosphate (PBS) buffer

solutions (25 mM, pH 7.0), 15 mM NaCl, 5 U mL-1

enzyme concentration, and 0.4 mM G3CNP. Alkaline

phosphatase: Either Tris-HCl or phosphate (PBS) buffer solutions (15 mM, pH 7.0), 10 mM NaCl, 0.5 U

mL-1

enzyme concentration, and 0.4 mM pNPP. Protease: Either Tris-HCl or phosphate (PBS) buffer

solution (20 mM, pH 7.0), 15 mM NaCl, 2 U mL-1

enzyme concentration, and 1.0 mM pNPA. The

temperature of each of the enzymatic reactions was fixed at 37 °C.

0.0

0.3

0.6

0.9

1.2

Tris-HCl PBS

Lip

ase

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

Buffer (50 mM pH 7.0)

a.

0

2

4

6

8

Tris-HCl PBS

-Am

yla

se A

ctiv

ity

(m

ol

CN

P m

in-1

mg

-1p

rote

in)

Buffer (25 mM pH 7.0)

b.

0

10

20

30

Tris-HCl PBS

Ph

osp

ha

tase

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

Buffer (15 mM pH 7.0)

c.

0

4

8

12

Tris-HCl PBS

Pro

tease

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

Buffer (20 mM pH 7.0)

d.

Page 53: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

44

Figure 2.3. Effect of the buffer concentration on the enzymatic activity of pancreatic lipase (a), -amylase

(b), alkaline phosphatase (c), and protease (d). The reaction conditions are presented below. Pancreatic

lipase: Tris-HCl buffer solution (pH 7.0) at concentrations ranging from 20 to 100 mM, 0.6 mM CaCl2,

600 U mL-1

enzyme concentration; and 4.0 mM pNPPA. -Amylase: Phosphate buffer solution (PBS; pH

7.0) at concentrations ranging from 5 to 45 mM, 15 mM NaCl, 5 U mL-1

enzyme concentration, and 0.4

mM G3CNP. Alkaline phosphatase: Tris-HCl buffer solution (pH 7.0) at concentrations ranging from 5 to

45 mM, 10 mM NaCl, 0.5 U mL-1

enzyme concentration, and 0.4 mM pNPP. Protease: Phosphate buffer

solution (PBS, pH 7.0) at concentrations ranging from 10 to 60 mM, 15 mM NaCl, 2 U mL-1

enzyme

concentration, and 1.0 mM pNPA. The temperature of each of the enzymatic reactions was fixed at 37 °C.

0

3

6

9

0 10 20 30 40 50

-Am

yla

se A

ctiv

ity

(m

ol

CN

P m

in-1

mg

-1p

rote

in)

PBS (mM)

b.

0

5

10

15

20

25

0 10 20 30 40 50

Ph

osp

ha

tase

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

Tris-HCl (mM)

c.

0

3

6

9

12

0 20 40 60 80

Pro

tea

se A

ctiv

ity

(m

ol p

NP

min

-1m

g-1

pro

tein

)

PBS (mM)

d.

0.0

0.3

0.6

0.9

1.2

0 30 60 90 120

Lip

ase

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

Tris-HCl (mM)

a.

Page 54: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

45

The pKa values have been reported to be 7.20 and 8.06 for PBS and Tris-HCl, respectively

(Good, Winget, Winter, Connolly, Izawa, & Singh, 1966). Therefore, PBS is expected to have a

higher buffer capacity than Tris-HCl at pH 7.0 because the pKa of PBS is closer to the pH of the

reaction than that of Tris-HCl. Therefore, the highest buffer capacity of PBS as compared to that

of Tris-HCl may be responsible for the higher enzymatic activity exhibited by -amylase (Figure

2.2b) and protease (Figure 2.2d), when they were prepared in PBS at pH 7.0.

ii) It has been established that the more concentrated a buffer system, the higher its capacity to

stabilize the pH (Iyer & Ananthanarayan, 2008). However, most enzymes accept only moderate

concentrations of buffer, commonly between 5 and 200 mM. At low concentrations of buffer, the

enzymatic activities can be inhibited by pH shifts during the reaction (Ballou & Luck, 1941). It is

important to consider that some enzymatic reaction themselves can cause pH shifts and

consequently, a continuous decrease of the enzymatic activity, e.g., if an acid or alkaline

component becomes released during a hydrolysis reaction, such as the liberation of fatty acids by

pancreatic lipase or the liberation of amino acids by protease (Li & McClements, 2010). In such

cases, the pH should be kept constant during the reaction by using a concentrated buffer solution.

Therefore, the high buffer concentrations obtained here for pancreatic lipase (Figure 2.3a) and

protease (Figure 2.3d) necessary for reaching their maximum activities may be required to

compensate the continuous acidification of the reaction medium by the release of acid products.

Conversely, high concentrations of buffer can destabilize the intermolecular interactions which

maintain the enzymes folded, leading to denaturation (Bauduin, Nohmie, Touraud, Neueder,

Kunz, & Ninham, 2006). For example, the destabilizing effect of Tris-HCl on the tertiary

structure of -amylase (Ballou & Luck, 1941), and the inhibitory effect of PBS on the activity of

alkaline phosphatase (Coburn, Mahuren, Jain, Zubovic, & Wortsman, 1998) have been reported.

The inhibition of the activity due to the extreme concentrations of buffer was observed for all of

the tested enzymes (Figure 2.3). Therefore, the low enzymatic activities exhibited by -amylase

(Figure 2.2b) and alkaline phosphatase (Figure 2.2c) can be attributed to the destabilizing

influence of Tris-HCl and the inhibitory effect of PBS, respectively.

Page 55: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

46

2.3.2. Effect of the cofactor type and concentration

After defining the appropriate type and concentration of buffer for each of the tested enzymes,

the effect of both cofactor type and its concentration was evaluated. Figure 2.4 shows the effect

of the cofactor type on the specific enzymatic activities of pancreatic lipase, -amylase, alkaline

phosphatase, and protease. For each of the tested enzymes, enzymatic activities in the absence of

cofactor (control) were negligible. However, the addition of a cofactor significantly increased the

enzymatic activities. NaCl, CaCl2, and a mixture of both NaCl and CaCl2 were evaluated as

cofactors of the digestive enzymes. For pancreatic lipase (Figure 2.4a) addition of NaCl or CaCl2

had little effect on the enzymatic activity. However, when a mixture of both NaCl and CaCl2 was

added, the enzymatic activity was significantly increased. The results obtained for -amylase

(Figure 2.4b), alkaline phosphatase (Figure 2.4c), and protease (Figure 2.4d) were quite

different to those obtained for pancreatic lipase. The individual addition of CaCl2 had little effect

on the enzymatic activities, whereas the individual addition of NaCl significantly increased the

activities of these enzymes. In addition, when a mixture of both NaCl and CaCl2 was added, the

enzymatic activities were fairly similar to those obtained with only NaCl. This result suggests

that NaCl is the only cofactor affecting the activity of -amylase, alkaline phosphatase, and

protease, whereas CaCl2 has no significant contribution to the overall activity of these enzymes.

The cofactor concentration dependence towards the activity of each of the tested digestive

enzymes resembles in some respect the buffer concentration dependence (Figure 2.5): increasing

when raising the cofactor concentration, reaching a maximum value, followed by a decrease. For

pancreatic lipase (Figure 2.5a) the maximum specific activity was reached at concentrations of

150 and 1.0 mM for NaCl and CaCl2, respectively; whereas for -amylase (Figure 2.5b),

alkaline phosphatase (Figure 2.5c), and protease (Figure 2.5d) the maximum specific activities

were reached at concentrations of 10, 15, and 20 mM NaCl, respectively. It has been reported that

pancreatic lipase is highly dependent on calcium concentration, whereas -amylase, alkaline

phosphatase and protease do not require a cofactor for its normal operation (Schonheyder &

Volqvartz, 1945). However, the enzymatic activities of all the digestive enzymes studied here

were considerably enhanced upon addition of cofactors (Figures 2.4 and 2.5).

Page 56: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

47

Figure 2.4. Effect of the cofactor type on the enzymatic activity of pancreatic lipase (a), -amylase (b),

alkaline phosphatase (c), and protease (d). The reaction conditions are presented below. Pancreatic lipase:

50 mM Tris-HCl buffer solution (pH 7.0), 0.6 mM NaCl, 0.6 mM CaCl2, or a mixture of NaCl and CaCl2

(each component at a concentration of 0.6 mM), 600 U mL-1

enzyme concentration, and 4.0 mM pNPPA.

-Amylase: 20 mM Phosphate buffer solution (PBS, pH 7.0), 15 mM NaCl, 15 mM CaCl2, or a mixture of

NaCl and CaCl2 (each component at a concentration of 15 mM), 5 U mL-1

enzyme concentration, and 0.4

mM G3CNP. Alkaline phosphatase: 20 mM Tris-HCl buffer solution (pH 7.0), 10 mM NaCl, 10 mM

CaCl2, or a mixture of NaCl and CaCl2 (each component at a concentration of 10 mM), 0.5 U mL-1

enzyme concentration, and 0.4 mM pNPP. Protease: 50 mM Phosphate buffer solution (PBS, pH 7.0), 15

mM NaCl, 15 mM CaCl2, or a mixture of NaCl and CaCl2 (each component at a concentration of 15 mM),

2 U mL-1

enzyme concentration, and 1.0 mM pNPA. The temperature of each of the enzymatic reactions

was fixed at 37 °C. Control corresponds to the sample with no addition of cofactor.

0.0

0.3

0.6

0.9

1.2

Control NaCl CaCl2 NaCl+CaCl2

Lip

ase

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

Cofactor (0.6 mM)

a.

0

3

6

9

Control NaCl CaCl2 NaCl+CaCl2

-A

my

lase

Act

ivit

y(

mo

l C

NP

min

-1m

g-1

pro

tein

)

Cofactor (15 mM)

b.

0

10

20

30

40

Control NaCl CaCl2 NaCl+CaCl2

Ph

osp

ha

tase

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

Cofactor (10 mM)

c.

0

2

4

6

8

10

Control NaCl CaCl2 NaCl+CaCl2

Pro

tea

se A

ctiv

ity

(m

ol p

NP

min

-1m

g-1

pro

tein

)

Cofactor (15 mM)

d.

Page 57: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

48

Figure 2.5. Effect of the cofactor concentration on the enzymatic activity of pancreatic lipase (a), -

amylase (b), alkaline phosphatase (c), and protease (d). The reaction conditions are presented below.

Pancreatic lipase: 50 mM Tris-HCl buffer solution (pH 7.0), NaCl at concentrations ranging from 30 to

270 mM, and CaCl2 at concentrations ranging from 0.3 to 2.7 mM, 600 U mL-1

enzyme concentration, and

4.0 mM pNPPA. -Amylase: 20 mM Phosphate buffer solution (PBS, pH 7.0), NaCl at concentrations

ranging from 3 to 23 mM, 5 U mL-1

enzyme concentration, and 0.4 mM G3CNP. Alkaline phosphatase: 20

mM Tris-HCl (pH 7.0), NaCl at concentrations ranging from 3 to 23 mM, 0.5 U mL-1

enzyme

concentration, and 0.4 mM pNPP. Protease: 50 mM Phosphate buffer solution (PBS, pH 7.0), NaCl at

concentrations ranging from 5 to 30 mM, 2 U mL-1

enzyme concentration, and 1.0 mM pNPA. The

temperature of each of the enzymatic reactions was fixed at 37 °C.

0 1 2 3

0.0

0.3

0.6

0.9

1.2

0 100 200 300

CaCl2 (mM)

Lip

ase

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

NaCl (mM)

NaCl

CaCl2

0

2

4

6

0 5 10 15 20 25

-Am

yla

se A

ctiv

ity

(m

ol

CN

P m

in-1

mg

-1p

rote

in)

NaCl (mM)

b.a.

0

10

20

30

0 5 10 15 20 25

Ph

osp

hata

se A

ctiv

ity

(m

ol p

NP

min

-1m

g-1

pro

tein

)

NaCl (mM)

c.

0

2

4

6

8

10

0 5 10 15 20 25 30 35

Pro

tease

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

NaCl (mM)

d.

Page 58: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

49

Interestingly, pancreatic lipase required both NaCl and CaCl2 to exhibit an appreciable catalytic

activity (Figures 2.4a and 2.5a). This result can be explained by considering two different

situations: i) There might be a synergistic effect between NaCl and CaCl2 probably due to the fact

that the two metal ions can act as cofactors of pancreatic lipase (Kapoor & Gupta, 2012), and ii)

it has been reported that NaCl can operate as a cofactor of pancreatic lipase by binding to its

catalytic site, whereas CaCl2 is capable to act as both a cofactor and an allosteric activator of the

catalytic reaction (Kimura, Futami, Tarui, & Shinomiya, 1982).

There are other divalent ions such as Fe2+

, Cu2+

, and Zn2+

that may behave as cofactors.

However, Ca2+

is very important for its biological prevalence on the activity of digestive

enzymes. Unlike Fe2+

and Cu2+

, Ca2+

has a single oxidation state (+2) under physiological

conditions, therefore, Ca2+

does not participate in redox reactions catalyzed by oxidoreductases,

being a selective cofactor for hydrolases (Kimura, Futami, Tarui, & Shinomiya, 1982). Similar to

Ca2+

, Zn2+

also has a single oxidation state (+2), however, Zn2+

is a cofactor often associated to

oxidoreductases because of its capacity to form coordination bonds (Broderick, 2001). It is also

important to stress that although digestive enzymes do not require NaCl as a cofactor, the activity

of all of them was stimulated upon addition of NaCl (Figures 2.4 and 2.5). Although Na+ is not a

very common ion acting as cofactor, it has been reported that NaCl is capable to add functionality

to enzymes by providing strongly electrophilic centers (Broderick, 2001). Therefore, NaCl is a

very versatile cofactor that can be used either to increase the activity of digestive enzymes or to

replace other cofactors if required.

The effect of the cofactor concentration on the activity of the digestive enzymes can be addressed

by considering similar criteria to those considered for the buffer concentration. Enzymes are less

active at low cofactor concentrations because the substrates cannot adequately orientate in the

active site (Broderick, 2001). Conversely, high concentrations of cofactor can destabilize the

intermolecular interactions which maintain the enzymes folded, leading to denaturation (Bauduin,

Nohmie, Touraud, Neueder, Kunz, & Ninham, 2006). The inhibition of the activity due to the

extreme concentrations of cofactor was observed for all of the tested enzymes (Figure 2.5).

Page 59: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

50

2.3.3. Effect of the enzyme concentration

It has been reported that enzymatic activity is directly proportional to the enzyme concentration,

showing a linear dependence at low concentrations of enzyme (Al-Zuhair, Ramachandran, &

Hasan, 2004; Bisswanger, 2014; Miranda, Ferreira, Cordeiro, & Freire, 2006; Rami Tzafriri &

Edelman, 2007). This behavior was observed for all of the tested digestive enzymes (Figure 2.6).

The linear dependence was maintained for all of the tested enzymes until the optimum enzyme

concentration was reached. From this optimum, the enzymatic activity decreased as the enzyme

concentration increased, indicating an inhibition of the enzymatic activities by either substrate

depletion or inhibition by product (Bisswanger, 2014; Cao & De La Cruz, 2013; Goddard &

Reymond, 2004). For pancreatic lipase (Figure 2.6a), -amylase (Figure 2.6b), alkaline

phosphatase (Figure 2.6c), and protease (Figure 2.6d) the maximum specific activities were

reached at enzyme concentrations of 1000, 10, 0.8, and 3 U mL-1

, respectively.

The concentration of all components directly involved in the enzymatic reaction should be

present at saturating concentrations, excepting the enzyme, i.e., the reaction rate must be

independent of the substrate, buffer, and cofactor concentrations (Bisswanger, 2014). Unlike the

other components involved in the enzymatic reaction, the concentration of the enzyme should be

as low as possible because only catalytic amounts of enzyme are required. This is because the

Michaelis-Menten model is derived on the assumption of minor, even negligible amounts of

enzyme to ensure the enzyme saturation with the substrate (Johnson & Goody, 2011). The

evaluation of the substrate concentrations, however, was postponed for further studies (Chapter

3). Finally, the best conditions for measuring the activities of the digestive enzymes evaluated in

this study are summarized in Table 2.1.

2.3.4. Overall effect of the optimization process

The optimization of the reaction conditions allowed to improve the activities of all of the

digestive enzymes studied during each step of the optimization process.

Page 60: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

51

Figure 2.6. Effect of the enzyme concentration on the enzymatic activity of pancreatic lipase (a), -

amylase (b), alkaline phosphatase (c), and protease (d). The reaction conditions are presented below.

Pancreatic lipase: 50 mM Tris-HCl buffer solution (pH 7.0); 150 mM NaCl and 1.0 mM CaCl2; enzyme

concentration ranging from 300 to 1500 U mL-1

; and 4.0 mM pNPPA. -Amylase: 20 mM Phosphate

buffer solution (PBS; pH 7.0); 10 mM NaCl; enzyme concentration ranging from 3 to 15 U mL-1

; and 0.4

mM G3CNP. Alkaline phosphatase: 20 mM Tris-HCl buffer solution (pH 7.0); 15 mM NaCl; enzyme

concentration ranging from 0.3 to 1.3 U mL-1

; and 0.4 mM pNPP. Protease: 50 mM Phosphate buffer

solution (PBS; pH 7.0); 20 mM NaCl; enzyme concentration ranging from 1 to 5 U mL-1

; and 1.0 mM

pNPA. The temperature of each of the enzymatic reactions was fixed at 37 °C.

0.00

0.05

0.10

0.15

0.20

0 6 12 18

-Am

yla

se A

ctiv

ity

(m

ol

CN

P m

in-1

)

-Amylase (U mL-1)

b.

0.00

0.02

0.04

0.06

0.08

0.0 0.5 1.0 1.5

Ph

osp

hata

se A

ctiv

ity

(m

ol p

NP

min

-1)

Phosphatase (U mL-1)

c.

0.00

0.05

0.10

0.15

0.20

0.25

0 2 4 6

Pro

tease

Act

ivit

y(

mo

l p

NP

min

-1)

Protease (U mL-1)

d.

0.0

0.5

1.0

1.5

2.0

0 600 1200 1800

Lip

ase

Act

ivit

y(

mo

l p

NP

min

-1)

Lipase (U mL-1)

a.

Page 61: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

52

Table 2.1. Optimum conditions for the measurement of the enzymatic activities of pancreatic lipase, -

amylase, alkaline phosphatase, and protease.

Enzyme Buffer (pH 7.0) Cofactor concentration Enzyme concentration

(U mL-1

)

Pancreatic lipase 50 mM Tris-HCl 150 mM NaCl; 1 mM CaCl2 1000

-Amylase 20 mM PBS1 10 mM NaCl 10

Alkaline phosphatase 20 mM Tris-HCl 15 mM NaCl 0.8

Protease 50 mM PBS 20 mM NaCl 3

1 Phosphate buffer solution (PBS).

The use of Tris-HCl allowed increasing the activity of pancreatic lipase and alkaline phosphatase

by 2.4 and 6.1 times, respectively, as compared to PBS, whereas PBS allowed to increase the

activity of amylase and protease 2.4 and 6.1 times, respectively, as compared to Tris-HCl.

Increasing the Tris-HCl concentration from 20 to 50 mM increased 1.6 times the activity of

pancreatic lipase, whereas increasing the Tris-HCl concentration from 5 to 20 mM increased 3.9

times the activity of alkaline phosphatase. In addition, increasing the PBS concentration from 5 to

20 mM increased 2.6 times the activity of -amylase, whereas increasing the PBS concentration

from 10 to 50 mM increased 2.1 times the activity of protease.

The cofactor optimization also presented a considerable effect on the activity of all of the tested

digestive enzymes. The enzymatic activities in presence of the optimum cofactor at the optimum

concentration (NaCl and CaCl2 for pancreatic lipase, and NaCl for -amylase, alkaline

phosphatase, and protease) were 8.1, 21.1, 28.3, and 9.3 times higher for pancreatic lipase, -

amylase, alkaline phosphatase, and protease, respectively, than those obtained when the cofactor

was suppressed. Increasing the enzyme concentration from 300 to 1000 U mL-1

increased 2.9

times the activity of pancreatic lipase, whereas increasing the enzyme concentration from 3 to 10

U mL-1

increased 2.8 times the activity of -amylase. In addition, increasing the enzyme

concentration from 0.3 to 0.8 U mL-1

increased 2.6 times the activity of alkaline phosphatase,

whereas increasing the enzyme concentration from 1 to 3 U mL-1

increased 2.9 times the activity

of protease.

Page 62: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

53

Finally, the optimization of the reaction conditions had no a significant effect on both the overall

maximum enzymatic activity and the reproducibility of the measurements. The average

maximum enzymatic activities during the optimization process of pancreatic lipase, alkaline

phosphatase, and protease were 0.88 0.05, 23.25 3.72, and 7.66 0.99 μmol pNP min-1

mg-1

protein, respectively, whereas the average maximum enzymatic activity during the optimization

process of -amylase was 6.17 0.96 μmol CNP min-1

mg-1

protein. In addition, the variation

coefficients of the enzymatic activities during the optimization process were 5.7, 15.6, 16.0, and

12.9% for pancreatic lipase, -amylase, alkaline phosphatase, and protease, respectively.

2.4. Conclusions

This is the first reported evidence on the evaluation of the reaction conditions affecting the

activity of the main digestive enzymes (pancreatic lipase, -amylase, alkaline phosphatase, and

protease) by using artificial chromogenic substrates. Artificial chromogenic substrates are known

to increase the sensitivity and reliability of enzymatic assays. However, the adjustment of the

experimental conditions affecting the activity of digestive enzymes is required since the

experimental conditions reported in literature for native substrates may differ considerably to

those required for artificial substrates. The enzymatic activities of each of the tested digestive

enzymes were enhanced by selecting appropriate buffer and cofactor type as well as the

concentration of buffer, cofactor, and enzyme. In addition, the optimization of the reaction

conditions allowed enhancing the specific activities of all of the tested digestive enzymes.

However, no significant effect was observed on the reproducibility of the measurements. The best

conditions obtained here for measuring the activities of digestive enzymes (summarized in Table

2.1) will be used in further studies to evaluate the effect of pectin on the enzymatic activities of

these enzymes (Chapter 3).

Page 63: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

54

Acknowledgements

We are grateful to Departamento Administrativo de Ciencia, Tecnología e Innovación de

Colombia (COLCIENCIAS) and Vicerrectoría Académica of Universidad Nacional de Colombia

for providing a fellowship to Mauricio Espinal-Ruiz supporting this work.

References

Al-Zuhair, S., Ramachandran, K. B., & Hasan, M. (2004). High enzyme concentration model for the

kinetics of hydrolysis of oils by lipase. Chemical Engineering Journal, 103(1–3), 7-11.

Ballou, G. A., & Luck, J. M. (1941). The effects of different buffers on the activity of amylase. Journal of

Biological Chemistry, 139(1), 233-240.

Bauduin, P., Nohmie, F., Touraud, D., Neueder, R., Kunz, W., & Ninham, B. W. (2006). Hofmeister

specific-ion effects on enzyme activity and buffer pH: Horseradish peroxidase in citrate buffer.

Journal of Molecular Liquids, 123(1), 14-19.

Bisswanger, H. (2014). Enzyme assays. Perspectives in Science, 1(1–6), 41-55.

Broderick, J. B. (2001). Coenzymes and Cofactors. In eLS): John Wiley & Sons, Ltd.

Cao, W., & De La Cruz, E. M. (2013). Quantitative full time course analysis of nonlinear enzyme cycling

kinetics. Scientific Reports, 3, 2658.

Coburn, S. P., Mahuren, J. D., Jain, M., Zubovic, Y., & Wortsman, J. (1998). Alkaline Phosphatase (EC

3.1.3.1) in Serum Is Inhibited by Physiological Concentrations of Inorganic Phosphate. The

Journal of Clinical Endocrinology & Metabolism, 83(11), 3951-3957.

Chaudhuri, G., Chatterjee, S., Venu-Babu, P., Ramasamy, K., & Thilagaraj, W. R. (2013). Kinetic

behaviour of calf intestinal alkaline phosphatase with pNPP. Indian journal of biochemistry &

biophysics, 50(1), 64-71.

Chiriac, V., & Balea, G. (1997). Buffer Index and Buffer Capacity for a Simple Buffer Solution. Journal

of Chemical Education, 74(8), 937.

Daniel, R. M., & Danson, M. J. (2013). Temperature and the catalytic activity of enzymes: A fresh

understanding. FEBS Letters, 587(17), 2738-2743.

Goddard, J.-P., & Reymond, J.-L. (2004). Recent advances in enzyme assays. Trends in Biotechnology,

22(7), 363-370.

Good, N. E., Winget, G. D., Winter, W., Connolly, T. N., Izawa, S., & Singh, R. M. M. (1966). Hydrogen

Ion Buffers for Biological Research*. Biochemistry, 5(2), 467-477.

Iyer, P. V., & Ananthanarayan, L. (2008). Enzyme stability and stabilization—Aqueous and non-aqueous

environment. Process Biochemistry, 43(10), 1019-1032.

Jacobson, M. P., Kalyanaraman, C., Zhao, S., & Tian, B. (2014). Leveraging structure for enzyme

function prediction: methods, opportunities, and challenges. Trends in Biochemical Sciences,

39(8), 363-371.

Johnson, K. A., & Goody, R. S. (2011). The Original Michaelis Constant: Translation of the 1913

Michaelis–Menten Paper. Biochemistry, 50(39), 8264-8269.

Kapoor, M., & Gupta, M. N. (2012). Lipase promiscuity and its biochemical applications. Process

Biochemistry, 47(4), 555-569.

Page 64: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 2

55

Kimura, H., Futami, Y., Tarui, S.-i., & Shinomiya, T. (1982). Activation of Human Pancreatic Lipase

Activity by Calcium and Bile Salts. Journal of Biochemistry, 92(1), 243-251.

Lallès, J.-P. (2014). Intestinal alkaline phosphatase: novel functions and protective effects. Nutrition

Reviews, 72(2), 82-94.

Li, Y., & McClements, D. J. (2010). New Mathematical Model for Interpreting pH-Stat Digestion

Profiles: Impact of Lipid Droplet Characteristics on in Vitro Digestibility. Journal of Agricultural

and Food Chemistry, 58(13), 8085-8092.

Martínez Cuesta, S., Rahman, Syed A., Furnham, N., & Thornton, Janet M. (2015). The Classification and

Evolution of Enzyme Function. Biophysical Journal, 109(6), 1082-1086.

Miranda, H. V., Ferreira, A. E. N., Cordeiro, C., & Freire, A. P. (2006). Kinetic assay for measurement of

enzyme concentration in situ. Analytical Biochemistry, 354(1), 148-150.

Morishita, Y., Iinuma, Y., Nakashima, N., Majima, K., Mizuguchi, K., & Kawamura, Y. (2000). Total and

Pancreatic Amylase Measured with 2-Chloro-4-nitrophenyl-4-O-β-d-galactopyranosylmaltoside.

Clinical Chemistry, 46(7), 928-933.

Nagaki, M., Kimura, K., Kimura, H., Maki, Y., Goto, E., Nishino, T., & Koyama, T. (2001). Artificial

substrates of medium-chain elongating enzymes, hexaprenyl- and heptaprenyl diphosphate

synthases. Bioorganic & Medicinal Chemistry Letters, 11(16), 2157-2159.

Rami Tzafriri, A., & Edelman, E. R. (2007). Quasi-steady-state kinetics at enzyme and substrate

concentrations in excess of the Michaelis–Menten constant. Journal of Theoretical Biology,

245(4), 737-748.

Reis, P., Holmberg, K., Watzke, H., Leser, M. E., & Miller, R. (2009). Lipases at interfaces: A review.

Advances in Colloid and Interface Science, 147–148, 237-250.

Robic, S., Linscott, K., Aseem, M., Humphreys, E., & McCartha, S. (2011). Bile Acids as Modulators of

Enzyme Activity and Stability. The Protein Journal, 30(8), 539-545.

Schonheyder, F., & Volqvartz, K. (1945). On the Activation of Pancreatic Lipase by Calcium Chloride at

Varying pH. Acta Physiologica Scandinavica, 10(1), 62-69.

Svendsen, L., Blombäck, B., Blombäck, M., & Olsson, P. I. (1972). Synthetic chromogenic substrates for

determination of trypsin, thrombin and thrombin-like enzymes. Thrombosis Research, 1(3), 267-

278.

Tsujita, T., Takaichi, H., Takaku, T., Sawai, T., Yoshida, N., & Hiraki, J. (2007). Inhibition of lipase

activities by basic polysaccharide. Journal of Lipid Research, 48(2), 358-365.

Vendruscolo, M. (2002). Energetics of enzyme stability. Trends in Biotechnology, 20(1), 1-2.

Verma, S. K., & Ghosh, K. K. (2010). α-Chymotrypsin catalyzed hydrolysis of p-nitrophenyl acetate in

cationic microemulsions. Colloids and Surfaces A: Physicochemical and Engineering Aspects,

368(1–3), 154-158.

Wenger, D. A., Sattler, M., Clark, C., & Wharton, C. (1976). I-Cell disease: Activities of lysosomal

enzymes toward natural and synthetic substrates. Life Sciences, 19(3), 413-420.

Zor, T., & Selinger, Z. (1996). Linearization of the Bradford Protein Assay Increases Its Sensitivity:

Theoretical and Experimental Studies. Analytical Biochemistry, 236(2), 302-308.

Page 65: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

Inhibition of digestive enzyme activities by pectins in model solutions

Published as:

Espinal-Ruiz, M., Parada-Alfonso, F.; Restrepo-Sánchez, L. P., & Narváez-Cuenca, C. E.

Bioactive Carbohydrates and Dietary Fibre. 4 (2014): 27 – 38.

Page 66: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

57

Abstract

The presence of dietary fiber (e.g., pectin) in the gastrointestinal tract may decrease the caloric

uptake and reduce the risk of developing cardiovascular diseases. These phenomena are governed

by several mechanisms, such as the regulation of the rate and extent of nutrient absorption and

the alteration of the normal activity of the gastrointestinal tract enzymes. In this study, we

evaluated the effect of pectins with five methoxylation degrees (MD) on the activities of lipase,

-amylase, alkaline phosphatase, and protease. The MD of pectins ranged (in % mol/mol) from

87.4% (high methoxylated pectin, HMP) to 7.1% (low methoxylated pectin, LMP). The

enzymatic activities were evaluated in model solutions after incubation with pectins. The

Michaelis-Menten constant (Km) remained unmodified whereas the apparent maximum velocity

(Vmaxapp) decreased with increasing pectin concentrations. The Vmaxapp represented 13.3, 38.6,

41.9, and 44.4% of the Vmax (without pectins) for lipase, -amylase, alkaline phosphatase, and

protease, respectively, when they were inhibited with 100 g mL-1

HMP. Kinetic analyses

showed that all of the tested pectins behaved as non-competitive inhibitors of digestive enzymes.

Increasing both the concentration and MD of pectins, the enzymatic activities were reduced by

decreasing the non-competitive inhibition constant (Ki). In plotting Ki versus MD, a straight line

was obtained, with slopes of 1.943, 1.558, 1.344, and 1.165 g mL-1

%-1

for lipase, -amylase,

alkaline phosphatase, and protease, respectively. Among them, lipase was most likely to be

inhibited by pectins. Our results suggested that pectins might be able to suppress digestion by

inhibiting digestive enzymes.

Keywords: Pectin, methoxylation degree, digestive enzymes, hydrophobic interactions, non-

competitive inhibition.

Page 67: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

58

3.1. Introduction

The main function of the gastrointestinal tract is the absorption of nutrients derived from food

digestion. This function is controlled by a series of digestive processes that occur in different

sections of the gastrointestinal tract. These digestive processes are controlled by the secretion of

digestive enzymes and their associated cofactors and by the stability of the pH and temperature

conditions of the gastrointestinal tract (Dawson, 1993). Lipases, -amylases, alkaline

phosphatases, and proteases are the main gastric and pancreatic enzymes present in the

gastrointestinal tract (Rothman, 1977). These enzymes are responsible for the hydrolysis of the

triglycerides, carbohydrates, and proteins that are consumed in diet, which are carriers of a great

caloric content. It has been postulated that the presence of any type of dietary fiber in the

gastrointestinal tract may result in the decrease of the total caloric uptake (Amarowicz, Kmita-

Glazewska, & Kostyra, 1990). This phenomenon is governed by several mechanisms, such as the

regulation of the nutrient absorption rate, perturbations of the intestinal physiological conditions,

the encapsulation of minerals and vitamins needed for metabolic processes, and alterations of the

normal activities of the digestive enzymes (Kumar & Chauhan, 2010).

Recent epidemiological studies have shown that the consumption of dietary fiber is associated

with the reduction of the risk of developing chronic cardiovascular diseases. Consequently, it has

been recommended to increase the intake of products of plant origin that have high dietary fiber

levels, particularly soluble dietary fiber, together with other phytochemical constituents (Kris-

Etherton, et al., 2002). Pectins are a type of soluble dietary fiber that cannot be digested in the

upper gastrointestinal tract due to their resistance to the hydrolytic action of digestive enzymes

(Holloway, Tasman-Jones, & Maher, 1983). Pectins also promote bacterial fermentation in the

large intestine, improving the proliferation of intestinal microbiota that is beneficial for human

health (Louis, Scott, Duncan, & Flint, 2007). In the gastrointestinal tract, pectins form a complex

three-dimensional matrix with fibrous and amorphous characteristics. Physicochemical

properties, such as the methoxylation degree (MD), acetylation degree, molecular weight

distribution, distribution of non-methoxylated galacturonic acid residues, methoxylation and

Page 68: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

59

acetylation distribution patterns, and gel-forming capacity (Brownlee, 2011), as well as the

structure of the three-dimensional matrix, play important roles in the homeostatic and therapeutic

functionality of pectins in human nutrition (Kay, 1982). Enhancement of the gastrointestinal

viscosity, inhibition of digestion and the absorption of nutrients, control of gastrointestinal

motility and immunity, regulation of the activity of the colonic microbiota, and regulation of a

systemic stimulus associated with the feeling of satiety are among the therapeutic properties and

physiologic effects that are beneficial for human health that have been attributed to pectins

(Brownlee, 2011).

Regulation of the gastrointestinal tract enzymatic activity by dietary fiber from different plant

materials was previously reported (Dunaif & Schneeman, 1981; Isaksson, Lundquist, & Ihse,

1982a, 1982b; Ikeda & Kusano, 1983; Tsujita, Takaichi, Takaku, Sawai, Yoshida, & Hiraki,

2007). The dietary fiber materials contribute to inhibiting the activity of gastrointestinal tract

enzymes. According to the afore-cited work, the occurrence of physical interactions (such as

ionic interactions, hydrogen bonding, dispersive forces, and hydrophobic interactions) might play

an important role in the capacity of the dietary fiber to inhibit enzymatic activities. However, the

experiments conducted in those studies, did not identify a kinetic mechanism underlying the

inhibition of such enzymes. Inhibition of the activity of the enzymes by ingested pectins might

play an important role in reducing the quantity of free fatty acids, monosaccharides, and amino

acids that can be absorbed at the gastrointestinal tract level (Ikeda & Kusano, 1983). Identifying

the mechanism by which pectins act as inhibitors of digestive enzymes in model solutions might

be useful in understanding the physiological phenomena involved in the in vivo regulation that

pectins exert on nutrient absorption at the gastrointestinal tract level (Brownlee, 2011).

The aim of this study was, therefore, to evaluate in model solutions the effects of both the

concentration and MD of pectins on the activities of lipase, -amylase, alkaline phosphatase, and

protease (chymotrypsin). This study also aimed to determine a kinetic mechanism by which

pectins inhibit the activity of enzymes, as well as to evaluate theoretically, through molecular

docking calculations, the influence of structural parameters on the enzyme-pectin surface

interaction.

Page 69: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

60

3.2. Materials and methods

3.2.1. Chemicals

The enzymes porcine pancreatic -amylase, from Sus scrofa (16 U mg-1

type VI-B, E.C. 3.2.1.1)

porcine pancreatic lipase, from Sus scrofa (100 U mg-1

type II, E.C. 3.1.1.3) bovine pancreatic

protease (chymotrypsin), from Bos taurus (5 U mg-1

type I, E.C. 3.4.21.1), and bovine intestinal

mucosa alkaline phosphatase, from Bos taurus (10 U mg-1

type I, E.C. 3.1.3.1); the artificial

substrates 2-chloro-p-nitrophenyl--D-maltotrioside (G3CNP), p-nitrophenyl palmitate (pNPPA),

p-nitrophenyl acetate (pNPA), and p-nitrophenyl phosphate (pNPP); the reaction products p-

nitrophenol (pNP) and 2-chloro-p-nitrophenol (CNP); and the protein determination reagent

brilliant blue G-250 (Coomassie Blue); as well as bovine serum albumin were purchased from

Sigma-Aldrich Chemical Company (St. Louis, MO, USA). A high-methoxylated citrus pectin

(HMP) was purchased from CIMPA (Bogotá, Colombia). Ox bile extract with a cholic acid

content higher than 55% (w/w) was purchased from MP Biomedicals (Solon, OH, USA). Other

chemicals were purchased from Merck KGaA (Darmstadt, Germany).

3.2.2. Preparation and characterization of pectins with different MD levels

3.2.2.1. De-esterification of pectins

Alkaline de-esterification of the HMP was performed as described by Dongowski (1997) to

obtain pectins with different MD levels. One gram of HMP was mixed with 50 mL of 0.25 M

NaOH (pH of the mixture 10.0) and stirred for 0, 9, 15, 30, or 45 min at 25 °C. The mixture

was neutralized to pH 7.0 with 3.0 M HCl and then, 150 mL of 80% (v/v) ethanol were added to

induce pectin precipitation. The partially de-esterified pectins were filtered and washed with 100

mL of 80% (v/v) ethanol. Pectins were dried at 70 °C for 5 h and then the MD, total uronic acid

content, acetylation degree, and molecular weight distribution were evaluated.

Page 70: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

61

3.2.2.2. Characterization of pectins

3.2.2.2.1. Determination of the total uronic acid content

The total uronic acid content was determined according to van den Hoogen et al. (1998). An

aliquot of 400 L of each pectin solution (100 g mL-1

) was mixed with 2 mL of concentrated

sulfuric acid (98% w/w) containing 120 mM sodium tetraborate and incubated for 60 min at

80°C. After cooling to room temperature, the background absorbance of the samples was

measured at 540 nm. Then, 400 L of m-hydroxydiphenyl reagent (100 L of 100 mg mL-1

m-

hydroxydiphenyl in dimethyl sulfoxide, mixed with 4.9 mL of 80% (v/v) sulfuric acid) was added

and mixed with the samples. After 15 min, the absorbance of the pink-colored samples was

measured at 540 nm. A calibration line was obtained using galacturonic acid at concentrations

ranging from 0.1 to 1.0 g mL-1

(7 data points, r2=0.999). Total uronic acid content was

expressed as the moles of uronic acid residues per 100 g of pectin.

3.2.2.2.2. Determination of the methoxylation and acetylation degrees

The MDs and acetylation degrees were determined according to Voragen, Schols, and Pilnik

(1986). A model LC-20AT liquid chromatograph (Shimadzu Corporation, Kyoto, Japan)

equipped with an Aminex HPX-87H column (300 mm x 7.8 mm x 9 m; Bio-Rad Labs,

Richmond, CA, USA) was used. The column was operated at 18 °C, at a flow rate of 0.3 mL min-

1, with 5 mM sulfuric acid as the eluent. The components eluted from the column were detected

using a RID-10A refractive index detector (Shimadzu Corporation) at 40 °C. Each pectin sample

(30 mg) was suspended in 1.0 mL of 0.4 M NaOH and stirred at 18 °C for 2 h. The suspension

was then centrifuged (4,000 x g; 4 °C; 30 min) and then, 20 L of the clear supernatant was

injected into the column. The amounts of methanol and acetic acid were determined using the

external standard method. Calibration lines were obtained with methanol and acetic acid at

concentrations ranging from 5 to 100 g mL-1

(7 data points, r2=0.999 for methanol and r

2=0.997

Page 71: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

62

for acetic acid). The MD and acetylation degree were expressed as moles of methyl and acetyl

esters, respectively, per 100 mol of uronic acid.

3.2.2.2.3. Determination of the molecular weight distribution

The molecular weight distribution was determined using high performance size exclusion

chromatography (HPSEC) according to the method of Houben, Jolie, Fraeye, Van Loey, and

Hendrickx (2011), using a LC-20AT liquid chromatograph model (Shimadzu Corporation),

equipped with a mixed-bed column of TSK-gel (GMPWXL, 300 mm x 7.8 mm, pore size 100-

1000 Å, particle size 13 m; Tosoh Biosciences, Stuttgart, Germany). A 20 L injection loop was

used. Elution was performed using 50 mM NaNO3 at a flow rate of 0.7 mL min-1

for 25 min at 35

°C. A RID-10A refractive index detector (Shimadzu Corporation) at 40 °C was used to monitor

the eluents. Dextran standards with molecular weights ranging from 1.1 to 400 kDa (Sigma-

Aldrich Chemical Company, St. Louis, MO, USA) were used to estimate the molecular weight

distribution of pectins. A straight calibration curve was obtained when plotting the log of the

molecular weight versus the elution time.

3.2.3. Kinetics of the inhibition of digestive enzyme activities by pectins in model solutions

3.2.3.1. Experimental design

The kinetic inhibition profile for each enzyme by pectins was obtained according to the

methodology proposed by Kakkar, Boxenbaum, and Mayersohn (1999). Several experimental

factors were used to evaluate the activity of lipase, -amylase, alkaline phosphatase, and protease

(chymotrypsin). These experimental factors were derived by using six substrate concentration

points (ranging from 2 to 10 mM pNPPA for lipase, from 0.2 to 1.0 mM G3CNP for -amylase,

from 0.2 to 1.0 mM pNPP for alkaline phosphatase, and from 0.4 to 2.0 mM pNPA for protease),

six concentration points for each pectin (ranging from 20 to 100 g mL-1

), and five types of each

pectin, with MDs ranging (in % mol/mol) from 7.1% (low methoxylated pectin, LMP) to 87.4%

Page 72: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

63

(high methoxylated pectin, HMP). The inhibitory mechanism of each pectin sample toward the

enzyme activities was evaluated by kinetic analysis using the Lineweaver-Burk plot.

3.2.3.2. Determination of the inhibition of digestive enzyme activities by pectins

3.2.3.2.1. Determination of the activity of lipase

The enzymatic activity of lipase was determined according to Tsujita, Takaichi, Takaku, Sawai,

Yoshida, and Hiraki (2007). A reaction mixture consisting of 500 L of pNPPA at concentrations

ranging from 6 to 30 mM, 500 L of each pectin solution at concentrations ranging from 60 to

300 g mL-1

, and 500 L of a working solution of lipase at 3,000 U mL-1

(one unit is the amount

of enzyme required to convert 1 mol of pNPPA) was prepared. All of the solutions were

prepared in 50 mM Tris-HCl buffer pH 7.0 containing 150 mM NaCl, 1 mM CaCl2, 100 g mL-1

Tween-20, and 3.5 mg mL-1

bile acid extract (equivalent to 5.0 mM cholic acid in the total

mixture). The mixture was incubated at 37 °C and the absorbance of the pNP produced was

measured at 415 nm every 5 s throughout 120 s.

3.2.3.2.2. Determination of the activity of -amylase

The enzymatic activity of -amylase was determined according to Morishita, Iinuma,

Nakashima, Majima, Mizuguchi, and Kawamura (2000). A reaction mixture consisting of 500 L

of G3CNP at concentrations ranging from 0.6 to 3.0 mM; 400 L of 13.1 mg mL-1

bile acid

extract (equivalent to 18.7 mM cholic acid in the total mixture), 500 L of each pectin solution at

concentrations ranging from 60 to 300 g mL-1

, and 100 L of a working solution of -amylase

of 150 U mL-1

(one unit is the amount of enzyme required to convert 1 mol of G3CNP) was

prepared. All of the solutions were prepared in 20 mM phosphate buffer pH 7.0 containing 10

mM NaCl and 15 % (v/v) glycerol. The mixture was incubated at 37 °C and the absorbance of the

CNP produced was measured at 405 nm every 5 s throughout 60 s.

Page 73: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

64

3.2.3.2.3. Determination of the activity of alkaline phosphatase

The enzymatic activity of alkaline phosphatase was determined according to Chaudhuri,

Chatterjee, Venu-Babu, Ramasamy, and Thilagaraj (2013). A reaction mixture consisting of 500

L of pNPP at concentrations ranging from 0.6 to 3.0 mM, 400 L of 13.1 mg mL-1

bile acid

extract (equivalent to 18.7 mM cholic acid in the total mixture), 500 L of each pectin solution at

concentrations ranging from 60 to 300 g mL-1

, and 100 L of a working solution of alkaline

phosphatase of 12 U mL-1

(one unit is the amount of enzyme required to convert 1 mol of

pNPP) was prepared. All of the solutions were prepared in 20 mM Tris-HCl buffer pH 7.0

containing 15 mM NaCl. The mixture was incubated at 37 °C and the absorbance of the pNP

produced was measured at 415 nm every 5 s throughout 60 s.

3.2.3.2.4. Determination of the activity of protease

The enzymatic activity of protease (chymotrypsin) was determined according to Verma & Ghosh

(2013). A reaction mixture consisting of 500 L of pNPA at concentrations ranging from 1.2 to

6.0 mM, 400 L of 13.1 mg mL-1

bile acid extract (equivalent to 18.7 mM cholic acid in the total

mixture), 500 L of each pectin solution at concentrations ranging from 50 to 300 g mL-1

, and

100 L of a working solution of protease of 45 U mL-1

(one unit is the amount of enzyme

required to convert 1 mol of pNPA) was prepared. All of the solutions were prepared in 50 mM

phosphate buffer pH 7.0 containing 20 mM NaCl. The mixture was incubated at 37 °C and the

absorbance of the pNP produced was measured at 415 nm every 5 s throughout 60 s.

3.2.3.2.5. Determination of the specific enzymatic activities

To determine the enzymatic activities, the slope of the straight lines obtained when plotting the

absorbance versus time data obtained from the enzymatic activity experiments was interpolated

in the calibration lines for each reaction product. The calibration lines were obtained using pNP

at concentrations ranging from 5 to 50 M (7 data points, r2=0.999) for lipase, alkaline

Page 74: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

65

phosphatase, and protease and using CNP at concentrations ranging from 5 to 60 M (7 data

points, r2=0.998) for -amylase. The protein concentrations of the working enzyme solutions

were determined using the Bradford method as modified by Zor & Selinger (1996), using bovine

serum albumin (BSA) as the standard. The specific enzymatic activities of lipase, alkaline

phosphatase, and protease were expressed as mol of pNP released per minute per mg protein

(mol pNP min-1

mg-1

protein), while the specific enzyme activity of -amylase was expressed as

the mol of CNP released per minute per mg protein (mol CNP min-1

mg-1

protein).

3.2.4. Theoretical studies of digestive enzyme inhibition by pectins

3.2.4.1. Structure of the gastrointestinal enzymes

The structure of each enzyme was obtained from the Protein Data Bank

(http://www.rcsb.org/pdb, consulted on March 5th

2013). Porcine pancreatic lipase from Sus

scrofa (EC 3.1.1.3, pdb code 1ETH), porcine pancreatic -amylase from Sus scrofa (EC 3.2.1.1,

pdb code 1DHK), bovine pancreatic chymotrypsin from Bos taurus (EC 3.4.21.4, pdb code

1S0Q), and human placental alkaline phosphatase from Homo sapiens (EC 3.1.3.1, pdb code

3MK1) were used for the molecular docking calculations.

3.2.4.2. Structure of the artificial substrates

The molecular structure of the artificial substrates pNPPA, G3CNP, pNPA, and pNPP was fully

optimized using the Hartree-Fock method, as implemented in GAMESS (General Atomic and

Molecular Electronic Structure System) version May 01 of 2012 (Schmidt et al., 1993) in the 6-

311G(d,p) basis set. Solvent effects were simulated by placing the substrates in dielectric

medium simulating water, using the IEF-PCM model (Cossi, Barone, Mennucci, & Tomasi,

1998). The geometric optimization and the energy calculations were performed using this

medium.

Page 75: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

66

3.2.4.3. Structure of pectins

Pectin fragments -D-GalA-(1-4)-[-D-GalA-(1-4)]18--D-GalA-(1-1)-OH with MDs ranging

(in % mol/mol) from 0% to 100% were constructed in the biomolecular force field GLYCAM 06

for carbohydrates (Kirschner et al., 2008). Energy minimization of pectin fragments was

performed using the AMBER 9 force field, and the assigning of partial charges was conducted

using Gasteiger function partial charges. In all cases, the pdb format files that were built were

employed in the molecular docking calculations.

3.2.4.4. Molecular docking protocol

The basic docking protocol was performed using the default settings provided by AutoDock

Tools, according to Neuhaus (2010). The enzymes, artificial substrates, and pectin fragments

were converted to pdbqt format in Autodock Tools. The Lamarckian Genetic Algorithm with a

population size of 150 dockings and five million energy evaluations was used. All other

parameters, e.g., the crossover rate and mutation rate, were obtained using the default settings.

The grid size for specifying the search space was set at 21 x 21 x 21 Å3 in a centered position

with default grid-point spacing of 0.375 Å. Autodock 4.0 was launched from Autodock Tools on

Devian-Linux operating system, and the docking logs were analyzed using the graphical user

interface of Autodock Tools. The docked energy was defined as the sum of the intermolecular

and the internal energies. For a representative docking instance, the orientation or pose with the

lowest estimated free energy (G) of binding, corresponding to the docking energy and unbound

free energy of the system, was chosen in each calculation. The binding free energy ratio (BFER)

was defined as the ratio of the free energy binding of each enzyme-pectin complex and the free

energy binding of the enzyme-substrate complex.

3.2.5. Data analysis

All measurements were conducted with three analytical and three technical (instrumental)

replicates, for a total of nine replicates (n=9). The mean values and their standard deviations were

Page 76: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

67

reported. Comparisons among the mean values were performed by one-way variance analysis

(ANOVA) and using Fisher's least significant difference test (p<0.05), using the R program

(version 2.13.1, 08 July 2011).

3.3. Results and discussion

3.3.1. Characterization of pectins

Alkaline hydrolysis of HMP was performed to obtain pectins with different MD, thus pectins

obtained had MDs (% mol/mol) ranging from 87.4 5.4% (HMP) to 7.1 2.7% (LMP) (Table

3.1). In addition to the MD, the acetylation degree and molecular weight distribution of the

obtained pectins were also evaluated because these molecules are potentially susceptible the

effects of the alkaline hydrolysis conditions, and these features determine the structure, three-

dimensional conformation, and functional properties of pectins (Mohnen, 2008). The alkaline

treatment did not affect (p<0.05) the acetylation degree [average of acetylation degree was 4.9

0.4% (mol/mol)] nor the molecular weight values (HPSEC profiles of the obtained pectins did not

show significant differences; Figure 3.1). Alkaline hydrolysis was performed at low temperature

(pH 10.0; 25 °C; 45 min) because it has been established that alkaline hydrolysis

Table 3.1. Effect of the alkaline hydrolysis of high methoxylated pectin on the methoxylation and

acetylation degrees.

Time (min) Methoxylation (% mol/mol)1 Acetylation (% mol/mol)

2

0 87.4 5.4ª 4.4 0.3ª

9 64.6 5.3b 4.8 0.3

a

15 39.1 5.0c 5.3 0.1

a

30 28.4 3.8d 4.9 0.4

a

45 7.1 2.7e 5.4 0.4

a

1 Expressed as moles of methyl esters per 100 mol of uronic acid.

2 Expressed as moles of acetyl esters per 100 mol of uronic acid.

Different letters within the same column indicate significant differences as calculated using Fisher's least

aaaaaaaaaaaaaaasignificant difference test (p<0.05).

Page 77: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

68

Figure 3.1. High performance size exclusion chromatography (HPSEC) elution profiles of pectins after 0

(continuous line) and 45 min (dotted line) of de-esterification with 0.25 M NaOH at 25 °C, to obtain

pectins with methoxylation degrees of 87.4 and 7.1% (mol/mol), respectively. The scale was shifted

upwards by 2 refractive index (RI) units for the 45 min profile (dotted line).

of the -(1,4) glycosidic bond of pectins should not be performed at temperatures above 60 °C to

prevent -elimination from becoming the primary mechanism of degradation (Krall &

McFeeters, 1998). We found that the molecular weight distribution was not affected by this

process (Figure 3.1). This result is most likely because hydrolysis of the -(1,4) glycosidic bond

of pectins, which is required to decrease the molecular weight, is highly efficient under extreme

temperature and acidity conditions (pH ≤ 2.0; 100 °C; 3 h) (Ramaswamy, Kabel, Schols, &

Gruppen, 2013). It is well known that the alkaline hydrolysis conditions may affect both the MD

and acetylation degree, depending on the temperature, alkali concentration, and the initial number

of available methyl and acetyl groups present in pectin. Nevertheless, in this study, we found that

the alkaline conditions did not affect significantly the acetylation degree, most likely due to the

low number of initial acetylation sites available compared to the high number of initial

methoxylation sites present in HMP (Sundar-Raj, Rubila, Jayabalan, & Ranganathan, 2012;

Garna, Mabon, Nott, Wathelet, & Paquot, 2006). In contrast to the stability of the acetylation

degree and molecular weight during the alkaline treatment, there was a statistically significant

0

2

4

6

8

10

0 5 10 15 20

RI

Res

po

pn

se

Time (min)

Page 78: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

69

decrease (p<0.05) of the MD over time. After 45 min of alkaline treatment, almost complete de-

esterification of HMP was obtained, yielding an LMP with a MD of 7.1 2.7% (mol/mol).

The MD was the only structural parameter that differentiated pectins used in the enzyme

inhibition experiments, meaning that the observed differences in the biological functionalities of

pectins are most likely due to differences in their MD. The structural importance of MD in

pectins lies in the fact that the methyl-ester groups neutralize the negative charges present in the

free carboxyl groups, decreasing their polar nature and increasing their hydrophobicity (Mohnen,

2008). Thus, the MD is a parameter that affects the biological functionality of pectins and it can

be a factor that determines the interaction of them with other biologically relevant biomolecules,

such as proteins (Benjamin, Lassé, Silcock, & Everett, 2012).

3.3.2. Kinetics of the inhibition of the digestive enzymatic activities by pectins

The effect of the MD and of the concentration of each pectin sample on the activities of the

enzymes lipase, -amylase, alkaline phosphatase, and protease (chymotrypsin) was evaluated.

The kinetic profiles of the in vitro digestion of artificial substrates by the enzymes in the presence

of HMP are shown in Figure 3.2. Increasing the concentration of HMP decreased (p<0.05) the

activity of lipase, -amylase, alkaline phosphatase, and protease. The same trend was observed

using each of the other pectins. Because the increase in the concentration of each pectin sample in

the reaction medium caused a decrease in the activity of each enzyme, one might suppose that the

tested pectins displayed inhibitory behavior toward the enzymatic activities and that pectins

therefore might behave kinetically as a defined type of enzyme inhibitor (competitive, non-

competitive or uncompetitive). Inhibitory behavior towards enzymes was previously reported for

other types of dietary fiber, such as cellulose (Dunaif & Schneeman, 1981), agar-agar,

carboxymethyl cellulose, sodium alginate, xylan, inulin (Ikeda & Kusano, 1983) pectin,

mucilages, polyethylene glycol (Isaksson, Ihse, & Lundquist, 1982a), and dietary fiber extracted

from wheat bran, oat bran, and alfalfa (Dunaif & Schneeman, 1981). In all cases, increasing the

substrate concentration did not overcome the inhibitory effect of pectins (the maximum velocity,

Vmax, of each enzyme inhibited by pectins did not equal the Vmax of the uninhibited enzyme at

Page 79: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

70

Figure 3.2. Effect of the concentration of high methoxylated pectin (HMP, methoxylation degree of

87.4% mol/mol) on the kinetic behaviors of lipase (a), -amylase (b), alkaline phosphatase (c), and

protease (d). The concentrations of HMP were 0 (), 20 (), 40 (), 60 (), 80 (), and 100 g mL-1

(). Points correspond to experimental data and lines correspond to fitted data according to the Michaelis-

Menten equation.

any substrate concentration). In addition, lipase was the enzyme most affected by the inhibitory

effect of each pectin sample among the enzymes studied (Figure 3.2a).

That HMP utilized the kinetic mechanism of non-competitive inhibition of the activities of the

studied enzymes was confirmed using the double-reciprocal plot method of Lineweaver-Burk

0

2

4

6

0.0 0.2 0.4 0.6 0.8 1.0

-A

my

lase

Act

ivit

y(

mo

l C

NP

min

-1m

g-1

pro

tein

)

G3CNP (mM)

b.

0.0

0.4

0.8

1.2

0 2 4 6 8 10

Lip

ase

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

pNPPA (mM)

a.

0

10

20

30

0.0 0.2 0.4 0.6 0.8 1.0

Ph

osp

ha

tase

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

pNPP (mM)

c.

0

3

6

9

0.0 0.5 1.0 1.5 2.0

Pro

tease

Act

ivit

y(

mo

l p

NP

min

-1m

g-1

pro

tein

)

pNPA (mM)

d.

Page 80: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

71

(Figure 3.3). All of the tested pectins behaved as non-competitive inhibitors of the enzymes. This

behavior suggested that pectins can interact with both the free enzyme and the enzyme-substrate

complex (ES) at a domain different from the catalytic site of the enzymes (Rehm & Becker,

1988). It was also possible to verify that the Vmax of the enzymes inhibited with pectins did not

equal the Vmax of the non-inhibited enzyme at any substrate concentration because the apparent

maximum velocity (Vmaxapp) decreased significantly as pectin concentrations increased (Figure

3.4; only the results from using HMP are shown). The presence of pectins in the reaction medium

did not significantly modify the ES affinity or the efficiency of ES complex formation

(kinetically represented by the Michaelis-Menten constant, Km), which are distinctive behaviors

of non-competitive enzyme inhibition.

The variance analysis (ANOVA, with p<0.05) performed for each kinetic profile (from Figure

3.4) revealed that neither the MD levels nor pectin concentration affected the Km value. The

observed kinetic mechanism of non-competitive inhibition explains the fact that this type of

inhibition cannot be overcome by increasing the substrate concentration because in this type of

inhibition pectins do not compete for the catalytic site of the enzyme. This result is important in

understanding the enzyme-pectin interaction phenomenon because it could be expected that an

interaction of enzymes with pectins could be only superficial (e.g., at an allosteric regulation site)

without any compromise of the catalytic site of the enzyme.

The decrease in the Vmaxapp (Equation 3.1) was dependent on the concentration of each pectin

sample. According to the non-competitive inhibition mechanism, this parameter corresponds to

the following expression:

[ ]

(3.1)

Where Ki is the non-competitive inhibition constant of an inhibited enzyme with each pectin

sample at a given concentration [Pectin]. This constant governs the process of non-competitive

inhibition of each enzyme. Using Equation 3.1, it was possible to find the Ki constant for each

Page 81: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

72

Figure 3.3. Lineweaver-Burk plot of lipase (a), -amylase (b), alkaline phosphatase (c), and protease (d)

inhibited with high methoxylated pectin (HMP, methoxylation degree of 87.4% mol/mol). The kinetic

profile obtained suggested that HMP behaves as a non-competitive inhibitor of digestive enzymes. The

concentrations of HMP were 0 (), 20 (), 40 (), 60 (), 80 (), and 100 g mL-1

(). Points

correspond to experimental data and lines correspond to fitted data according to the Lineweaver-Burk

equation.

pectin sample with a given MD. Figure 3.5 shows that a linear tendency, with a negative slope,

was revealed upon plotting the Ki against MD. An increase in the MD caused a significant

decrease in the Ki. In the context of non-competitive enzymatic inhibition, the Ki can be

interpreted as the inhibitor concentration required to reduce the enzymatic activity by 50%

0

2

4

6

8

10

-1.2 -0.8 -0.4 0.0 0.4 0.8

1/L

ipa

se A

ctiv

ity

1/pNPPA (mM-1)

a.

0.0

0.5

1.0

1.5

-3 0 3 6

1/

-Am

yla

se A

ctiv

ity

1/G3CNP (mM-1)

b.

0.00

0.05

0.10

0.15

0.20

-4 -2 0 2 4 6

1/P

ho

sph

ata

se A

ctiv

ity

1/pNPP (mM-1)

c.

0.0

0.1

0.2

0.3

0.4

0.5

-4 -2 0 2 4

1/P

rote

ase

Act

ivit

y

1/pNPA (mM-1)

d.

Page 82: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

73

Figure 3.4. Effect of the concentration of high methoxylated pectin (HMP, methoxylation degree of

87.4% mol/mol) on the apparent maximum velocity (Vmaxapp) and Michaelis-Menten constant (Km) of

lipase (a), -amylase (b), alkaline phosphatase (c), and protease (d). In non-competitive inhibition, the Km

remains constant whereas the Vmaxapp decreases when the concentration of HMP increases. An ANOVA

analysis (p<0.05) showed that the HMP concentration did not significantly affect the Km.

(IC50, Cheng & Prussof, 1973), namely, pectins with a high Ki (LMP) are less efficient as a non-

competitive inhibitor of enzymes than pectins with a low Ki (HMP). All of the tested enzymes

were inhibited by pectins with any MD by means of the same molecular mechanism (non-

competitive inhibition). The efficiency of the inhibition was determined from the slope (m) of the

straight line that was obtained when plotting Ki vs MD (Figure 3.5).

0.0

0.5

1.0

1.5

2.0

0.0

0.5

1.0

1.5

0 20 40 60 80 100

Km

(mM

)Vm

ax

Ap

p

(m

ol p

NP

min

-1m

g-1

pro

tien

)

Pectin (g mL-1)

Vmax App

Km

a.

0.0

0.2

0.4

0.6

0.8

1.0

0

2

4

6

8

10

0 20 40 60 80 100

Km

(mM

)Vm

ax

Ap

p

(m

ol

CN

P m

in-1

mg

-1p

roti

en)

Pectin (g mL-1)

Vmax App

Km

b.

0.0

0.2

0.4

0.6

0.8

1.0

0

10

20

30

40

0 20 40 60 80 100

Km

(mM

)Vm

ax

Ap

p

(m

ol p

NP

min

-1m

g-1

pro

tien

)

Pectin (g mL-1)

Vmax App

Km

c.

0.0

0.2

0.4

0.6

0.8

1.0

0

2

4

6

8

10

0 20 40 60 80 100

Km

(mM

)Vm

ax

Ap

p

(m

ol p

NP

min

-1m

g-1

pro

tien

)

Pectin (g mL-1)

Vmax App

Km

d.

Page 83: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

74

Figure 3.5. Effect of the methoxylation degree (MD) of pectins on the non-competitive inhibition constant

(Ki) of lipase (a), -amylase (b), alkaline phosphatase (c), and protease (d). The efficiency of inhibition

was measured as the slope (m) of the straight line obtained when plotting Ki vs MD. High methoxylated

pectin (HMP) has the highest ability to inhibit the activities of the digestive enzymes (r2=0.984, 0.986,

0.961, and 0.930 for lipase, -amylase, alkaline phosphatase, and protease, respectively).

Interestingly, each enzyme was inhibited with different efficiencies. Lipase (m = -1,943 g mL-1

%-1

) was more likely to be inhibited by pectins than were -amylase (m = -1,558 g mL-1

%-1

),

alkaline phosphatase (m = -1,344 g mL-1

%-1

), or protease (m = -1,165 g mL-1

%-1

). These

differences could be attributed to the interaction capacity of each enzyme and pectins, which is

0

50

100

150

200

0 20 40 60 80 100

Ki(

g m

L-1

)

MD (% mol/mol)

a.

m = -1.943 0.112

0

50

100

150

200

0 20 40 60 80 100

Ki(

g m

L-1

)

MD (% mol/mol)

b.

m = -1.558 0.109

0

50

100

150

200

0 20 40 60 80 100

Ki(

g m

L-1

)

MD (% mol/mol)

c.

m = -1.344 0.154

0

50

100

150

200

0 20 40 60 80 100

Ki(

g m

L-1

)

MD (% mol/mol)

d.

m = -1.165 0.139

Page 84: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

75

mediated by the structural characteristics of each of the molecule involved in the interaction

(Rodríguez-Patiño & Pilosof, 2011).

Ikeda & Kusano (1983) have suggested two hypotheses to explain the inhibitory effect of dietary

fiber on the activity of trypsin (a protease). The first hypothesis states that the inhibition of

enzymatic activity may result from the interaction of dietary fiber with the substrate, preventing

the enzyme-substrate interaction, and the second hypothesis states that there is a direct interaction

between dietary fiber and trypsin, making the enzyme unable to perform its catalytic function.

Regarding these hypothesis, we found that the inhibitory effect of pectins on the activities of the

enzymes appears to be due to a non-competitive enzyme-pectin interaction. The non-competitive

inhibitory mechanism of pectins on the activities of the enzymes revealed in this study (Figure

3.3) is most likely due to the large size of pectins, which generally have sizes and molecular

dimensions comparable to those of the enzymes, suggesting that the interaction of pectin with an

enzyme may be superficial (Zhao, Diao, & Zong, 2013). Nevertheless, it is important to consider

that the overall inhibitory effect of pectins toward a gastrointestinal tract activity might be caused

by several mechanisms, such as modification of the composition and structure of the interface,

substrate coating with a pectin layer or embedding of the substrates within pectin particles

(Miled, Beisson, de Caro, de Caro, Arondel, & Verger, 2001; Reis, Holmberg, Watzke, Leser, &

Miller, 2009; McClements & Li, 2010). Also, it has been previously established that the

inhibition of some digestive processes can be enhanced by the increase of viscosity of the

gastrointestinal fluids and by flocculation of lipids due to the presence of polysaccharides in the

gastrointestinal tract (McClements, 2000). At sufficiently high concentration, polysaccharides

form a three-dimensional network of interacting or entangled molecules than traps substrates and

enzymes and effectively inhibit their movements and interactions (Dickinson, 2009).

3.3.3. Nature of the enzyme-pectin interaction

The interaction of pectins with enzymes may be controlled by the structural parameters of both

pectins (MD, acetylation degree, and molecular weight distribution) and the digestive enzymes

(size, molecular weight, surface hydrophobicity, and three-dimensional structure).

Page 85: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

76

Table 3.2. Structural properties of lipase, -amylase, alkaline phosphatase, and protease. Lipase has the

greater probability of having surface interactions with hydrophobic pectins (such as high methoxylated

pectin, HMP) due to their larger size and its large exposed surface area. The high content of hydrophobic

amino acids confers upon the lipase enzyme its predominantly hydrophobic character and its high capacity

for interaction with HMP.

Enzyme MW1 pI

2 V

3 As

4 THA (%)

5 Gtr w→o

6

Lipase 99.7 (896) 6.11 (-12.0) 170,960 31,063 58.1 (521) -74.5

-Amylase 55.4 (496) 6.17 (-5.2) 111,734 16,811 56.0 (278) -40.5

Alkaline phosphatase 52.7 (484) 6.12 (-7.8) 76,503 14,611 52.5 (254) -35.7

Protease 23.3 (223) 8.34 (+6.3) 33,795 6,753 49.8 (111) -16.1

1 MW: Molecular weight (kDa). The number of total amino acids is in parentheses. Obtained from Protein Data Bank. 2 pI: Isoelectric point. The estimated charge at pH 7.0 is in parentheses. Obtained from Protein Data Bank. 3 V: Protein volume (Å3). Calculated according to Voss & Gerstein (2010). 4 As: Accessible surface area (Å2). Calculated according to Miller, Janin, Lesk & Chothia (1987). 5 THA: Total hydrophobic amino acids (sum of Phe, Trp, His, Tyr, Ala, Ile, Leu, Val, Pro, and Gly). The percent of hydrophobic

residues is given as number of hydrophobic residues per 100 residues. The number of hydrophobic residues is in parentheses.

Obtained from Protein Data Bank. 6 Gtr w→o (kcal mol-1): free energy transfer from water to the bilayer interface (surface hydrophobicity). Calculated according to

Eisenhaber (1996).

Table 3.2 shows some of the structural characteristics of the enzymes of interest in this study.

Size is a structural parameter that might be important in the interaction of each enzyme with

pectins. The size of the enzymes may be represented by several structural characteristics, such as

the molecular weight, volume and accessible surface area. Lipase, with the highest degree of

inhibition by pectins, is the enzyme with the highest molecular weight, volume, and surface

accessible area of those studied. The large size of the lipase molecule might result in an increased

probability of interaction with pectins and therefore, an increased susceptibility of being inhibited

by them. Thus, as the size (As) of the enzyme decreases (lipase, 31,063 Å2; -amylase, 16,811

Å2; alkaline phosphatase, 14,611 Å

2; and protease, 6,753 Å

2), its ability to interact with and to be

inhibited by pectins also decreases.

It was also observed that each enzyme was more susceptible to inhibition by HMP than by LMP.

This behavior could be governed by the physical nature of the non-covalent intermolecular forces

operating in the enzyme-pectin interaction (McClements, 2006). The difference in the

hydrophobicity of LMP and HMP might also explain the non-covalent interaction of pectins with

each protein. Due to the presence of negatively charged free carboxyl groups, LMP is hydrophilic

Page 86: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

77

in nature and can interact more efficiently with proteins via an electrostatic mechanism. In

contrast, due to the presence of methoxylated carboxyl (carbomethoxyl ester group) groups of

neutral nature, HMP is more hydrophobic and can interact more efficiently with proteins via a

hydrophobic mechanism. Lipase, -amylase, and alkaline phosphatase, which have isoelectric

points (pI) slightly below the pH (pH 7.0) of the model solution, would have a negative

electrostatic charge (net charge of the enzymes at pH 7.0 in all cases is very close to neutrality).

The negative electrostatic nature of these enzymes could hinder their interaction with LMP due to

negative electrostatic repulsions. Such repulsive electrostatic forces could explain the weak

interaction between the enzymes and the LMP.

The total hydrophobic amino acids (THA) and the free transfer energy of the enzyme from the

aqueous phase (w) toward the organic (lipidic) phase (o, Gtr wo) were calculated to estimate the

hydrophobicity (Eisenhaber, 1996) of each enzyme. Lipase (containing 58.1% THA) is the

enzyme with the greatest number of hydrophobic amino acids (Phe, Trp, His, Tyr, Ala, Ile, Leu,

Val, Pro, and Gly) present in its structure; whereas, as the hydrophobicity of the enzymes

decreased (-amylase, 56.0% THA; alkaline phosphatase, 52.5% THA; and protease, 49.0%

THA), their ability to interact and be inhibited by HMP also decreased. The parameter Gtr wo

(Equation 3.2) is defined as:

(

) (Eq. 3.2)

Where So corresponds to the solubility of the enzyme in the organic (lipidic) phase and Sw

corresponds to the solubility of the enzyme in the aqueous phase. In the case of hydrophobic

enzymes, for which So > Sw, the Gtr wo is negative, whereas in hydrophilic enzymes, for which

So < Sw, the Gtr wo is positive. The more negative the Gtr wo, the greater the So, as well as the

hydrophobicity of the enzyme, meaning that a protein has a strong tendency to be transferred

from within the aqueous solution to the interface.

Page 87: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

78

Examining this parameter showed that lipase (Gtr wo = -74.51 kcal mol-1

) is the enzyme with

the highest hydrophobicity, whereas -amylase, alkaline phosphatase, and protease were the

enzymes with low hydrophobicity. Interestingly, the enzymes with low Gtr wo values (-

amylase, -40.46 kcal mol-1

; alkaline phosphatase, -35.66 kcal mol-1

; and protease, -16.07 kcal

mol-1

) were those which were less inhibited by HMP than lipase. This finding might indicate that

electrostatic interactions are less important in the overall enzyme-pectin interaction as compared

to hydrophobic interactions, which are generally dominant (Miled, Beisson, de Caro, de Caro,

Arondel, & Verger, 2001; McClements, 2006), and that the mechanism that could be determinant

in the enzyme-pectin interactions might be hydrophobic in nature (McClements, 2006; Hur, Lim,

Decker, & McClements, 2011).

The interfacial phenomena that occur at the oil-water interface of the simulated gastrointestinal

fluid utilized in this study also might govern the efficiency of the enzyme-pectin interaction.

Lipase is the enzyme that is most affected by interfacial characteristics, such as the substrate, bile

salts, and lipase interfacial concentrations, as well as by its hydrophobic character and its high

susceptibility to be affected by the highly hydrophobic HMP molecule (McClements, 2006; Reis,

Holmberg, Watze, Leser, & Miller, 2009). Lipases are water-soluble enzymes with a limited

activity toward substrates in aqueous media, but they exhibit high activity when the substrate is at

a concentration high enough to form micelles in the presence of a surfactant or when the substrate

is present in an emulsified medium (Miled, Beisson, de Caro, de Caro, Arondel, & Verger, 2001).

This particular behavior occurs because lipases are enzymes that are resistant to the denaturing

conditions of interfaces, such as the high concentration of surfactant agents (Reis, Holmberg,

Watzke, Leser, & Miller, 2009). It can be suggested, therefore, that lipases are enzymes that are

particularly susceptible to being inhibited by pectins, not only due to the direct effect they have

on the activity of these enzymes, but also due to the effect they have on the stability of the

emulsion formed under the simulated gastrointestinal tract conditions. It has also been suggested

that pectins can behave as anionic surfactants that may compete with a lipase for a position in the

interface, displacing the enzyme and inhibiting its activity by decreasing its interfacial

concentration (Reis, Holmberg, Watze, Leser, & Miller, 2009).

Page 88: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

79

In this study we did not conduct any emulsion preparation for any enzyme test. For the lipase

model, the chromogenic group of the pNPPA allows this compound to remain in aqueous

solution. It was only necessary the addition of a small amount of Tween 20 to improve the

solubility of pNPPA in the buffer solution. It is known, nevertheless, that to catalyze its

enzymatic reaction, lipase must absorb to the oil-water interface so that it is in close proximity to

lipid droplets. Lipase generally does this as part of a complex with co-lipase and possibly with

bile-salts (Singh, Ye, & Horne, 2009). Competitive adsorption processes occur at the lipid-

droplet surfaces among the enzyme complex, bile salts, phospholipids, digestion products, and

other surface-active substances such as pectins, which could interfere with lipase adsorbing to the

droplet surfaces. Thus, in an emulsified system (not tested by us), lipid-droplet surfaces might be

coated with pectin layers that inhibit the direct access of the lipase/co-lipase complex to the lipid

droplets (Singh, Ye, & Horne, 2009; McClements & Li, 2010).

3.3.4. Theoretical studies of digestive enzyme inhibition by pectins

To evaluate the differences in the enzyme-pectin interaction that were experimentally observed in

this study, a theoretical model was proposed. This model allowed calculating the interaction

energy for each enzyme-pectin pair. The interaction energies were calculated using molecular

docking methodology, by searching for the site on the enzyme where the interaction with pectin

was most likely to occur. Figure 3.6 shows the structures of the enzyme-HMP interaction. In all

cases, the enzyme-pectin interaction occurred at a site different from the catalytic site of each

enzyme (where substrate is located), which is consistent with a non-competitive inhibition

mechanism. Due to the large size that pectins can have, they were not expected to have a

significant interaction with the catalytic site of an enzyme, which generally represents only a

small fraction of the total surface area of an enzyme. The sites of the enzymes for which pectins

had more affinity were protein domains with a relatively high abundance of superficial

hydrophobic amino acids. For example, in the case of lipase, it was found that HMP had high

affinity for a highly hydrophobic domain that includes Gly47, Leu136, Leu140, Trp131, Tyr142,

Gly354, Tyr373, Gly415, Trp436, Val437, and Leu443, all of which are amino acids with a

hydrophobic nature.

Page 89: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

80

Figure 3.6. Structure of the enzyme-substrate-pectin complexes formed by the substrates and lipase (a),

-amylase (b), alkaline phosphatase (c), or protease (d) when inhibited by high methoxylated pectin

(HMP) with a methoxylation degree of 100% (mol/mol). In all cases, pectin fragments were docked on the

surface of the enzymes at a site different from the active (catalytic) site.

This observation suggested that the hydrophobic interactions in the enzyme-pectin coupling are

the interactions that could mostly govern the non-competitive enzyme inhibition that was

experimentally observed. As was experimentally observed, although all of the enzymes were

inhibited by pectins with any MD by means of the same molecular mechanism, each enzyme was

inhibited with different efficiencies. Figure 3.7 shows the calculated enzyme-pectin interaction

energies expressed as the BFER. The BFER value corresponds to the relationship between the

free energy of the enzyme-pectin interaction and the free energy of the enzyme-substrate

interaction. The BFER is a measure of how many fold greater is the affinity of enzyme for pectin

than that for its respective substrate. It was observed that for any MD, all the enzymes had greater

affinity for pectins than for the substrates. The BFER increased as the MD increased.

a. b.

c. d.

Page 90: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

81

Figure 3.7. Effect of the methoxylation degree (MD, ranging from 0% to 100% mol/mol) of pectins on the

free energy of binding of the enzyme-pectin complexes. The binding free energy ratio (BFER) was

defined as the ratio of the free energy of binding of an enzyme-pectin complex and the free energy of

binding of the enzyme-substrate complex (dimensionless). Lipase (), -amylase (), alkaline

phosphatase (), and protease ().

These results suggested that an increase in the MD promotes enzyme-pectin interactions and

increases the efficiency of the non-competitive enzyme inhibition. The trend of inhibition

observed in our experiments (Figure 3.5) was theoretically validated (Figures 3.6 and 3.7). For a

pectin with a theoretical MD of 100% (mol/mol), lipase had 18.4 times greater affinity for the

inhibitor than for the substrate, whereas -amylase (12.3 times), alkaline phosphatase (8.0 times),

and protease (4.2 times) had weaker interactions with the inhibitor (Figure 3.7). Furthermore, for

a pectin with a theoretical MD of 0% (mol/mol), lipase had 6.2 times greater affinity for the

inhibitor than for the substrate, whereas -amylase (4.3 times), alkaline phosphatase (2.7 times)

and protease (1.4 times) were less inhibited (Figure 3.7). Thus, it could be suggested that the

enzyme-pectins hydrophobic interactions are critical to the efficiency with which pectins inhibit

the activity of the enzymes via a non-competitive mechanism.

0

5

10

15

20

0 40 80 120

BF

ER

MD (% mol/mol)

Page 91: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

82

Figure 3.8. Non-competitive inhibition of digestive enzyme activities upon addition of pectin. Digestive

enzyme (yellow), pectin (white), substrate (red), and product (blue).

Finally, our results allow us to suggest that the mechanism by which pectin is able to reduce the

digestive enzyme activities is the non-competitive inhibition (Figure 3.8): Pectin binds to the

digestive enzymes somewhere other than the active site, or the substrate cannot correctly

orientate on the active site of the digestive enzymes.

3.4. Conclusions

The kinetic analyses performed in this study suggested that pectins function as non-competitive

inhibitors of the gastrointestinal tract enzymes. The high degree of methoxylation of pectins and

the hydrophobicity of the enzymes significantly contribute to the enzyme-pectin surface

interactions and to the efficiency of the inhibitory effect of pectins on the in vitro activity of the

enzymes in model solutions. The order of the magnitude with which the enzymes were inhibited

by pectins was lipase>-amylase>alkaline phosphatase>protease. Based on our results, we

suggest that pectins might be able to inhibit nutrient digestion in the small intestine by inhibiting

gastrointestinal tract enzymatic activities.

Enzymatic

reaction

Inhibition of

enzymatic activity

No orientation of

the substrate

Page 92: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

83

Acknowledgments

The authors are grateful to COLCIENCIAS and the Universidad Nacional de Colombia for

providing a fellowship to Mauricio Espinal-Ruiz that supported this study.

References

Amarowicz, R., Kmita-Glazewska, H., & Kostyra, H. (1990). Influence of fiber on the enzymatic

digestion of casein. Fagopyrum, 10, 69-72.

Benjamin, O., Lassé, M., Silcock, P., & Everett, D. W. (2012). Effect of pectin adsorption on the

hydrophobic binding sites of -lactoglobulin in solution and in emulsion systems. International

Dairy Journal, 26, 36-40.

Brownlee, I. A. (2011). The physiological roles of dietary fiber. Food Hydrocolloids, 25, 238-250.

Chaudhari, G., Chatterjee, S., Venu-Babu, P., Ramasamy, K., & Thilagaraj, W. R. (2013). Kinetic

behavior of calf intestinal alkaline phosphatase with pNPP. Indian Journal of Biochemistry &

Biophysics, 50, 64-71.

Cheng, Y. C., & Prussof, W. H. (1973). Relationship between the inhibition constant (Ki) and the

concentration of inhibitor which causes 50 per cent inhibition (I50) of an enzymatic reaction.

Biochemical Pharmacology, 22, 3099-3108.

Cossi, M., Barone, V., Mennucci, B., & Tomasi, J. (1998). Ab initio study of ionic polarizable continuum

dielectric model. Chemical Physics Letters, 286, 253-260.

Dawson, D. C. (1993). Gastrointestinal physiology: the molecular basis of G.I. transport. Annual Review

of Physiology, 55, 571–573.

Dickinson, E. (2009). Hydrocolloids as emulsifiers and emulsion stabilizers. Food Hydrocolloids, 23,

1473-1482.

Dongowski, G. (1997). Effect of the pH on the in vitro interactions between bile acids and pectin.

European Food Research and Technology, 205, 185-192.

Dunaif, G., & Schneeman, B. O. (1981). The effect of dietary fiber on human pancreatic enzyme activity

in vitro. The American Journal of Clinical Nutrition, 34, 1034-1035.

Eisenhaber, F. (1996). Hydrophobic regions on protein surfaces: derivation of the solvation energy from

their area distribution in crystallographic protein structures. Protein Science, 5, 1676-1686.

Garna, H., Mabon, N., Nott, K., Wathelet, B., & Paquot, M. (2006). Kinetic of the hydrolysis of pectin

galacturonic acid chains and quantification by ionic chromatography. Food Chemistry, 96, 477-

484.

Holloway, W. D., Tasman-Jones, C., & Maher, K. (1983). Pectin digestion in humans. The American

Journal of Clinical Nutrition, 37, 253-255.

Houben, K., Jolie, R. P., Fraeye, I., Van Loey, A. M., & Hendrickx, M. E. (2011). Comparative study of

the cell wall composition of broccoli, carrot, and tomato: structural characterization of the

extractable pectins and hemicelluloses. Carbohydrate Research, 346, 1105-1111.

Hur, S. J., Lim, B. O., Decker, E. A., & McClements, D. J. (2011). In vitro human digestion models for

food applications. Food Chemistry, 125, 1-12.

Ikeda, K., & Kusano, T. (1983). In vitro inhibition of digestive enzymes by indigestible polysaccharides.

Cereal Chemistry, 60, 260-263.

Page 93: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

84

Isaksson, G., Lundquist, I., & Ihse, A. (1982a). Effect of dietary fiber on pancreatic enzyme activity in

vitro. Gastroenterology, 82, 918-924.

Isaksson, G., Lundquist, I., & Ihse, A. (1982b). In vitro inhibition of pancreatic enzyme activities by

dietary fiber. Digestion, 24, 54-59.

Kakkar, T., Boxenbaum, H., & Mayersohn, M. (1999). Estimation of Ki in a competitive enzyme-

inhibition model: comparisons among three methods of data analysis. Drug Metabolism and

Disposition, 27, 756-762.

Kay, R. M. (1982). Dietary fiber. Journal of Lipid Research, 23, 221-242.

Kirschner, K. N., Yongye, A. B., Tschampel, S. M., González-Outeiriño, J., Daniels, C. R., Foley, L., &

Woods, R. J. (2008). GLYCAM 06: a generalizable biomolecular force field. Carbohydrates.

Journal of Computational Chemistry, 29, 622-655.

Krall, S. M., & McFeeters, R. F. (1998). Pectin hydrolysis: effect of temperature, degree of methoxylation,

pH, and calcium on hydrolysis rates. Journal of Agricultural and Food Chemistry, 46, 1311-1315.

Kris-Etherton, P. M., Hecker, K. D., Bonanome, A., Coval, S. M., Binkoski, A. E., Hilpert, K. F., Griel,

A. E., & Etherton, T. D. (2002). Bioactive compounds in foods: their role in the prevention of

cardiovascular disease and cancer. The American Journal of Medicine, 113, 71S-88S.

Kumar, A., & Chauhan, G. S. (2010). Extraction and characterization of pectin from Apple pomace and its

evaluation as lipase (steapsin) inhibitor. Carbohydrate Polymers, 82, 454-459.

Louis, P., Scott, K. P., Duncan, S. H., & Flint, H. J. (2007). Understanding the effects of diet on bacterial

metabolism in the large intestine. Journal of Applied Microbiology, 102, 1197-1208.

McClements, D. J. (2000). Comments on viscosity enhancement and depletion flocculation by

polysaccharides. Food Hydrocolloids, 14, 173-177.

McClements, D. J. (2006). Non-covalent interactions between proteins and polysaccharides.

Biotechnology Advances, 24, 621-625.

McClements, D. J., & Li, Y. (2010). Review of in vitro digestion models for rapid screening of emulsion-

based systems. Food & Function, 1, 32-59.

Miled, N., Beisson, F., de Caro, J., de Caro, A., Arondel, V., & Verger, R. (2001). Interfacial catalysis by

lipases. Journal of Molecular Catalysis B: Enzymatic, 11, 165-171.

Miller, S., Janin, J., Lesk, A. M., & Chothia, C. (1987). Interior and surface of monomeric proteins.

Journal of Molecular Biology, 196, 641-656.

Mohnen, D. (2008). Pectin structure and biosynthesis. Current Opinion in Plant Biology, 11, 266-277.

Morishita, Y., Iinuma, Y., Nakashima, N., Majima, K., Mizuguchi, K., & Kawamura, Y. (2000). Total and

pancreatic amylase measured with 2-chloro-4-nitrophenyl-4-O--D-galactopyranosyl maltoside.

Clinical Chemistry, 46, 928-933.

Neuhaus, F. C. (2010). Roles of Arg301 in substrate orientation and catalysis in subsite 2 of D-alanine: D-

alanine (D-lactate) ligase from Leuconostoc mesenteroides: A molecular docking study. Journal

of Molecular Graphics and Modelling, 28, 728-734.

Ramaswamy, U. R., Kabel, M. A., Schols, H. A., & Gruppen, H. (2013). Structural features and water

holding capacities of pressed potato fibre polysaccharides. Carbohydrate Polymers, 93, 589-596.

Rehm, H., & Becker, C. M. (1988). Interpreting non-competitive inhibition. Trends in Pharmacological

Sciences, 9, 316-317.

Reis, P., Holmberg, K., Watzke, H., Leser, M. E., & Miller, R. (2009). Lipases at interfaces: A review.

Advances in Colloid and Interface Science, 147-148, 237-250.

Rodríguez-Patiño, J. M., & Pilosof, A. M. R. (2011). Protein – polysaccharide interactions at fluid

interfaces. Food Hydrocolloids, 25, 1925-1937.

Rothman, S. S. (1977). The digestive enzymes of the pancreas: a mixture of inconstant proportions.

Annual Review of Physiology, 39, 373-389.

Schmidt, M. W., Baldridge, K. K., Boatz, J. A., Elbert, S. T., Gordon, M. S., Jensen, J. H., Koseki, S.,

Matsunaga, N., Nguyen, K. A., Su, S., Windus, T. L., Dupuis, M., & Montgomery, J. A. (1993).

Page 94: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 3

85

General atomic and molecular electronic structure system. Journal of Computational Chemistry,

14, 1347-1363.

Singh, H., Ye, A., & Horne, D. (2009). Structuring food emulsions in the gastrointestinal tract to modify

lipid digestion. Progress in Lipid Research, 48, 92-100.

Sundar-Raj, A. A., Rubila, S., Jayabalan, R., & Ranganathan, T. V. (2012). A review on pectin: chemistry

due to general properties of pectin and its pharmaceutical uses. Scientific Reports, 1, 550-553.

Tsujita, T., Takaichi, H., Takaku, T., Sawai, T., Yoshida, N., & Hiraki, J. (2007). Inhibition of lipase

activities by basic polysaccharides. Journal of Lipid Research, 48, 358-365.

van den Hoogen, B. M., van Weeren, P. R., Lopes-Cardozo, M., van Golde, L. M. G., Barneveld, A., &

van de Lest, C. H. A. (1998). A microtiter plate assay for the determination of uronic acids.

Analytical Biochemistry, 257, 107-111.

Verma, S. K., & Ghosh, K. K. (2013). Activity, stability and kinetic parameters of -chymotripsin

catalyzed reactions in AOT/isooctane reverse micelles with nonionic and zwitterionic mixed

surfactants. Journal of Chemical Science, 125, 1-8.

Voragen, A. G. J., Schols, H. A., & Pilnik, W. (1986). Determination of the degree of methoxylation and

acetylation of pectins by HPLC. Food Hydrocolloids, 1, 65-70.

Voss, N. R., & Gerstein, M. (2010). 3V: cavity, channel and cleft volume calculator and extractor. Nucleic

Acids Research, 38, W555-W562.

Zhao, G. Y., Diao, H. J., & Zong, W. (2013). Nature of pectin-protein-cathechin interactions in model

systems: Pectin-protein-cathechin interactions. Food Science and Technology International, 19,

153-165.

Zor, T., & Selinger, Z. (1996). Linearization of the Bradford protein assay increases its sensitivity:

theoretical and experimental studies. Analytical Biochemistry, 236, 302-308.

Page 95: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

Impact of dietary fibers [methyl cellulose, chitosan, and pectin]

on digestion of lipids under simulated gastrointestinal conditions

Published as:

Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L. P., Narváez-Cuenca, C. E.,

& McClements, D. J. Food & Function. 5 (2014): 3083 – 3095.

Page 96: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

87

Abstract

A simulated in vitro digestion model was used to elucidate the impact of dietary fibers on the

digestion rate of emulsified lipids. The influence of polysaccharide type (chitosan (cationic),

methyl cellulose (non-ionic), and pectin (anionic)) and initial concentration (0.4 to 3.6% (w/w))

was examined. 2% (w/w) corn oil-in-water emulsions stabilized by 0.2% (w/w) Tween-80 were

prepared, mixed with polysaccharides, and then subjected to an in vitro digestion model (37 °C):

initial (pH 7.0); oral (pH 6.8; 10 min); gastric (pH 2.5; 120 min); and, intestinal (pH 7.0; 120

min) phases. The impact of polysaccharides on lipid digestion, -potential, particle size,

viscosity, and stability was determined. The rate and extent of lipid digestion decreased with

increasing pectin, methyl cellulose, and chitosan concentrations. The free fatty acids released

after 120 min of lipase digestion were 46, 63, and 81% (w/w) for methyl cellulose, pectin, and

chitosan, respectively (3.6% (w/w) initial polysaccharide), indicating that methyl cellulose had

the highest capacity to inhibit lipid digestion, followed by pectin and chitosan. In the absence of

polysaccharides, lipid droplets remained stable to flocculation throughout the digestion model.

Methyl cellulose and pectin promoted depletion flocculation of the lipid droplets, whereas

chitosan promoted bridging flocculation. These results have important implications for

understanding the influence of dietary fibers on lipid digestion, since they promote droplet

flocculation and therefore inhibit digestion. The control of lipid digestibility within the

gastrointestinal tract might be important for the development of reduced-calorie emulsion-based

functional food products.

Keywords: Pectin, chitosan, methyl cellulose, emulsion, lipid digestion, gastrointestinal tract,

flocculation.

Page 97: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

88

4.1. Introduction

Diets rich in fat have been associated with high incidences of obesity and elevated risks of

coronary heart disease, diabetes, and certain forms of cancer (Bray, Paeratakul, & Popkin, 2004;

van Dam & Seidell, 2007). A potential strategy for combating these chronic diseases is therefore

to reduce the total level of fat present in food products (Chung, Degner, & McClements, 2013;

Mao & Julian McClements, 2012; Wu, Degner, & McClements, 2013). However, the

development of these products is challenging because fats have a major impact on the

physicochemical, sensory, and nutritional properties of foods (Heertje, 2014; Narine &

Marangoni, 1999). For instance, fat contributes to the desirable texture of dairy products

(Rousseau, 2000), the mouthfeel and texture of bakery products (Ghotra, Dyal, & Narine, 2002),

and the creamy texture, milky appearance, desirable flavor, and satiating effects of emulsion-

based products, such as sauces, spreads, dressings, and dips (Ghosh & Rousseau, 2011). Foods

with reduced fat levels must therefore be carefully formulated to ensure that they maintain their

desirable physicochemical, sensory, and nutritional properties (e.g., appearance, flavor, texture,

shelf life, and satiety effects), otherwise they will not be acceptable to consumers (Heertje, 2014).

Rather than simply reducing the total amount of fat present within foods, it may also be possible

to improve their healthfulness using other strategies associated with controlling fat digestion. For

example, if the rate and extent of lipid digestion within the small intestine can be decreased then

the spike in blood lipid levels that normally occurs can be reduced (Michas, Micha, & Zampelas,

2014). In addition, retarded lipid digestion may also increase the feelings of satiety and satiation,

which may lead to lower overall calorie consumption (Kritchevsky, 1988; Li, Hu, &

McClements, 2011; Li & McClements, 2014). Dietary fibers are known to have an impact on the

behavior of lipids within the gastrointestinal tract and can therefore be used to modulate the

response of humans to ingested lipids (Galisteo, Duarte, & Zarzuelo, 2008; Gunness & Gidley,

2010; Kritchevsky, 1988; Slavin, 2005). Dietary fibers may influence lipid digestion through a

wide variety of different mechanisms (Gunness & Gidley, 2010): i) they may bind to species that

play a critical role in digestion such as bile salts, phospholipids, enzymes or calcium

(Dongowski, 2007); ii) they may increase the viscosity of the intestinal phase, and thereby alter

Page 98: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

89

mass transport processes (Fabek, Messerschmidt, Brulport, & Goff, 2014; Kristensen & Jensen,

2011); iii) they may form protective coatings around lipid droplets thereby inhibiting lipase

access (Li & McClements, 2014; Simo, Mao, Tokle, Decker, & McClements, 2012; Tokle,

Lesmes, & McClements, 2010); iv) they may promote lipid droplet aggregation thereby changing

the amount of lipid surface exposed to lipase (Hur, Lim, Decker, & McClements, 2011;

McClements & Li, 2010a); v) they may inactive digestive enzymes (Beysseriat, Decker, &

McClements, 2006; Brownlee, 2011; Espinal-Ruiz, Parada-Alfonso, Restrepo-Sánchez, &

Narváez-Cuenca, 2014); and vi) they may alter their microbial population within the large

intestine (Castillo, Martín-Orúe, Anguita, Pérez, & Gasa, 2007). The ability of dietary fibers to

impact lipid digestion through these and other mechanisms ultimately depends on their molecular

and physicochemical properties (Mudgil & Barak, 2013). At present, there is a relatively poor

understanding of the relationship between dietary fiber structure and their impact on the lipid

digestion process.

In the present study, we used a simulated static in vitro digestion model to study the influence of

a cationic (chitosan), non-ionic (methyl cellulose), and anionic (pectin) polysaccharides on the

digestion of emulsified lipid droplets. The aim of the study was to obtain a better understanding

of the role of dietary fiber characteristics on the gastrointestinal fate of ingested lipids. The

knowledge gained from this study might be useful for the fabrication of healthier functional food

products designed to promote health and wellness (Chung, Degner, & McClements, 2013; Khan,

Grigor, Winger, & Win, 2013).

4.2. Materials and methods

4.2.1. Chemicals

Corn oil was purchased from a commercial food supplier (Mazola, ACH Food Companies Inc.,

Memphis, TN, USA) and stored at 4 °C until use. The manufacturer reported that the corn oil

contained approximately 14, 29, and 57% (w/w) of saturated, monounsaturated, and

polyunsaturated fatty acids, respectively. Powdered methyl cellulose (M0262, 41 kDa molecular

Page 99: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

90

weight, 27.5-31.5% (mol/mol) methylation degree, viscosity of 2% (w/w) aqueous solution

=400 cps, and powdered chitosan (448877, medium molecular weight (190-310 kDa), 75-85%

deacetylation, viscosity of 1% (w/w) solution in 1% (w/v) acetic acid =200-800 cps, were

purchased from Sigma-Aldrich Chemical Company (St Louis, MO, USA).

Commercial powdered high methoxylated pectin (Genu Pectin (Citrus), USP/100) was kindly

donated by CP Kelco (Lille Skensved, Denmark) and was used without further purification. The

composition of this material as provided by the manufacturer was 6.9% (w/w) moisture, 89.0%

(mol/mol) galacturonic acid, and 8.6% (w/w) methoxyl groups, which corresponds to a degree of

methoxylation of approximately 62% (mol/mol). The average molecular weight was reported by

the manufacturer as 200 kDa. Fat soluble fluorescent dye Nile Red (N3013), lipase from porcine

pancreas (Type II, L3126, triacylglycerol hydrolase E.C. 3.1.1.3), bile extract (porcine, B8631),

mucin from porcine stomach (Type II, M2378, bound sialic acid ≤ 1.2% w/w), and pepsin A from

porcine gastric mucose (P7000, endopeptidase E.C. 3.4.23.1, activity ≥ 250 units mg-1

solid) were

purchased from Sigma-Aldrich Chemical Company (St Louis, MO, USA). The supplier has

reported that lipase activity is 100-400 units mg-1

protein (using olive oil) and 30-90 units mg-1

protein (using triacetin) for 30 min incubation (one unit of lipase activity was defined as the

amount of enzyme required for the release of 1 eq of fatty acid from either triacetin (pH 7.4) or

olive oil (pH 7.7) in 1 h at 37 °C).

The composition of the bile extract has been reported as 49% (w/w) total bile salt (BS),

containing 10-15% glycodeoxycholic acid, 3-9% taurodeoxycholic acid, 0.5-7.0% deoxycholic

acid, 1-5% hydrodeoxycholic acid, and 0.5-2.0% cholic acid; 5% (w/w) phosphatidyl choline

(PC); Ca2+

≤0.06% (w/w); critical micelle concentration of bile extract 0.07 0.04 mM; and mole

ratio of BS to PC being around 15:1. All other chemicals were purchased from Sigma-Aldrich

Chemical Company (St Louis, MO, USA). Double distilled water was used to prepare all

solutions.

Page 100: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

91

4.2.2. Solution and emulsion preparation

4.2.2.1. Polysaccharide stock solutions preparation

Pectin, chitosan, and methyl cellulose stock solutions (4% w/w) was prepared by dispersing 10 g

of powdered pectin, chitosan, or methyl cellulose into 240 g of 5 mM phosphate buffer pH 7.0 for

pectin and methyl cellulose, and 5 mM acetate buffer pH 4.0 for chitosan. The solutions were

stirred at 800 rpm for 12 h (overnight) at room temperature to ensure complete dispersion and

dissolution. Pectin, chitosan, and methyl cellulose stock solutions were finally adjusted to pH 7.0

using 1.0 M sodium hydroxide and hydrochloric acid solutions, and were equilibrated for 10 min

before analysis.

4.2.2.2. Stock emulsion preparation

A stock emulsion was prepared by mixing together 20% (w/w) corn oil and 80% (w/w) buffered

emulsifier solution (5 mM phosphate buffer pH 7.0, containing 2.5% (w/w) Tween 80) for 5 min

using a bio-homogenizer (Speed 2, Model MW140/2009-5, Biospec Products Inc., ESGC,

Switzerland). The coarse emulsion obtained was then passed 5 times through a high-pressure

homogenizer (Microfluidizer M-110L processor, Microfluidics Inc., Newton, MA, USA)

operating at 11,000 psi (75.8 MPa).

4.2.2.3. Polysaccharide-emulsion mixtures preparation

Polysaccharide-emulsion mixtures were prepared by mixing the stock emulsion (containing

20.0% (w/w) corn oil and 2.0% (w/w) Tween 80) with buffered stock solutions of 4.0% (w/w)

chitosan (cationic), methyl cellulose (non-ionic), or pectin (anionic), to obtain systems of varying

composition: 2.0% (w/w) corn oil, 0.2% (w/w) Tween 80, and 0.2-3.6% (w/w) polysaccharide

(corresponding to mass ratio polysaccharide to corn oil ranging from 0.1 to 1.8). The emulsion-

polysaccharide mixtures were then stirred with a high-speed stirrer (Fisher Steadfast Stirrer,

Model SL 1200, Fisher Scientific, Pittsburgh, PA, USA) at 1000 rpm and stored overnight at

Page 101: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

92

room temperature. The emulsion-polysaccharide mixtures were characterized to obtain the initial

phase, prior to subjection to the static in vitro digestion model.

4.2.3. Static in vitro digestion model

Each emulsion sample (initial phase) was passed through a simulated static in vitro digestion

model that consisted of oral (Section 4.2.3.1), gastric (Section 4.2.3.2), and intestinal (Section

4.2.3.3) phases. Measurements of emulsion microstructure and stability, particle size distribution,

particle charge, and viscosity were performed after each phase (Section 4.2.4). The standardized

static in vitro digestion model used in this study was a modification of those described previously

(Li, Hu, & McClements, 2011; Minekus, Alminger, Alvito, Ballance, Bohn, Bourlieu, et al.,

2014).

4.2.3.1. Oral phase

Simulated saliva fluid (SSF, pH 6.8) containing 3.0% (w/w) mucin was prepared according to the

composition shown in Table 4.1. The SSF composition was based on those reported in previous

studies (Mao & McClements, 2012). Each emulsion (initial phase) was mixed with SSF (ratio 1:1

w/w) and the resulting mixture containing 1% (w/w) corn oil and 0.1-1.8% (w/w) pectin was used

for characterization after the incubation period. The oral phase model consisted of a conical flask

containing emulsion-SSF mixture incubated at 37 °C with continuous shaking at 100 rpm for 10

min in a temperature controlled air incubator (Excella E24 Incubator Shaker, New Brunswick

Scientific, NJ, USA) to mimic the conditions in the mouth. The resulting oral phase (bolus) was

used in the gastric phase (Section 4.2.3.2).

4.2.3.2. Gastric phase

Simulated gastric fluid (SGF) was prepared by adding 2.0 g NaCl, 7.0 mL concentrated HCl

(37% w/w), and 3.2 g pepsin A (from porcine gastric mucose, 250 units mg-1

) to a flask and then

diluting with double distilled water to a volume of 1.0 L, and finally adjusting to pH 1.2

Page 102: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

93

Table 4.1. Chemical composition of simulated saliva fluid (SSF) used to simulate oral conditions.

Compound Chemical formula Concentration (g L-1

)1

Sodium chloride NaCl 1.594

Ammonium nitrate NH4NO3 0.328

Potassium dihydrogen phosphate KH2PO4 0.636

Potassium chloride KCl 0.202

Potassium citrate K3C6H5O7•H2O 0.308

Uric acid sodium salt C5H3N4O3Na 0.021

Urea H2NCONH2 0.198

Lactic acid sodium salt C3H5O3Na 0.146

Porcine gastric mucin (Type II) ---- 30

1The SSF was prepared in double distilled water and then pH 6.8 was adjusted using 0.1 M NaOH.

using 1.0 M HCl. Samples taken from the oral phase (bolus) were mixed with SGF (ratio 1:1

w/w) so that the final mixture contained 12.0 mM NaCl, 0.16% (w/w) pepsin A (corresponding to

an enzymatic activity of 400 units mL-1

), 0.5% (w/w) corn oil, and 0.05-0.90% (w/w) pectin. This

mixture was then adjusted to pH 2.5 using 1.0 M NaOH and incubated at 37 °C with continuous

shaking at 100 rpm for 2 h (this time represents the half emptying of a moderately nutritious and

semi-solid meal (Hur, Lim, Decker, & McClements, 2011)). Since lipase activity is markedly

lower in the gastric compartment compared to that in the duodenal tract, the addition of gastric

lipase in this phase can be omitted (McClements & Li, 2010a). Samples were taken for

characterization at the end of the incubation period (gastric phase). The resulting gastric phase

(chyme) was used in the intestinal phase (Section 4.2.3.3).

4.2.3.3. Intestinal phase

Samples obtained from the gastric phase (20.0 mL chyme containing 0.5% (w/w) corn oil and

0.05-0.90% (w/w) pectin) were incubated for 2 h at 37 °C in a simulated small intestine fluid

(SIF) containing 2.5 mL pancreatic lipase (24.0 mg mL-1

), 3.5 mL bile extract solution (54.0 mg

mL-1

), and 1.5 mL salt solution containing 0.25 M CaCl2 and 3.0 M NaCl, to obtain a final

Page 103: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

94

composition of the intestinal fluid in the reaction vessel of 0.36% (w/w) corn oil, 0.05-0.65%

(w/w) pectin, 2.0 mg mL-1

pancreatic lipase (corresponding to an enzymatic activity of 550 units

mL-1

), 7.0 mg mL-1

bile extract, 15.0 mM CaCl2, and 150.0 mM NaCl. The free fatty acids (FFA)

released were monitored by determining the amount of 0.10 M NaOH needed to maintain a

constant pH 7.0 within the reaction vessel using an automatic titration unit (pH stat titrator, 835

Titrando, Metrohm USA, Inc., Riverview, FL, USA). All additives were dissolved in phosphate

buffer solution (5 mM, pH 7.0) before use. Lipase addition and initialization of the titration

program were carried out only after the addition of all pre-dissolved ingredients and balancing

the pH to 7.0. Samples were taken for physicochemical and structural characterization at the end

of the digestion period (intestinal phase). The volume of 0.10 M NaOH added to the emulsion

was recorded over time and then was used to calculate the concentration of FFA generated by

lipolysis. The amount of FFA (% w/w) released was calculated using the following equation:

(

) (4.1)

Where, CNaOH is the concentration of the sodium hydroxide (0.10 M), MWLipid is the average

molecular weight of corn oil (872 g mol-1

), WLipid is the initial weight of corn oil in the intestinal

phase (0.10 g), and VNaOH is the volume of NaOH (L) titrated into the reaction vessel to

neutralize the FFA released, assuming that all triacylglycerols (TAG) are hydrolyzed in two

molecules of FFA and one molecule of monoacylglycerol (MAG). Titration blanks were

performed by inactivating lipase in boiling water for 15 min prior to initialization of the titration

program.

4.2.4. Emulsion characterization

4.2.4.1. Creaming stability measurements

Ten milliliters of emulsion was transferred into a test tube (internal diameter 15 mm, height 125

mm), tightly sealed with a plastic cap, and then stored at room temperature for 24 h, after which

Page 104: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

95

appreciable phase separation into an opaque layer at the top, a turbid layer in the middle, and a

transparent layer at the bottom was observed in some of the systems. We defined the serum layer

to be the sum of the turbid and transparent layers. The total height of the emulsion (HE) and the

height of the serum layer (HS) were measured using a laser vertical profiling system (Turbiscan

Classic MA2000, Formulaction, Wynnewood, PA, USA). The extent of creaming was then

characterized by the creaming index (CI) as follows: ⁄ . The creaming index

provided indirect information about the extent of droplet aggregation.

4.2.4.2. Emulsion microstructure

The microstructure of the emulsions was characterized by confocal microscopy. An optical

microscopy (C1 Digital Eclipse, Nikon, Tokyo, Japan) with a 60x objective lens was used to

capture images of the emulsions. Emulsions were gently stirred to form a homogeneous mixture

without introducing air bubbles. A small aliquot of the emulsions (6 L) was then transferred to a

glass microscope slide and covered with a glass cover slip. The cover slip was fixed to the slide

using nail polish to avoid evaporation. A small amount of immersion oil (Type A, Nikon,

Melville, NY, USA) was placed on the top of cover slip. Emulsions samples were stained with fat

soluble fluorescent dye Nile Red (0.1% (w/w) dissolved in ethanol) to visualize the location of

the oil phase. All confocal images were taken using an excitation (543 nm) argon laser and

emitted light was collected between 555-620 nm, and then characterized using the instrument

software (EZ CS1 version 3.8, Niko, Melville, NY, USA).

4.2.4.3. Apparent viscosity measurements

The apparent viscosity () of samples was measured using a dynamic shear rheometer (Kinexus

Rotational Rheometer, Malvern Instruments Ltd., Worcestershire, UK). A cup and bob geometry

consisting of a rotating inner cylinder (diameter 25.0 mm) and a static outer cylinder (diameter

27.5 mm) was used. The samples were loaded into the rheometer measurement cell and allowed

to equilibrate at 37 °C for 5 min before the beginning all experiments. Samples underwent a

Page 105: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

96

constant shear treatment (10 s-1

; 10 min) prior to analysis to standardize the shear rate of each

sample. The apparent viscosity was then obtained from measurements with a shear rate of 10 s-1

.

4.2.4.4. Particle size distribution measurements

The emulsions were diluted to a droplet concentration of approximately 0.005% (w/w) using

buffer solution at the appropriate pH prior to analysis to avoid multiple scatterings effects. The

particle size distribution of emulsions was then measured using a static light scattering instrument

(Mastersizer 2000, Malvern Instruments Ltd., Worcestershire, UK). A refractive index ratio of

1.47 (corn oil) was used in the calculations of the particle size distribution. Background

corrections and system alignment were performed prior to each measurement when the

measurement cell was filled with the appropriate buffer solution. Particle sizes were reported as

particle size distribution profiles (volume fraction (%) vs. particle diameter (m)) for a mass ratio

polysaccharide:corn oil of 1.8.

4.2.4.5. Particle electrical charge measurements

The surface electrical charge (-potential) of emulsions was determined using a particle

electrophoresis instrument (Zetasizer NanoSeries, Malvern Instruments Ltd., Worcestershire,

UK). The emulsions were diluted to a droplet concentration of approximately 0.005% (w/w)

using buffer solution at the appropriate pH prior to analysis. Diluted emulsions were injected into

the measurement chamber, equilibrated for 120 s and then the -potential was determined by

measuring the direction and velocity that the droplets moved in the applied electric field. Each -

potential measurement was calculated from the average of 20 continuous readings made per

sample. To determine the effect of pH on the -potential of the polysaccharides (0.5% w/w), a

titration between pH 2.0-8.0 was performed with an automatic titration unit (Multi Purpose

Titrator MPT-2, Malvern Instruments Ltd., Worcestershire, UK) and 0.25 M NaOH. The -

potential was recorded at each pH after 60 s equilibrium.

Page 106: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

97

4.2.5. Data analysis

All measurements were performed at least three times using freshly prepared samples. Averages

and standard deviations were calculated from these triplet measurements.

4.3. Results and discussion

4.3.1. Electrical properties of dietary fibers

Initially, we measured the -potential versus pH profiles of the three polysaccharides used in this

study to characterize their electrical properties (Figure 4.1). The -potential of the chitosan went

from highly positive at pH 2.0 to close to zero at pH 8.0, which can be attributed to the presence

of cationic amino groups (–NH2 + H2O –NH3⊕

+ OH⊝) with a pKa value around pH 6.5 along

the polymer backbone (Yuan, Gao, Decker, & McClements, 2013). The -potential of the methyl

cellulose was close to zero across the entire pH range studied due to the fact that it is a neutral

polymer with no charged groups. The -potential of the pectin went from close to zero at pH 2.0

to highly negative at pH 8.0, which can be attributed to the presence of anionic carboxyl groups

(–COOH + H2O –COO⊝ + H3O⊕) with a pKa value around pH 3.5 (Jones, Lesmes, Dubin, &

McClements, 2010). Visual observations of the samples indicated that they remained transparent

across the entire pH range studied, suggesting that self-association, precipitation, and

sedimentation did not occur.

4.3.2. Influence of dietary fibers on physicochemical properties of lipid droplets in

simulated gastrointestinal tract (GIT)

In this series of experiments, we examined the influence of the three polysaccharides on the

physicochemical and structural properties of lipid droplets as they passed through a simulated

GIT. Different types and amounts of dietary fiber were mixed with stock emulsion, and then the

properties of the resulting mixtures were characterized as they were passed through the simulated

Page 107: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

98

Figure 4.1. Influence of pH on the surface electrical charge (-potential) of diluted 0.5% (w/w) chitosan,

methyl cellulose, and pectin solutions.

oral, gastric, and small intestine stages. The particle size distribution, microstructure, charge, and

stability of the samples were measured after each stage of the GIT model.

4.3.2.1. Initial samples

The particle size distribution measured by static light scattering (SLS) indicated that all of the

initial emulsions contained relatively small droplets, with a monomodal distribution with a peak

around 310 nm (Figure 4.2a). Confocal microscopy images suggested that there were very large

flocs present in the emulsions containing methyl cellulose and pectin, and some small flocs in the

emulsions containing chitosan (Figure 4.3). The fact that droplet flocculation was not evident in

the light scattering data, but was in the microscopy images, can be attributed to the fact that the

emulsions were highly diluted prior to SLS measurements, which will breakdown any weakly

flocculated droplets (McClements, 2000).

-60

-40

-20

0

20

40

60

2 3 4 5 6 7 8

-P

ote

nti

al

(mV

)

pH

Chitosan

Methyl Cellulose

Pectin

Page 108: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

99

Figure 4.2. Influence of chitosan, methyl cellulose, and pectin (mass ratio polysaccharide:corn oil of 1.8)

on the particle size distribution of emulsions under simulated gastrointestinal conditions consisting of an

initial (a), oral (b), gastric (c), and intestinal (d) phases. Control corresponds to the emulsion without the

addition of polysaccharides. The scale was shifted upwards by 25, 50, and 75% for chitosan, methyl

cellulose, and pectin, respectively.

One would not expect an electrostatic attraction between anionic or neutral polymers and oil

droplets stabilized by a non-ionic surfactant. We therefore attribute the extensive droplet

flocculation observed in the emulsions containing methyl cellulose or pectin to a depletion effect

(Furusawa, Ueda, & Nashima, 1999), i.e., the generation of an osmotic attraction between the

droplets due to the exclusion of non-adsorbed polymers from the droplet

0

20

40

60

80

100

0.05 0.5 5 50 500

Volu

me

Fra

ctio

n (

%)

Particle Diameter (m)

Pectin

Methyl Cellulose

Chitosan

Control

0

20

40

60

80

100

0.05 0.5 5 50 500

Volu

me

Fra

ctio

n (

%)

Particle Diameter (m)

Pectin

Methyl Cellulose

Chitosan

Control

0

20

40

60

80

100

0.05 0.5 5 50 500

Volu

me

Fra

ctio

n (

%)

Particle Diameter (m)

Pectin

Methyl Cellulose

Chitosan

Control

c.

0

20

40

60

80

100

0.05 0.5 5 50 500

Volu

me

Fra

ctio

n (

%)

Particle Diameter (m)

Pectin

Methyl Cellulose

Chitosan

Control

d.

a. b.

Page 109: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

100

Figure 4.3. Influence of chitosan, methyl cellulose, and pectin (mass ratio polysaccharide:corn oil of 1.8)

on the microstructure of emulsions observed by confocal fluorescence microscopy under simulated

gastrointestinal conditions consisting of an initial (a), oral (b), gastric (c), and intestinal (d) phases.

Control corresponds to the emulsion without the addition of polysaccharides.

surfaces (Jenkins & Snowden, 1996; McClements, 2000). Conversely, the small amount of

flocculation observed in the emulsions containing chitosan may be attributed to either a depletion

or bridging effect (Furusawa, Ueda, & Nashima, 1999). Measurement of the -potential of the

Tween 80-stabilized oil droplets indicated that they had a slight negative charge (-6.0 mV) at

neutral pH, which may have been due to the presence of anionic impurities (such as fatty acids)

in the oil or surfactant, or due to preferential adsorption of hydroxyl ions (rather than hydronium

ions) from water by the oil droplet surfaces (Nikiforidis & Kiosseoglou, 2011). Thus, there may

have been a weak electrostatic attraction between the anionic fat droplets and cationic chitosan

molecules initially leading to some bridging flocculation in this system (Biggs, Habgood,

Jameson, & Yan, 2000). In addition, any non-adsorbed chitosan molecules may have promoted

ChitosanMethyl

CellulosePectin

a. Initial

b. Oral

c. Gastric

d. Intestinal

Control

Page 110: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

101

Figure 4.4. Influence of the concentration (mass ratio polysaccharide (P):corn oil (CO)) of chitosan,

methyl cellulose, and pectin on creaming stability of emulsions under simulated gastrointestinal conditions

consisting of an initial (a), oral (b), gastric (c), and intestinal (d) phases.

depletion flocculation (Biggs, Habgood, Jameson, & Yan, 2000; Renault, Sancey, Badot, &

Crini, 2009). However, the fact that much less flocculation occurred within the sample containing

chitosan suggests that neither depletion nor bridging effects were particularly strong (Furusawa,

Ueda, & Nashima, 1999). Bridging flocculation may have been limited due to the relatively weak

electrostatic interactions at this pH (Biggs, Habgood, Jameson, & Yan, 2000), whereas depletion

flocculation may have been limited because of the relatively low molecular weight of the

chitosan used (Li & McClements, 2013; Renault, Sancey, Badot, & Crini, 2009; Yuan, Gao,

Decker, & McClements, 2013).

ChitosanMethyl

CellulosePectin

a. Initial

b. Oral

c. Gastric

d. Intestinal

0.0 0.1 0.6 1.2 1.8 0.0 0.1 0.6 1.2 1.8 0.0 0.1 0.6 1.2 1.8Mass Ratio (P:CO)

Page 111: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

102

Measurements of the creaming stability of the initial emulsions in the presence of the different

polysaccharides also supported the observation that flocculation occurred in some of the samples

(Figure 4.4). In the absence of dietary fiber, the emulsions appeared homogenous after storage

and could therefore be considered to be stable to creaming. The initial emulsions containing

chitosan were stable to creaming at all dietary fiber concentrations studied, which suggests that

extensive droplet flocculation did not occur. The stability of the chitosan emulsions could be

attributed to an increase in aqueous phase viscosity, since the chitosan gave a larger increase than

the pectin (Figure 4.5). The emulsions containing methyl cellulose and pectin were stable to

creaming at low levels (0.4% w/w), but highly susceptible to creaming at higher levels (Figure

4.4). At low levels of these polysaccharides, the depletion attraction is not strong enough to

overcome the steric and/or electrostatic repulsion between the oil droplets and therefore

flocculation does not occur. At higher polysaccharide levels, the depletion attraction is strong

enough to promote flocculation and therefore rapid creaming occurs because of the resulting

increase in particle size (McClements, 2000). The thickness of the cream layer increases at high

polysaccharide levels because of the formation of a three-dimensional network of strongly

aggregated droplets that inhibits their movement (Mao & McClements, 2012). Viscosity

measurements of the samples containing high levels of the polysaccharides indicated that they

were relatively viscous, and could therefore inhibit droplet creaming (Kawakatsu, Trägårdh, &

Trägårdh, 2001) (Figure 4.5).

Measurements of the electrical charge in the emulsion-polysaccharide systems showed that there

was little change in the -potential when methyl cellulose or chitosan was added, but that there

was an appreciable increase in the negative charge when pectin was added (Figure 4.6). These

results suggest that methyl cellulose and chitosan did not strongly interact with the emulsion

droplets, which can be attributed to the relatively low charge of the fat droplets and

polysaccharides at this pH. The large increase in negative charge that occurred when pectin was

added can probably be attributed to the fact that the micro-electrophoresis instrument measured

the electrical characteristics of the pectin molecules rather than those of the fat droplets (Tsai,

Chen, Kuo, Lin, Wang, Hsien, et al., 2014).

Page 112: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

103

Figure 4.5. Influence of the concentration of chitosan, methyl cellulose, and pectin on the apparent

viscosity of emulsions under simulated gastrointestinal conditions consisting of an initial (a), oral (b),

gastric (c), and intestinal (d) phases. Apparent viscosity () was obtained from measurements with a shear

rate of 10 s-1

.

4.3.2.2. Oral phase

The emulsion samples were then subjected to a simulated oral phase, and their physicochemical

and structural properties were measured. The particle size distribution measured by SLS indicated

that the majority of fat droplets in all of the emulsions remained relatively small, but that there

was a population of highly aggregated droplets in all of the systems (Figure 4.2b).

0.00

0.25

0.50

0.75

1.00

1.25

1.50

0.0 0.2 0.4 0.6 0.8 1.0

Ap

pa

ren

t V

isco

sity

(P

a s

)

Polysaccharide (% w/w)

Chitosan

Methyl Cellulose

Pectin

c.

0.000

0.005

0.010

0.015

0.020

0.025

0.0 0.2 0.4 0.6 0.8 1.0

Ap

pa

ren

t V

isco

sity

(P

a s

)

Polysaccharide (% w/w)

Chitosan

Methyl Cellulose

Pectin

d.

0

100

200

300

400

500

600

700

0.0 1.0 2.0 3.0 4.0

Ap

pa

ren

t V

isco

sity

(P

a s

)

Polysaccharide (% w/w)

Chitosan

Methyl Cellulose

Pectin

a.

0

10

20

30

40

50

60

70

0.0 0.5 1.0 1.5 2.0

Ap

pa

ren

t V

isco

sity

(P

a s

)

Polysaccharide (% w/w)

Chitosan

Methyl Cellulose

Pectin

b.

Page 113: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

104

Figure 4.6. Influence of the concentration of chitosan, methyl cellulose, and pectin on the surface

electrical charge (-potential) of emulsions under simulated gastrointestinal conditions consisting of an

initial (a), oral (b), gastric (c), and intestinal (d) phases.

The confocal microscopy images confirmed that large flocs were present in all of the emulsions

containing polysaccharides, but that there were also some smaller flocs in the control emulsion

containing no dietary fiber (Figure 4.3). Visual observations indicated that all the emulsions

were highly unstable to gravitational separation: after storage they all had a thin white layer of fat

droplets at the top and a watery serum layer at the bottom (Figure 4.4). These results suggest that

the conditions in the oral phase promoted extensive droplet flocculation in all of the emulsions. In

-60

-40

-20

0

20

40

60

0.0 1.0 2.0 3.0 4.0

-P

ote

nti

al

(mV

)

Polysaccharide (% w/w)

Chitosan

Methyl Cellulose

Pectin

a.

-60

-40

-20

0

20

40

60

0.0 0.5 1.0 1.5 2.0

-P

ote

nti

al

(mV

)

Polysaccharide (% w/w)

Chitosan

Methyl Cellulose

Pectin

b.

-60

-40

-20

0

20

40

60

0.0 0.2 0.4 0.6 0.8 1.0

-P

ote

nti

al

(mV

)

Polysaccharide (% w/w)

Chitosan

Methyl Cellulose

Pectin

c.

-60

-40

-20

0

20

40

60

0.0 0.2 0.4 0.6 0.8

-P

ote

nti

al

(mV

)

Polysaccharide (% w/w)

Chitosan

Methyl Cellulose

Pectin

d.

Page 114: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

105

the control emulsion, droplet aggregation can be attributed to depletion flocculation induced by

the presence of the mucin molecules in the simulated oral fluids (Vingerhoeds, Blijdenstein, Zoet,

& van Aken, 2005). In the other emulsions, droplet flocculation may have been a result of

depletion and bridging flocculation caused by the mucin and dietary fiber molecules

(McClements, 2000; Vingerhoeds, Blijdenstein, Zoet, & van Aken, 2005). The presence of mucin

would have increased the osmotic attraction between the fat droplets due to the presence of non-

adsorbed polysaccharides in the aqueous phase. In addition, there may have been some

electrostatic attraction between the anionic mucin and cationic chitosan in the emulsions

containing chitosan (Svensson, Thuresson, & Arnebrant, 2008).

Similar to the initial samples (Section 4.3.2.1), measurements of the electrical charge

characteristics of the emulsion-polysaccharide systems showed that there was little change in the

-potential when methyl cellulose or chitosan was added, but that there was a large increase in

negative charge when pectin was added (Figure 4.6b). Again, these results suggest that methyl

cellulose and chitosan did not strongly interact with the fat droplets under oral conditions, which

can be attributed to the relatively low charge of the fat droplets (-12.0 mV) and these two

polysaccharides (Figure 4.1) at this pH. The large increase in negative charge that was observed

when pectin was added to the emulsions can again be attributed to the fact that the micro-

electrophoresis instrument was more sensitive to the pectin molecules than the fat droplets

(Tholstrup Sejersen, Salomonsen, Ipsen, Clark, Rolin, & Balling Engelsen, 2007).

Shear viscosity measurements indicated that all of the samples containing polysaccharides were

relatively viscous after exposure to oral conditions (Figure 4.5, >1 Pa s). The increase in

viscosity in the presence of the polysaccharides depended on dietary fiber type: methyl

cellulose>chitosan>pectin. These differences can be attributed to differences in the molecular

characteristics of the dietary fibers, such as molecular weight, conformation, and molecular

interactions. In general, the apparent viscosity of a polymer solution increases with increasing

molecular weight, decreasing branching, and increasing interactions (Dunstan, Chai, Lee, &

Boger, 1995).

Page 115: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

106

4.3.2.3. Gastric phase

After passage through the oral phase, the samples were subjected to a simulated gastric phase,

and again changes in their physicochemical and structural properties were measured. Both the

light scattering and confocal microscopy measurements indicated that extensive droplet

aggregation occurred in all of the systems (Figures 4.2c and 4.3). The irregular shape of the

particles observed in the confocal microscopy images suggested that the droplets were

flocculated, rather than coalesced under gastric conditions. Visual observations indicated that all

the control and chitosan emulsions were relatively stable to gravitational separation: after storage

they had a fairly uniform cloudy appearance throughout (Figure 4.4). On the other hand, the

emulsions containing methyl cellulose or pectin had white sediments at the bottom of the test

tubes after exposure to the gastric phase (Figure 4.4). The amount of sediment present in these

samples increased as the polysaccharide concentration increased. These results suggest that the

flocs formed by these two polysaccharides in the simulated gastric fluids were large and dense

enough to rapidly sediment. Furthermore, the flocs formed in the control and chitosan emulsions

did not appear to be strongly susceptible to gravitational separation, perhaps because of their

smaller size or lower density contrast (McClements, 2000).

Electrical charge measurements of the emulsion-polysaccharide systems under gastric conditions

showed that there was little change in -potential when methyl cellulose or pectin was added, but

that there was a large increase in positive charge when chitosan was added (Figure 4.6c). These

results suggest that methyl cellulose and pectin did not strongly interact with the fat droplets

through electrostatic interactions under gastric conditions, which can be attributed to the

relatively low charge of the fat droplets (-1.0 mV) and these polysaccharides (Figure 4.1) at pH

3.0. The large increase in positive charge that occurred when chitosan was added to the emulsions

can be attributed to the fact that the chitosan molecules became strongly cationic under acidic

conditions (Figure 4.1). The measured positive charge may therefore have been indicative of

interactions between the fat droplets and chitosan (Yuan, Gao, Decker, & McClements, 2013), or

due to the fact that the micro-electrophoresis instrument was more sensitive to the chitosan

molecules than the fat droplets. The viscosity of all the emulsions was relatively low under

Page 116: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

107

simulated gastric conditions, which can be attributed to the fact that the samples were diluted at

each stage of the gastrointestinal tract model so the polymer concentration would be relatively

low, i.e., below the polymer overlap region(Simo, Mao, Tokle, Decker, & McClements, 2012).

4.3.2.4. Intestinal phase

After passage through the gastric phase, the samples were subjected to a simulated small intestine

phase, and changes in their physicochemical and structural properties were again measured. Light

scattering and confocal microscopy measurements suggested that extensive droplet aggregation

occurred in all of the systems, but that there were distinct differences between their

microstructures (Figures 4.2d and 4.3). The fat phase was fairly evenly distributed throughout

the sample in the control emulsion containing no polysaccharide (Figure 4.3) and many small

particles were detected by SLS (Figure 4.2d). Presumably, the majority of these particles were

mixed micelles formed by the lipid digestion process (Hur, Lim, Decker, & McClements, 2011;

McClements & Li, 2010a; Minekus, et al., 2014). Mixed micelles consist of small (<10 nm)

micelle structures, as well as much larger (50-5000 nm) liposome structures (McClements & Li,

2010b). They consist of phospholipids and bile salts from the intestinal fluids, as well as free

fatty acids and monoacylglycerols resulting from digestion of the triacylglycerols (Almgren,

2000; Yang & McClements, 2013).

The mixed emulsions containing chitosan contained some irregular shaped particles, but these

were appreciably smaller than those observed in the mixed emulsions containing either pectin or

methyl cellulose (Figure 4.3). The particles in these systems were probably a mixture of

undigested fat droplets and mixed micelles. Visual observations indicated that the control

emulsions and the emulsions containing chitosan had a relatively uniform yellowish brown

appearance (Figure 4.4). The emulsions containing methyl cellulose or pectin also had a

yellowish brown color but there was evidence of some sediment at the bottom of the test tubes

after exposure to the intestinal phase (Figure 4.4). The brownish yellow color can be attributed to

the presence of bile salts, since the stock solution of these digestive components had a dark

brown color.

Page 117: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

108

Electrical charge measurements indicated that the control emulsions had a relatively high

negative charge (-35.0 mV) under simulated intestinal conditions (Figure 4.6d), which can be

attributed to the presence of anionic substances at the particle surfaces, such as free fatty acids,

phospholipids, and bile salts. The -potential changed appreciably with increasing polysaccharide

concentration, with the direction of the change depending on initial polysaccharide type. The

particles became more positive when chitosan was added, more negative when pectin was added,

and changed little when methyl cellulose was added (Figure 4.6d). These results suggest that

methyl cellulose did not strongly interact with the fat droplets through electrostatic interactions

under intestinal conditions, which can be attributed to its neutral charge characteristics (Figure

4.1) at pH 7.0. On the other hand, the increase in positive charge on the particles when chitosan

was added to the emulsions may have been due to the fact that cationic chitosan molecules

interacted with the anionic lipid particles. The increase in negative charged when increasing

amounts of pectin were added may have been due to binding of pectin to the negative lipid

particles, but this is unlikely due to strong electrostatic repulsion between them (Beysseriat,

Decker, & McClements, 2006; Simo, Mao, Tokle, Decker, & McClements, 2012). Instead, the

micro-electrophoresis instrument may have been more sensitive to the pectin molecules than the

lipid particles. In addition, the viscosity of all the emulsions was relatively low under simulated

intestinal conditions (Figure 4.5d), which can be attributed to the progressive dilution that occurs

after passage through each stage of the gastrointestinal model (Hur, Lim, Decker, & McClements,

2011; McClements & Li, 2010a; Minekus, et al., 2014). Finally, we examined the influence of

polysaccharide type and concentration on the rate and extent of lipid digestion using a pH stat

method (Figure 4.7). In the absence of polysaccharide, the emulsions were rapidly and

completely digested. Indeed, the fat phase was almost fully digested within the first 5 minutes of

incubation. In the presence of polysaccharides, there was a decrease in both the rate and extent of

lipid digestion (principally the extent rather than rate), with the extent of the digestion process

depending on the polysaccharide concentration. In addition, there was a slight decrease in the

total amount of fatty acids produced after 2 hours of digestion with increasing chitosan

concentration, but a much more appreciable decrease with increasing pectin or methyl cellulose

concentration (Figure 4.8). These results suggest that both pectin and methyl cellulose were able

to appreciably inhibit lipid digestion.

Page 118: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

109

Figure 4.7. Influence of the concentration of chitosan (a), methyl cellulose (b), and pectin (c) on in vitro

hydrolysis (percentage of free fatty acids (FFA) released by the pH stat method) of lipid droplets (0.5 %

w/w) under simulated gastrointestinal conditions.

0

20

40

60

80

100

0 20 40 60 80 100 120

FF

A R

elea

sed

(%

w/w

)

Digestion Time (min)

0.00%

0.07%

0.15%

0.44%

0.65%

0

20

40

60

80

100

0 20 40 60 80 100 120

FF

A R

elea

sed

(%

w/w

)

Digestion Time (min)

0.00%

0.07%

0.15%

0.44%

0.65%

0

20

40

60

80

100

0 20 40 60 80 100 120

FF

A R

elea

sed

(%

w/w

)

Digestion Time (min)

0.00%

0.07%

0.15%

0.44%

0.65%

c.

b.

a.

Page 119: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

110

Figure 4.8. Influence of the concentration of chitosan, methyl cellulose, and pectin on free fatty acids

(FFA) released after 2 hours of digestion (intestinal phase).

Based on the confocal microscopy images (Figure 4.3), the creaming stability of the emulsions

throughout the gastrointestinal digestion (Figure 4.4), and the aggregation state of the lipid

droplets (Figure 4.2), the most likely mechanism for this phenomenon is the ability of these

polysaccharides to promote extensive droplet flocculation by depletion attraction. Presumably,

the lipid droplets are trapped within large flocs that reduce the ability of the lipase molecules to

interact with the fat droplet surfaces (by steric hindrance) and digest the lipids (Figures 4.7 and

4.8). One would expect that the inhibition of lipid digestion would increase as the floc size

increased, and as the packing of the droplets and polymers within the flocs increased.

4.3.3. Potential mechanisms

Overall, this study has shown that different polysaccharides have different effects on the extent of

lipid digestion. In particular, our results suggest that both pectin and methyl cellulose were able

to appreciably inhibit lipid digestion. In this section, we examine some potential mechanisms that

may account for the observed influence of polysaccharides on lipid digestion.

0

20

40

60

80

100

0.0 0.2 0.4 0.6 0.8

Fin

al

Dig

esti

on

(%

FF

A)

Polysaccharide (% w/w)

Chitosan

Methyl Cellulose

Pectin

Page 120: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

111

4.3.3.1. Rheology

The type and amount of polysaccharides present in the initial system influenced the rheological

properties of the fluids in the various stages of the simulated gastrointestinal tract (Figure 4.5).

Changes in the rheology of the gastrointestinal fluids may impact the rate and extent of lipid

digestion through a number of mechanisms. At the molecular level, an increase in the micro-

viscosity of a sample will slow down the movement of any molecular species involved in the

lipid digestion process, e.g., bile salts and lipase towards the droplet surfaces, or free fatty acids

and monoacylglycerols away from the droplet surfaces. Consequently, dietary fibers could

decrease the rate and extent of lipid digestion due to their ability to slow down molecular

diffusion. However, it should be stressed that polysaccharides may cause a large increase in the

macro-viscosity of a sample, but have little effect on the micro-viscosity since small molecules

can easily diffuse through the large pores in polymer networks. An increase in the macro-

viscosity associated with the presence of dietary fibers may influence the intimate mixing of the

samples with the digestive components, which could also inhibit the ability of lipase to get the

lipid droplets surfaces. In the small intestine phase, the increase in apparent viscosity due to the

presence of the different polysaccharides was relatively modest (Figure 4.5), and therefore we do

not believe that this mechanism played a major role in influencing lipid digestion.

4.3.3.2. Flocculation

The presence of polysaccharides within the gastrointestinal fluids may have promoted

flocculation of the lipid droplets due to bridging, depletion, or other flocculation mechanism. The

ability of lipase to interact with the lipid droplet surfaces and digest the encapsulated triglycerides

may be reduced if the droplets trapped within the large flocs (Figure 4.9). One would expect that

the inhibition of lipid digestion would increase as the floc size increased, and as the packing of

droplets and polymers within the flocs increased, since these factors would reduce the ability of

lipase molecules to rapidly diffuse through the entire flocs. Based on our confocal microscopy

images (Figure 4.3) and other measurements, this mechanism appears to be important in

accounting for the observed inhibition of lipid digestion, since the emulsions

Page 121: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

112

Figure 4.9. Schematic representation of the inhibition of lipid droplets digestion extent upon addition of

polysaccharides. Lipid droplets of the corn oil-in-water emulsion stabilized by Tween 80 (a), digestion of

lipid droplets by lipase (b), polysaccharides may lead a decrease on the digestion extent of lipid droplets

by embedding them into their structure (c).

containing methyl cellulose and pectin were highly flocculated (Figure 4.3) and also had reduced

digestion rates (Figure 4.8).

4.3.3.3. Electrostatic interactions

One would expect cationic chitosan molecules to interact with various anionic species involved in

the lipid digestion process, such as lipid droplets, bile salts, phospholipids, free fatty acids, and

mixed micelles. These interactions may either inhibit or promote lipid digestion depending on

their nature. For example, chitosan may bind free fatty acids produced during triglyceride

lipolysis and remove them from the lipid droplet surfaces, thereby allowing the lipase to continue

acting on the non-digested triglycerides. On the other hand, if chitosan forms a protective layer

around the lipid droplet surfaces, then it may inhibit lipid digestion by preventing the lipase from

reaching the non-digested triglycerides within the lipid droplets. One would also expect anionic

pectin molecules to interact with any cationic species involved in the lipid digestion process. For

example, anionic pectin may strongly bind calcium ions and prevent them for precipitating long-

chain fatty acids at the lipid droplet surfaces. As a result, lipid digestion may be inhibited because

a. b. c.

Page 122: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

113

the formation of a layer of free fatty acids around the lipid droplets can prevent lipase from

reaching the non-digested triglycerides. Electrostatic interactions may therefore also play an

important role in the ability of certain polysaccharides to inhibit lipid digestion. In future studies,

it would be useful to carry out a more detailed study of the interactions of dietary fibers with

specific digestive components so as to better understand the potential importance of these

interactions.

Finally, through this study it was possible to obtain a better understanding of the role of dietary

fiber characteristics (viscosity and surface electrical charge) on the gastrointestinal fate of

ingested lipids. The knowledge gained from this study might be useful for the design,

development, and fabrication of healthier functional food products specifically engineered for

controlling obesity and for promoting health and wellness of the consumers.

4.4. Conclusions

The objective of this work was to study the impact of three polysaccharides (chitosan, methyl

cellulose, and pectin) on the physicochemical characteristics and microstructure of emulsified

lipids during passage through a simulated gastrointestinal tract. Pectin and methyl cellulose

promoted depletion flocculation when present at sufficiently high concentrations, whereas

chitosan promoted bridging flocculation under acidic pH conditions. Pectin and methyl cellulose

reduced the extent of lipid digestion appreciably, whereas chitosan caused a slight decrease.

These results have important implications for understanding the influence of dietary fibers on

lipid digestion, since they promote droplet flocculation and therefore inhibit digestion. Our

results suggest that droplet flocculation may have restricted the access of lipase to the fat droplet

surfaces, thereby reducing the magnitude of the hydrolysis of the emulsified lipids (Figure 4.9).

In addition, electrostatic interactions of polysaccharides with oppositely charged species involved

in lipid digestion may also impact the digestion process. This information may be used for

designing functional foods that give healthier lipid profiles and thereby promote health and

wellness of the consumers.

Page 123: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

114

Acknowledgments

The authors are grateful to COLCIENCIAS and Universidad Nacional de Colombia for providing

a fellowship to Mauricio Espinal-Ruiz supporting this work. We also thank the United States

Department of Agriculture (NIFA Program) for supporting this research.

References

Almgren, M. (2000). Mixed micelles and other structures in the solubilization of bilayer lipid membranes

by surfactants. Biochimica et Biophysica Acta (BBA) - Biomembranes, 1508(1–2), 146-163.

Beysseriat, M., Decker, E. A., & McClements, D. J. (2006). Preliminary study of the influence of dietary

fiber on the properties of oil-in-water emulsions passing through an in vitro human digestion

model. Food Hydrocolloids, 20(6), 800-809.

Biggs, S., Habgood, M., Jameson, G. J., & Yan, Y.-d. (2000). Aggregate structures formed via a bridging

flocculation mechanism. Chemical Engineering Journal, 80(1–3), 13-22.

Bray, G. A., Paeratakul, S., & Popkin, B. M. (2004). Dietary fat and obesity: a review of animal, clinical

and epidemiological studies. Physiology & Behavior, 83(4), 549-555.

Brownlee, I. A. (2011). The physiological roles of dietary fibre. Food Hydrocolloids, 25(2), 238-250.

Castillo, M., Martín-Orúe, S. M., Anguita, M., Pérez, J. F., & Gasa, J. (2007). Adaptation of gut

microbiota to corn physical structure and different types of dietary fibre. Livestock Science,

109(1–3), 149-152.

Chung, C., Degner, B., & McClements, D. J. (2013). Designing reduced-fat food emulsions: Locust bean

gum–fat droplet interactions. Food Hydrocolloids, 32(2), 263-270.

Dongowski, G. (2007). Interactions between dietary fibre-rich preparations and glycoconjugated bile acids

in vitro. Food Chemistry, 104(1), 390-397.

Dunstan, D. E., Chai, E., Lee, M., & Boger, D. V. (1995). The rheology of engineered polysaccharides.

Food Hydrocolloids, 9(4), 225-228.

Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L.-P., & Narváez-Cuenca, C.-E. (2014).

Inhibition of digestive enzyme activities by pectic polysaccharides in model solutions. Bioactive

Carbohydrates and Dietary Fibre, 4(1), 27-38.

Fabek, H., Messerschmidt, S., Brulport, V., & Goff, H. D. (2014). The effect of in vitro digestive

processes on the viscosity of dietary fibres and their influence on glucose diffusion. Food

Hydrocolloids, 35(0), 718-726.

Furusawa, K., Ueda, M., & Nashima, T. (1999). Bridging and depletion flocculation of synthetic latices

induced by polyelectrolytes. Colloids and Surfaces, A: Physicochemical and Engineering Aspects,

153(1–3), 575-581.

Galisteo, M., Duarte, J., & Zarzuelo, A. (2008). Effects of dietary fibers on disturbances clustered in the

metabolic syndrome. The Journal of Nutritional Biochemistry, 19(2), 71-84.

Ghosh, S., & Rousseau, D. (2011). Fat crystals and water-in-oil emulsion stability. Current Opinion in

Colloid & Interface Science, 16(5), 421-431.

Ghotra, B. S., Dyal, S. D., & Narine, S. S. (2002). Lipid shortenings: a review. Food Research

International, 35(10), 1015-1048.

Page 124: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

115

Gunness, P., & Gidley, M. J. (2010). Mechanisms underlying the cholesterol-lowering properties of

soluble dietary fibre polysaccharides. Food & Function, 1(2), 149-155.

Heertje, I. (2014). Structure and function of food products: A review. Food Structure, 1(1), 3-23.

Hur, S. J., Lim, B. O., Decker, E. A., & McClements, D. J. (2011). In vitro human digestion models for

food applications. Food Chemistry, 125(1), 1-12.

Jenkins, P., & Snowden, M. (1996). Depletion flocculation in colloidal dispersions. Advances in Colloid

and Interface Science, 68(0), 57-96.

Jones, O. G., Lesmes, U., Dubin, P., & McClements, D. J. (2010). Effect of polysaccharide charge on

formation and properties of biopolymer nanoparticles created by heat treatment of β-

lactoglobulin–pectin complexes. Food Hydrocolloids, 24(4), 374-383.

Kawakatsu, T., Trägårdh, G., & Trägårdh, C. (2001). The formation of oil droplets in a pectin solution and

the viscosity of the oil-in-pectin solution emulsion. Journal of Food Engineering, 50(4), 247-254.

Khan, R. S., Grigor, J., Winger, R., & Win, A. (2013). Functional food product development –

Opportunities and challenges for food manufacturers. Trends in Food Science & Technology,

30(1), 27-37.

Kristensen, M., & Jensen, M. G. (2011). Dietary fibres in the regulation of appetite and food intake.

Importance of viscosity. Appetite, 56(1), 65-70.

Kritchevsky, D. (1988). Dietary Fiber. Annual Review of Nutrition, 8(1), 301-328.

Li, Y., Hu, M., & McClements, D. J. (2011). Factors affecting lipase digestibility of emulsified lipids

using an in vitro digestion model: Proposal for a standardised pH-stat method. Food Chemistry,

126(2), 498-505.

Li, Y., & McClements, D. J. (2013). Influence of non-ionic surfactant on electrostatic complexation of

protein-coated oil droplets and ionic biopolymers (alginate and chitosan). Food Hydrocolloids,

33(2), 368-375.

Li, Y., & McClements, D. J. (2014). Modulating lipid droplet intestinal lipolysis by electrostatic

complexation with anionic polysaccharides: Influence of cosurfactants. Food Hydrocolloids,

35(0), 367-374.

Mao, Y., & Julian McClements, D. (2012). Fabrication of Reduced Fat Products by Controlled

Heteroaggregation of Oppositely Charged Lipid Droplets. Journal of Food Science, 77(5), E144-

E152.

Mao, Y., & McClements, D. J. (2012). Influence of electrostatic heteroaggregation of lipid droplets on

their stability and digestibility under simulated gastrointestinal conditions. Food & Function,

3(10), 1025-1034.

McClements, D. J. (2000). Comments on viscosity enhancement and depletion flocculation by

polysaccharides. Food Hydrocolloids, 14(2), 173-177.

McClements, D. J., & Li, Y. (2010a). Review of in vitro digestion models for rapid screening of

emulsion-based systems. Food & Function, 1(1), 32-59.

McClements, D. J., & Li, Y. (2010b). Structured emulsion-based delivery systems: Controlling the

digestion and release of lipophilic food components. Advances in Colloid and Interface Science,

159(2), 213-228.

Michas, G., Micha, R., & Zampelas, A. (2014). Dietary fats and cardiovascular disease: Putting together

the pieces of a complicated puzzle. Atherosclerosis, 234(2), 320-328.

Minekus, M., Alminger, M., Alvito, P., Ballance, S., Bohn, T., Bourlieu, C., Carriere, F., Boutrou, R.,

Corredig, M., Dupont, D., Dufour, C., Egger, L., Golding, M., Karakaya, S., Kirkhus, B., Le

Feunteun, S., Lesmes, U., Macierzanka, A., Mackie, A., Marze, S., McClements, D. J., Menard,

O., Recio, I., Santos, C. N., Singh, R. P., Vegarud, G. E., Wickham, M. S. J., Weitschies, W., &

Brodkorb, A. (2014). A standardised static in vitro digestion method suitable for food - an

international consensus. Food & Function, 5(6), 1113-1124.

Page 125: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 4

116

Mudgil, D., & Barak, S. (2013). Composition, properties and health benefits of indigestible carbohydrate

polymers as dietary fiber: A review. International Journal of Biological Macromolecules, 61(0),

1-6.

Narine, S. S., & Marangoni, A. G. (1999). Relating structure of fat crystal networks to mechanical

properties: a review. Food Research International, 32(4), 227-248.

Nikiforidis, C. V., & Kiosseoglou, V. (2011). Competitive displacement of oil body surface proteins by

Tween 80 – Effect on physical stability. Food Hydrocolloids, 25(5), 1063-1068.

Renault, F., Sancey, B., Badot, P. M., & Crini, G. (2009). Chitosan for coagulation/flocculation processes

– An eco-friendly approach. European Polymer Journal, 45(5), 1337-1348.

Rousseau, D. (2000). Fat crystals and emulsion stability — a review. Food Research International, 33(1),

3-14.

Simo, O. K., Mao, Y., Tokle, T., Decker, E. A., & McClements, D. J. (2012). Novel strategies for

fabricating reduced fat foods: Heteroaggregation of lipid droplets with polysaccharides. Food

Research International, 48(2), 337-345.

Slavin, J. L. (2005). Dietary fiber and body weight. Nutrition, 21(3), 411-418.

Svensson, O., Thuresson, K., & Arnebrant, T. (2008). Interactions between chitosan-modified particles

and mucin-coated surfaces. Journal of Colloid and Interface Science, 325(2), 346-350.

Tholstrup Sejersen, M., Salomonsen, T., Ipsen, R., Clark, R., Rolin, C., & Balling Engelsen, S. (2007).

Zeta potential of pectin-stabilised casein aggregates in acidified milk drinks. International Dairy

Journal, 17(4), 302-307.

Tokle, T., Lesmes, U., & McClements, D. J. (2010). Impact of Electrostatic Deposition of Anionic

Polysaccharides on the Stability of Oil Droplets Coated by Lactoferrin. Journal of Agricultural

and Food Chemistry, 58(17), 9825-9832.

Tsai, R.-Y., Chen, P.-W., Kuo, T.-Y., Lin, C.-M., Wang, D.-M., Hsien, T.-Y., & Hsieh, H.-J. (2014).

Chitosan/pectin/gum Arabic polyelectrolyte complex: Process-dependent appearance,

microstructure analysis and its application. Carbohydrate Polymers, 101(0), 752-759.

van Dam, R. M., & Seidell, J. C. (2007). Carbohydrate intake and obesity. European Journal of Clinical

Nutrition, 61(S1), S75-S99.

Vingerhoeds, M. H., Blijdenstein, T. B. J., Zoet, F. D., & van Aken, G. A. (2005). Emulsion flocculation

induced by saliva and mucin. Food Hydrocolloids, 19(5), 915-922.

Wu, B.-c., Degner, B., & McClements, D. J. (2013). Creation of reduced fat foods: Influence of calcium-

induced droplet aggregation on microstructure and rheology of mixed food dispersions. Food

Chemistry, 141(4), 3393-3401.

Yang, Y., & McClements, D. J. (2013). Vitamin E and Vitamin E acetate solubilization in mixed micelles:

Physicochemical basis of bioaccessibility. Journal of Colloid and Interface Science, 405(0), 312-

321.

Yuan, F., Gao, Y., Decker, E. A., & McClements, D. J. (2013). Modulation of physicochemical properties

of emulsified lipids by chitosan addition. Journal of Food Engineering, 114(1), 1-7.

Page 126: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

Impact of pectin properties on lipid digestion under simulated gastrointestinal

conditions: Comparison of citrus and banana passion fruit

(Passiflora tripartita var. mollissima) pectins

Published as:

Espinal-Ruiz, M., Restrepo-Sánchez, L. P., Narváez-Cuenca, C. E., & McClements D. J.

Food Hydrocolloids. 52 (2016): 329 – 342.

Page 127: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

118

Abstract

Medium methoxylated pectin (52% mol/mol, MMP) was isolated from banana passion fruit

(Passiflora tripartita var. mollisima) by hot acidic extraction. The impact of MMP on lipid

digestion was compared to that of commercial citrus pectins with high (71% mol/mol, HMP) and

low (30% mol/mol, LMP) methoxylation degree. A static in vitro digestion model was used to

elucidate the impact of pectin properties (methoxylation degree and molecular weight) on the

gastrointestinal fate of emulsified lipids. A 2.0% (w/w) oil-in-water emulsion stabilized with

0.2% (w/w) Tween 80 was prepared, mixed with 1.8% (w/w) pectin samples, and then subjected

to the static in vitro digestion model (37 °C): initial (pH 7.0); oral (pH 6.8, 10 min, mucin);

gastric (pH 2.5, 120 min, pepsin); and intestinal (pH 7.0, 120 min, bile salts, and pancreatic

lipase) phases. The impact of the three pectin samples on surface particle charge (-potential),

particle size distribution of lipid droplets, microstructure, rheology, and lipid digestion (free fatty

acids (FFAs) released) was determined. The rate and extent of lipid digestion decreased with

increasing simultaneously both the molecular weight and pectin methoxylation, with the FFAs

released after 120 min of intestinal digestion being 47, 70, and 91% (w/w) for HMP, MMP, and

LMP, respectively. These results have important implications for understanding the influence of

pectin on lipid digestion. The control of lipid digestibility within the gastrointestinal tract might

be important for the designing and development of novel functional foods to control bioactive

release or to modulate satiety.

Keywords: Pectin, Passiflora tripartita var. mollisima, emulsion, lipid digestion, gastrointestinal

tract, depletion flocculation.

Page 128: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

119

5.1. Introduction

Tropical fruits are a good source of bioactive agents suitable for utilization in the food,

pharmaceutical, and cosmetic industries (Schieber, Stintzing, & Carle, 2001). Certain bioactive

agents found in tropical fruits have been shown to inhibit cardiovascular diseases and some types

of cancer (Rufino, Alves, de Brito, Pérez-Jiménez, Saura-Calixto, & Mancini-Filho, 2010).

Banana passion fruit (Passiflora tripartita var. mollissima) may be a particularly good source of

bioactive agents because of its relatively high levels of phenolics, carotenoids, and dietary fibers

(Gil, Restrepo, Millán, Alzate, & Rojano, 2014), which are known to be beneficial to human

health and wellbeing (Wootton-Beard & Ryan, 2011). Previous studies have shown that dietary

fibers from fruits have a positive effect on the treatment of diseases such as hyperlipidemia,

coronary heart disease, and certain types of cancer (Kumar, Sinha, Makkar, de Boeck, & Becker,

2011). The major source of non-cellulosic dietary fiber in fruits are pectins (Voragen, Timmers,

Linssen, Schols, & Pilnik, 1983). Pectins are acidic hetero-polysaccharides composed mainly of

-(1,4) linked D-galacturonic acid (GalA) residues (Ridley, O'Neill, & Mohnen, 2001). The

carboxyl moieties of the GalA unit may be esterified with methanol, which alters the electrical

characteristics of the molecule. Overall, the degree and patterning of methoxylation, as well as

the molecular weight, are important parameters determining the functional attributes of different

pectins (Funami, Nakauma, Ishihara, Tanaka, Inoue, & Phillips, 2011).

Although is usually accepted that pectin cannot be digested by the human gastrointestinal tract

(GIT), it is possible to get some nutrients from pectins due to the presence of symbiotic bacteria

in the GIT. Some bacteria are able to produce a group of enzymes that break down pectin into

simple sugars (mainly galacturonic acid), which are then in turn fermented to create short chain

fatty acids that human cells can absorb and which can contribute as much as 10 percent of the

calories required by cells. It has been established that two species of gut bacteria in mammals

have evolved to break down some particular kinds of foods. Bacteroides ovatus and Bacteroides

thetaiotaomicron are able to break down hemicelluloses and pectins, as well as other complex

carbohydrates that human intestinal cells secrete as mucus. These bacteria are a crucial part of the

Page 129: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

120

bacterial cells that make up our gastrointestinal tract, so further understanding of the metabolism

of these gut bacteria could help to improve the influence of pectin on human nutrition (Inman,

2011).

Overconsumption of fat is a major contributing factor to obesity, cardiovascular disease, and

diabetes (Bray & Popkin, 1998). For this reason, there has been considerable interest in the

development of effective strategies to reduce the caloric content of foods, or to reduce the spike

in blood lipids that occurs after consuming a fatty meal. Several studies have suggested that

certain types of dietary fibers can inhibit the digestion and absorption of lipids (Beysseriat,

Decker, & McClements, 2006; Edashige, Murakami, & Tsujita, 2008; Tsujita, Sumiyoshi, Han,

Fujiwara, Tsujita, & Okuda, 2003; Yonekura & Nagao, 2009). Numerous physicochemical and

physiological mechanisms may contribute to this effect, including the ability of dietary fibers to

alter the rheology of the gastrointestinal fluids, to bind digestive components (such as bile salts

and digestive enzymes), to alter the aggregation state of lipid droplets, to form protective coatings

around lipid droplets, and to be fermented within the large intestine by colonic bacteria

(Grabitske & Slavin, 2009; Lattimer & Haub, 2010; McClements, Decker, & Park, 2009). In a

recent study, we showed that pectin reduced the rate and extent (principally the extent rather than

rate) of the digestion of emulsified lipids (Espinal-Ruiz, Parada-Alfonso, Restrepo-Sanchez,

Narvaez-Cuenca, & McClements, 2014). Increased consumption of pectin may therefore prove to

be one strategy of reducing the caloric content of fatty food products or of modulating blood lipid

levels (Mesbahi, Jamalian, & Farahnaky, 2005).

The lipids in food may be consumed in a wide variety of different physical structures such as oils

(edible oils), bulk fats (margarine and butter), or emulsified fats (milk, cream, soups, and sauces).

Nevertheless, most fatty foods are broken down into oil-in-water emulsions within mouth during

mastication and within the stomach and small intestine during the digestion process (McClements

& Li, 2010). Consequently, lipid digestion within the gastrointestinal tract typically involves

digestion of emulsified fats. Lipid digestion involves several sequential steps that include various

physicochemical and biochemical events (Torcello-Gomez, Maldonado-Valderrama, Martin-

Rodriguez, & McClements, 2011).

Page 130: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

121

In the mouth, an ingested food is mixed with saliva (pH 7.0), undergoes temperature changes

(T 37 °C), and is subjected to mechanical forces that may alter the structure, physical state, and

interfacial properties of the lipid phase (Li, Kim, Park, & McClements, 2012). In the stomach, the

lipids are mixed with a highly acidic aqueous solution that contains minerals, biopolymers,

surface active compounds, and digestive enzymes (Singh, Ye, & Horne, 2009). The lipid phase

may undergo further changes in structure due to droplet disruption and coalescence processes, as

well as changes in the nature and composition of the surface active materials adsorbed at the

lipid-water interface (Singh & Sarkar, 2011). In particular, gastric lipase adsorbs to the lipid-

water interface and initiates the lipid digestion process, converting some of the triacylglycerols

(TAGs) to diacylglycerols (DAGs), monoacylglycerols (MAGs), and free fatty acids (FFAs)

(Wilde & Chu, 2011). In the small intestine, the emulsified lipids are mixed with digestive juices

that contain pancreatic lipase, colipase, bile salts, and phospholipids (Golding & Wooster, 2010).

The bile salts and phospholipids compete and displace any surface active material present at the

lipid-water interface, and the lipase-colipase complex binds to the lipid droplet surfaces (Reis,

Holmberg, Watzke, Leser, & Miller, 2009). The pancreatic lipase converts TAGs into MAGs and

FFAs, which leave the lipid droplet surfaces and are incorporated into mixed micelle structures

consisting of phospholipids and bile salts, which then transport them to the epithelial cells, where

they are adsorbed (Yao, Xiao, & McClements, 2014).

In this study, a simulated in vitro gastrointestinal model was used to evaluate the impact of

commercial high (HMP) and low (LMP) methoxylated pectins from citrus, and medium

methoxylated pectin (MMP) isolated from banana passion fruit (Passiflora tripartita var.

mollisima) on the gastrointestinal fate of emulsified lipids. These three pectin samples were

selected because of their different charge (methoxylation) and size (molecular weight)

characteristics, and because they can be used as functional ingredients in food and beverage

products (Willats, Knox, & Mikkelsen, 2006). We hypothesized that these three pectins would

have different effects on lipid digestion due to their different molecular and physicochemical

characteristics. In particular, we focused on their influence on the rheology of the gastrointestinal

fluids, the aggregation stability of lipid droplets in different stages of the gastrointestinal tract,

and the rate and extent of lipid digestion. The aim of the study was to obtain a better

Page 131: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

122

understanding of the role of pectin characteristics on the gastrointestinal fate of ingested lipids.

The knowledge obtained in this study might be useful for the design, fabrication, and

implementation of pectin-based functional foods designed to promote health and wellness by

modulating lipid digestion.

5.2. Materials and methods

5.2.1. Chemicals

Corn oil was purchased from a commercial food supplier (Mazola, ACH Food Companies Inc.,

Memphis, TN, USA) and stored at 4 °C until use. The manufacturer reported that the corn oil

contained approximately 14, 29, and 57% (w/w) of saturated, monounsaturated, and

polyunsaturated fatty acids, respectively. Commercial powdered high methoxylated pectin (HMP,

Genu Citrus Pectin USP/100) was kindly donated by CP Kelco Co. (Lille Skensved, Denmark)

and was used without further purification. The methoxylation degree of this material was 71%

(mol/mol) and the average molecular weight 181 kDa. Commercial powdered low methoxylated

pectin (LMP) was kindly donated by TIC Gums Inc. (Belcamp, MD, USA) and was also used

without further purification. The methoxylation degree of this material was 30% (mol/mol) and

the average molecular weight 130 kDa. Fat soluble fluorescent dye Nile Red (N3013), lipase

from porcine pancreas (Type II, L3126, triacylglycerol hydrolase E.C. 3.1.1.3), bile extract

(porcine, B8631), mucin from porcine stomach (Type II, M2378, bound sialic acid ≤ 1.2%), and

pepsin A from porcine gastric mucose (P7000, endopeptidase E.C. 3.4.23.1, activity ≥ 250 units

mg-1

solid) were purchased from Sigma-Aldrich Chemical Company (St Louis, MO, USA). One

unit of activity of pepsin A will increase A280nm to 0.001 per min at pH 2.0 and 37 °C, using

hemoglobin as substrate. The supplier reported that the lipase activity at 37 °C was 100-400 units

mg-1

protein (pH 7.7 using olive oil) and 30-90 units mg-1

protein (pH 7.4 using triacetin) for 30

min incubation (one unit of activity of lipase corresponds to the release of 1 μeq of free fatty

acids). The composition of the bile extract has been reported as 49% (w/w) total bile salt (BS),

containing 10-15% glycodeoxycholic acid, 3-9% taurodeoxycholic acid, 0.5-7% deoxycholic

Page 132: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

123

acid, 1-5% hydrodeoxycholic acid, and 0.5-2% cholic acid; 5% (w/w) phosphatidyl choline (PC);

Ca2+ ≤ 0.06% (w/w); critical micelle concentration of bile extract 0.07 0.04 mM; and mole ratio

of BS to PC being around 15:1. Dextran analytical standards (25, 150, and 410 kDa) for high

performance size exclusion chromatography (HPSEC) were purchased from Sigma-Aldrich

Chemical Company (St Louis, MO, USA). All other chemicals were purchased from Sigma-

Aldrich Chemical Company (St Louis, MO, USA). Distilled water was used to prepare all

solutions.

5.2.2. Extraction of pectin from banana passion fruit and characterization of the pectin

samples (LMP, MMP, and HMP)

5.2.2.1. Extraction of pectin from banana passion fruit (MMP)

Five grams of homogenized banana passion fruit epicarp (Passiflora tripartita var. mollissima)

were mixed with 30 mL of 0.1 M HCl (pH 1.0) and stirred for 60 min at 90 °C. The mixture was

neutralized to pH 7.0 with 1 M NaOH solution and then 100 mL of 95% (v/v) ethanol were added

to induce pectin precipitation. The pectin obtained after 12 h of precipitation was filtered, washed

with 100 mL of 70% (v/v) ethanol, and then dried at 45 °C for 12 h. The extraction yield was

64% (w/w). Pectin samples (LMP, MMP, and HMP) were characterized as follows:

5.2.2.2. Molecular weight and gyration radius

The average molecular weight, the molecular weight distribution, and the gyration radius (rg)

were determined using high performance size exclusion chromatography (HPSEC), using a 1260

Infinity liquid chromatograph (Agilent Technologies Inc., Santa Clara, CA, USA). One-hundred

microliters of 0.5% (w/w) pectin samples were injected into a packed column (OHpak SB-806M

HQ, 8.0 mm × 300 mm, Shoko America Inc., Torrance, CA, USA) and the elution was performed

using 200 mM NaCl at a flow rate of 1 mL min-1

for 25 min at 20 °C. An Optilab T-rex

differential refractive index detector (Wyatt Technology Co., Santa Barbara, CA, USA) at 40 °C;

and a Dawn Heleos-II multi-angle laser light scattering detector (MALLS, Wyatt Technology

Page 133: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

124

Co., Santa Barbara, CA, USA) at 30 °C were used to monitor the eluents. Both multi-angle laser

light scattering signal at 90° and dextran analytical standards (25, 150, and 410 kDa; 0.5% w/w)

were used to estimate the average molecular weight, molecular weight distribution, and gyration

radius of the pectin samples. Although the multi-angle light scattering detector has 18 optimized

scattering angles, the scattered light signal at 90° was selected for the quantification of the

average molecular weight because it allows to obtain a reliable and accurate measure of the

scattered light (Yoo & Jane, 2002).

5.2.2.3. Fourier transform-infrared (FT-IR) spectra

Pectin samples were dried in a desiccator containing blue silica gel prior to FT-IR analysis. KBr-

pectin disc mixtures (90:10 w/w) were prepared and then the FT-IR spectra were collected at the

transmittance mode in a Nicolet iS10 FT-IR Spectrometer (Thermo Fisher Scientific, Waltham,

MA, USA) at a 4 cm-1

resolution. Eighty interferograms were measured to obtain a high signal to

noise ratio.

5.2.2.4. Methoxylation and acetylation degrees

The methoxylation and acetylation degrees of pectin samples were determined by ion exchange

chromatography (Voragen, Schols, & Pilnik, 1986). Pectin samples (30 mg) were suspended in 1

mL of an isopropanol-water mixture (1:1 v/v) containing 0.4 M NaOH and stored at room

temperature for 2 h. The suspension was centrifuged (20 min, 18000 g, 4 °C) and then 20 L of

the clear supernatant was injected into the column. A model LC-20AT liquid chromatograph

(Shimadzu Corporation, Kyoto, Japan) equipped with an Aminex HPX-87H column (300 × 7.8

mm × 9 m, Bio-Rad Laboratories, Hercules, CA, USA) was used. The column was operated at

room temperature and a flow rate of 0.6 mL min-1

with 4 mM H2SO4 as the eluent. Components

eluting from the column were detected using a RID-10A refractive index detector (Shimadzu

Corporation, Kyoto, Japan) thermostated at 40 °C. The amounts of methanol and acetic acid

released after saponification were determined using an external standard method. Calibration

lines were obtained at concentrations ranging from 5 to 40 mM, and from 0.1 to 0.8 mM for

Page 134: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

125

methanol and acetic acid, respectively. The methoxylation and acetylation degrees were

expressed as moles of methyl and acetyl esters, respectively, per 100 mol of uronic acid, and

were corrected for free methanol and acetic acid. The uronic acid content was determined

spectrophotometrically (van den Hoogen, van Weeren, Lopes-Cardozo, van Golde, Barneveld, &

van de Lest, 1998). Briefly, an aliquot of 400 L of each pectin sample solution (100 g mL-1

)

was mixed with 2 mL of 98% (w/w) H2SO4 containing 120 mM sodium tetraborate

(Na2B4O710H2O) and incubated for 60 min at 80 °C. After cooling down to room temperature,

the background absorbance of the samples was measured at 540 nm. Then, 400 L of m-

hydroxydiphenyl reagent (prepared by mixing 100 L of 100 mg mL-1

m-hydroxydiphenyl in

dimethyl sulfoxide with 4.9 mL of 80% (w/w) H2SO4) was added and mixed with the samples.

After 15 min, the absorbance of the pink-colored samples was measured at 540 nm. A calibration

line was obtained using GalA at final concentrations ranging from 0.1 to 1.0 g mL-1

.

5.2.3. Solutions and emulsions preparation

5.2.3.1. Pectin stock solutions

Pectin stock solutions (2.0% w/w) were prepared by dispersing 1 g of powdered pectins (LMP,

MMP, and HMP) into 49 g of 5 mM phosphate buffer (pH 7.0). The solutions were stirred at 800

rpm overnight at room temperature to ensure complete dispersion and dissolution. Stock

solutions were finally adjusted to pH 7.0 using 1 M NaOH solution.

5.2.3.2. Stock emulsion

A stock emulsion was prepared by mixing 20% (w/w) corn oil and 80% (w/w) buffered emulsifier

solution (5 mM phosphate buffer pH 7.0, containing 2.5% (w/w) Tween 80) together for 5 min

using a bio-homogenizer (Speed 2, Model MW140/2009-5, Biospec Products Inc., ESGC,

Switzerland). The coarse emulsion obtained was then passed 5 times through a high-pressure

homogenizer (Microfluidizer M-110L processor, Microfluidics Inc., Newton, MA, USA)

operating at 11,000 psi (75.8 MPa).

Page 135: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

126

5.2.3.3. Pectin-emulsion mixtures

Pectin-emulsion mixtures were prepared by mixing the stock emulsion containing 20% (w/w)

corn oil with buffered stock solutions of 2.0% (w/w) pectin (mass ratio 1:9), to obtain emulsions

containing 2.0% (w/w) corn oil and 1.8% (w/w) pectin. The pectin-emulsion mixtures were then

stirred with a high-speed stirrer (Fisher Steadfast Stirrer, Model SL 1200, Fisher Scientific Inc.,

Pittsburgh, PA, USA) at 800 rpm and stored overnight (approximately 12 h) at room temperature.

The pectin-emulsion mixtures were characterized to obtain the initial phase, prior to subjection to

the static in vitro digestion model.

5.2.4. Static in vitro digestion model

Each emulsion sample (initial phase) was passed through a simulated static in vitro digestion

model that consisted of oral, gastric, and intestinal phases. Measurements of emulsion

microstructure and stability, particle size distribution, particle charge, and viscosity were

performed after each phase. The standardized static in vitro digestion model used in this study

was a slight modification of that described previously (Espinal-Ruiz, Parada-Alfonso, Restrepo-

Sanchez, Narvaez-Cuenca, & McClements, 2014; Minekus, Alminger, Alvito, Ballance, Bohn,

Bourlieu, et al., 2014).

5.2.4.1. Oral phase

Simulated saliva fluid (SSF, pH 6.8) containing 3.0% (w/w) mucin was prepared according to the

composition shown in Table 5.1. Each emulsion (20 mL of initial phase) was mixed with 20 mL

of SSF and the resulting mixture containing 1.0% (w/w) corn oil and 0.9% (w/w) pectin was used

for characterization after the incubation period. The oral phase consisted of a flask containing

emulsion-SSF mixture incubated at 37 °C with continuous shaking at 100 rpm for 10 min in a

temperature controlled air incubator (Excella E24 Incubator Shaker, New Brunswick Scientific

Co., New Brunswick, NJ, USA) to mimic the conditions in the mouth. Although 10 min of

incubation time is somewhat longer than in vivo (approximately 1 min), however, accuracy and

Page 136: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

127

Table 5.1. Chemical composition of simulated saliva fluid (SSF) used to simulate oral conditions.

Compound Chemical formula Concentration (g L-1

)1

Sodium chloride NaCl 1.594

Ammonium nitrate NH4NO3 0.328

Potassium dihydrogen phosphate KH2PO4 0.636

Potassium chloride KCl 0.202

Potassium citrate K3C6H5O7•H2O 0.308

Uric acid sodium salt C5H3N4O3Na 0.021

Urea H2NCONH2 0.198

Lactic acid sodium salt C3H5O3Na 0.146

Porcine gastric mucin (Type II) ---- 30

1The SSF was prepared in double distilled water and then pH 6.8 was adjusted using 0.1 M NaOH.

reproducibility in a laboratory situation may be compromised if using any shorter digestion time.

In addition, it has been recommended an oral digestion time of 10 min in order to compensate the

lack of a proper mechanical action for static models, which in most cases is difficult to simulate

(Minekus, et al., 2014). The resulting oral phase (bolus) was used in the gastric phase.

5.2.4.2. Gastric phase

Simulated gastric fluid (SGF) was prepared by adding 2.0 g NaCl, 7.0 mL concentrated HCl

(37% w/w), and 3.2 g pepsin A (from porcine gastric mucose, 250 units mg-1

) to a flask and then

diluting with double distilled water to a volume of 1.0 L, and finally adjusting to pH 1.2 using 1

M HCl. Samples taken from the oral phase (20 mL bolus) were mixed with 20 mL of SGF so that

the final mixture contained 0.50% (w/w) corn oil and 0.45% (w/w) pectin. These mixtures were

then adjusted to pH 2.5 using 1 M NaOH and incubated at 37 °C with continuous shaking at 100

rpm for 2 h. Samples were taken for characterization at the end of the incubation period (gastric

phase). The resulting gastric phases (chyme) were used in the intestinal phase.

Page 137: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

128

5.2.4.3. Intestinal phase

Samples obtained from the gastric phase (30 mL chyme containing 0.50% (w/w) corn oil and

0.45% (w/w) pectin) were incubated for 2 h at 37 °C in a simulated small intestine fluid (SIF)

containing 2.5 mL pancreatic lipase (24 mg mL-1

), 3.5 mL bile extract solution (54 mg mL-1

), and

1.5 mL salt solution containing 0.25 M CaCl2 and 3.0 M NaCl, to obtain a final composition of

the intestinal fluid in the reaction vessel of 0.40% (w/w) corn oil and 0.36% (w/w) pectin. The

lipolytic reaction was conducted at constant (pH 7.0) using an automatic titration unit (pH stat

titration unit, 835 Titrando, Metrohm USA, Inc., Riverview, FL, USA) and then the FFAs

released were monitored by determining the amount of 0.1 M NaOH needed to maintain the

constant pH within the reaction vessel. All additives were dissolved in 5 mM phosphate buffer

solution (pH 7.0) before use. Lipase addition and initialization of the titration program were

carried out only after the addition of all pre-dissolved ingredients and balancing the pH to 7.0.

Samples were taken for physicochemical and structural characterization at the end of the

digestion period (intestinal phase). The volume of 0.1 M NaOH added to the emulsion was

recorded over time and then was used to calculate the concentration of FFAs generated by

lipolysis. The amount of FFAs released was calculated using the following equation:

(

) (5.1)

Here, VNaOH is the volume of NaOH (in L) titrated into the reaction vessel to neutralize the FFAs

released, assuming that all TAGs are hydrolyzed in two molecules of FFAs and one molecule of

MAG, CNaOH is the concentration of the sodium hydroxide (0.1 M), MWLipid is the average

molecular weight of corn oil (872 g mol-1

), and wLipid is the initial weight of corn oil in the

intestinal phase (0.15 g). Titration blanks were performed by inactivating pancreatic lipase

solution in boiling water for 15 min prior to initialization of the titration program.

Page 138: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

129

5.2.5. Emulsion characterization

5.2.5.1. Gravitational separation

Ten milliliters of each sample were transferred into a glass test tube, sealed with a plastic cap,

and then stored at room temperature for 24 h. Digital photographs (Lumix DMC-ZS8 Digital

Camera, Panasonic Corporation, Newark, NJ, USA) of the samples were taken after storage to

record their stability to gravitational separation.

5.2.5.2. Microstructure

The microstructure of the samples was characterized by confocal fluorescence microscopy. An

optical microscopy (C1 Digital Eclipse, Nikon Co., Tokyo, Japan) with a 60× objective lens was

used to capture images of the emulsions. Emulsions were gently stirred to form a homogeneous

mixture without introducing air bubbles and then the emulsions were stained with fat soluble

fluorescent dye Nile Red (0.1% (w/w) dissolved in 90% (v/v) ethanol) to visualize the location of

the oil phase. A small aliquot of the stained emulsions (5 L) was then transferred to a glass

microscope slide and covered with a glass cover slip. The cover slip was fixed to the slide using

nail polish to avoid evaporation. A small amount of immersion oil (Type A, Nikon Co., Melville,

NY, USA) was placed on the top of cover slip. All fluorescence confocal images were taken

using an excitation argon laser (543 nm) and emitted light was collected between 555 to 620 nm,

and then characterized using the instrument software (EZ CS1 version 3.8, Niko Co., Melville,

NY, USA).

5.2.5.3. Apparent viscosity

The apparent viscosity of samples was measured using a dynamic shear rheometer (Kinexus

Rotational Rheometer, Malvern Instruments Ltd., Worcestershire, United Kingdom). A cup and

bob geometry consisting of a rotating inner cylinder (diameter 25.0 mm) and a static outer

cylinder (diameter 27.5 mm) was used. The samples were loaded into the rheometer measurement

Page 139: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

130

cell and allowed to equilibrate at 37 °C for 5 min before the beginning of all experiments.

Samples underwent a constant shear treatment (10 s-1

for 10 min) prior to analysis to standardize

the shear rate of each sample. The apparent viscosity () was then obtained from measurements

with a shear rate of 10 s-1

selected to mimic oral conditions (Pal, 2011).

5.2.5.4. Particle size distribution

The samples were diluted to a droplet concentration of approximately 0.005% (w/w) using buffer

solution at the appropriate pH prior to analysis to avoid multiple scatterings effects. The particle

size distribution of emulsions was then measured using a static light scattering instrument

(Mastersizer 2000, Malvern Instruments Ltd., Worcestershire, United Kingdom). Refractive

indices of 1.47 (corn oil) and 1.33 (water) were used for the calculations of the particle size

distribution. Background corrections and system alignment were performed prior to each

measurement when the measurement cell was filled with the appropriate buffer solution. Particle

sizes were reported as particle size distribution profiles (volume fraction (%) vs. particle diameter

(μm)) and surface-weighted mean diameter (d32, nm).

5.2.5.5. Surface electrical charge

The surface electrical charge (-potential, mV) of emulsions was determined using a particle

micro-electrophoresis instrument (Zetasizer NanoSeries, Malvern Instruments Ltd.,

Worcestershire, United Kingdom). The emulsions were diluted to a droplet concentration of

approximately 0.005% (w/w) using buffer solution at the appropriate pH prior to analysis. Diluted

emulsions were injected into the measurement chamber, equilibrated for 120 s and then the -

potential was determined by measuring the direction and velocity that the droplets moved in the

applied electric field. Each -potential measurement was calculated from the average of 20

continuous readings made per sample. To determine the effect of pH on the surface electrical

charge (-potential) of pectin solutions (0.5% w/w), a titration between pH 2.0 to 8.0 with 0.25 M

NaOH was performed with an automatic titration unit (Multi Purpose Titrator MPT-2, Malvern

Page 140: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

131

Instruments Ltd., Worcestershire, United Kingdom). The -potential was recorded at each pH

after 60 s equilibrium.

5.2.6. Data analysis

All digestions and measurements were performed at least three times using freshly prepared

samples. Averages and standard deviations were calculated from these triplet measurements.

5.3. Results and discussion

5.3.1. Characterization of the pectin samples

5.3.1.1. FT-IR analysis of functional groups

The FT-IR spectra of the different pectin samples are shown in Figure 5.1. It was observed the

presence of the monosaccharide units making up pectin such as GalA, xylose, arabinose, and

rhamnose, which exhibit intense signals between 1200 and 950 cm-1

wavenumber values and

constituting the fingerprint region specific for each polysaccharide. However, the most

representative signals of the FT-IR spectra of pectin samples are those related to the carboxyl (–

COOH) and carbomethoxyl (–COOCH3) groups (Manrique & Lajolo, 2002). Common features

of all the spectra were: a peak around 3410 cm-1

due to an O-H stretching vibration; a peak

around 2900 cm-1

due to C-H stretching of –CH2 groups; and two peaks at 1610 and 1410 cm-1

due to symmetrical stretching vibrations of the O=C‒O structure. The signal that appears at 1730

cm-1

can be assigned to the C=O stretching vibration of carbomethoxyl group (and also, if

present, of protonated carboxylic group) and shows clear evidence that HMP and MMP (higher

signal strength) were more methoxylated than LMP. FT-IR spectra of aliphatic carboxylic acids

(anionic form) exhibit a characteristic pair of strong intensity signals at 1610 and 1410 cm-1

corresponding, respectively, to asymmetrical and symmetrical stretching vibrations of the

Page 141: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

132

Figure 5.1. Fourier Transform Infrared (FT-IR) spectra of low methoxylated pectin (LMP), medium

methoxylated pectin (MMP) isolated from banana passion fruit (Passiflora tripartita var. mollisima), and

high methoxylated pectin (HMP). The scale was shifted upwards by 100, and 200% for MMP, and HMP

respectively.

carboxylate group (–COO⊝). Considering that a total ionization of –COOH groups could be

attained in the partially methoxylated pectin (e.g. MMP), the 1730 cm-1

signal would be

generated exclusively by the carbomethoxyl group ( ac ur ov , Cape , asin ov , ellner,

Ebringerová, 2000). The similarity of the fingerprint region of MMP with commercial HMP and

LMP pectins, and the relative intensity of the signal at 1730 cm-1

(intermediate intensity between

LMP and HMP) demonstrated that the polysaccharide obtained by acidic extraction from

Passiflora tripartita var. mollisima corresponds to medium methoxylated pectin (MMP).

5.3.1.2. Electrical characteristics of pectin samples

In this series of experiments, we used chemical analysis and micro-electrophoresis to establish

differences in the electrical characteristics of the three pectins. The degree of methoxylation of

the three pectins was 71, 52, and 30% (mol/mol) for the HMP, MMP and LMP samples,

respectively (Table 5.2).

0

50

100

150

200

250

300

5001500250035004500

Tra

nsm

ita

nce

(%

)

Wavenumber (cm-1)

HMP

MMP

LMP

3410

2900 1730 16101410

Page 142: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

133

Table 5.2. Physicochemical properties of high methoxylated pectin (HMP), medium methoxylated pectin

(MMP) isolated from banana passion fruit (Passiflora tripartita var. mollisima), and low methoxylated

pectin (LMP).

Parameter HMP MMP LMP

Average molecular weight (kDa)1 181 148 130

Methoxylation degree (% mol/mol) 71 52 30

Acetylation degree (% mol/mol) 0.4 6.6 0.1

Surface charge at pH 7 (, mV) -28.2 -34.8 -47.5

1Obtained from the multi-angle laser light scattering detector (reported

according to the signal at 90°).

Measurements of the -potential versus pH profiles of the three different pectin samples are

shown in Figure 5.2. In general, all of the samples had their highest negative charges close to pH

8.0, and became less negatively charged (more positive charge) as the pH was decreased with the

steepest change in charge occurring below pH 4.5. This effect can be attributed to protonation of

the carboxyl groups (–COO⊝ + H3O⊕ –COOH + H2O) on the pectin molecules when the pH

is reduced below their pKa values (typically around pH 3.5). As expected, the magnitude of the

negative charge increased with decreasing methoxylation (HMP, = -28.2 mV; MMP, = -34.8

mV; and LMP, = -47.5 mV), since then there were more non-esterified carboxyl groups present

that could be ionized (Table 5.2). It is also important to consider that only MMP had a

considerably degree of acetylation (6.6 % mol/mol) compared to HMP and LMP. Because of the

neutral nature of the acetyl group (–COCH3) this parameter is not expected to contribute to the

overall charge of pectin. Acetyl groups, however, play an important role in the structural

conformation of pectin since this residue is related to controlling the formation of branched

structures (Mohnen, 2008).

5.3.1.3. Molecular weight of pectin samples

The molecular weight distribution of the three pectin samples was determined by HPSEC

(Figure 5.3) and the average molecular weight was calculated. The average molecular weight

values reported in the Table 5.2 were obtained from the multi-angle laser light scattering detector

Page 143: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

134

Figure 5.2. Influence of the pH on the electrical charge (-potential) of high methoxylated pectin (HMP),

medium methoxylated pectin (MMP) isolated from banana passion fruit (Passiflora tripartita var.

mollisima), and low methoxylated pectin (LMP).

signal at 90° (MALLS). The values obtained from the dextran analytical standards were fairly

similar than those obtained from MALLS. All three pectins had mono-modal distributions, but

the width of the distribution was broader for MMP than for LMP and HMP. The average

molecular weights of the three pectins were fairly similar, with all being in the range of 130 to

181 kDa. In general, HMP had the higher average molecular weight (181 kDa), followed by

MMP (148 kDa) and LMP (130 kDa).

5.3.2. Influence of pectin type on gastrointestinal fate of emulsified lipids

In this section, the influence of pectin type on the potential gastrointestinal fate of corn oil-in-

water emulsions was determined using an in vitro digestion model that simulated oral, gastric,

and small intestinal phases. Changes in particle size, electrical charge, microstructure,

appearance, and rheology of the emulsions were measured after their exposure to each stage of

the gastrointestinal tract (Figures 5.4 to 5.8).

-50

-40

-30

-20

-10

0

2 3 4 5 6 7 8

-P

ote

nti

al

(mV

)

pH

HMP

MMP

LMP

Page 144: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

135

Figure 5.3. High performance size exclusion chromatography (HPSEC) profiles of high methoxylated

pectin (HMP), medium methoxylated pectin (MMP) isolated from banana passion fruit (Passiflora

tripartita var. mollisima), and low methoxylated pectin (LMP). The signal corresponds to the differential

refractive index (DRI) detector. Molecular weight scale on top x-axis is based on dextran standards (25,

150, and 410 kDa).

5.3.2.1. Particle size, microstructure, and appearance of emulsions

Initially, all of the emulsions had similar particle size distributions and mean particle diameters

(Figures 5.4 and 5.8a), which suggested that the droplets were stable to coalescence or Ostwald

ripening. However, the confocal microscopy images indicated that all the emulsions containing

pectin were highly flocculated (Figure 5.5), and photographs of the emulsions showed that they

were highly susceptible to creaming (Figure 5.6). The initial oil droplets were coated with a non-

ionic surfactant (Tween 80), and therefore it seems likely that the origin of aggregation was

depletion flocculation rather than bridging flocculation (Blijdenstein, Winden, Vliet, Linden, &

van Aken, 2004). Indeed, calculations of the strength of the osmotic attraction between the

droplets in the presence of pectin support this hypothesis (Section 5.3.3). When the emulsions

were diluted for particle size analysis by laser light scattering the flocs would have been disrupted

0.0

0.5

1.0

1.5

2.0

2.5

0 5 10 15 20 25

DR

I (x

10

5)

Time (min)

HMP

MMP

LMP

25150410

Page 145: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

136

Figure 5.4. Influence of high methoxylated pectin (HMP), medium methoxylated pectin (MMP) isolated

from banana passion fruit (Passiflora tripartita var. mollisima), and low methoxylated pectin (LMP) on

the particle size distribution of emulsions under simulated gastrointestinal conditions consisting of initial

(a), oral (b), gastric (c), and intestinal phases (d). Control corresponds to the emulsions without addition of

pectin. The scale was shifted upwards by 25, 50, and 75% for HMP, MMP, and LMP respectively.

because the amount of pectin present would have fallen below the critical flocculation

concentration (McClements, 2000). After exposure to oral conditions, all of the emulsions

(including the ones containing no pectin) were highly flocculated (Figure 5.5) and exhibited

some creaming (Figure 5.6), but the individual droplets remained relatively small after dilution

0

20

40

60

80

100

10 100 1000 10000

Vo

lum

e F

ract

ion

(%

)

Particle Diameter (nm)

LMP

MMP

HMP

Control

0

20

40

60

80

100

10 100 1000 10000

Vo

lum

e F

ract

ion

(%

)

Particle Diameter (nm)

LMP

MMP

HMP

Control

a. b.

0

20

40

60

80

100

10 100 1000 10000

Vo

lum

e F

ract

ion

(%

)

Particle Diameter (nm)

LMP

MMP

HMP

Control

0

20

40

60

80

100

10 100 1000 10000

Vo

lum

e F

ract

ion

(%

)

Particle Diameter (nm)

LMP

MMP

HMP

Control

c. d.

Page 146: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

137

Figure 5.5. Influence of high methoxylated pectin (HMP), medium methoxylated pectin (MMP) isolated

from banana passion fruit (Passiflora tripartita var. mollisima), and low methoxylated pectin (LMP) on

the microstructure of emulsions observed by confocal fluorescence microscopy under simulated

gastrointestinal conditions consisting of initial (a), oral (b), gastric (c), and intestinal (d) phases. Control

corresponds to the emulsions without addition of pectin.

(Figures 5.4 and 5.8a). This result suggests that the mucin molecules present within the SSF

promoted droplet flocculation through depletion and/or bridging flocculation (Vingerhoeds,

Blijdenstein, Zoet, & van Aken, 2005). Again, the most likely mechanism is depletion

flocculation due to the fact that the fat droplets had a low negative charge, and the flocs easily

dissociated upon dilution for particle size measurements (Jenkins & Snowden, 1996; Klinkesorn,

Sophanodora, Chinachoti, & McClements, 2004; McClements, 2000). After exposure to gastric

conditions, the particle size distribution became broader with many larger particles being present

(Figures 5.4 and 5.8a), and there was evidence of flocculation (Figure 5.5) and creaming

separation (Figure 5.6). The structure and shape of the flocs observed in the gastric phase

LMP MMP HMP

a. Initial

b. Oral

c. Gastric

d. Intestinal

Control

Page 147: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

138

Figure 5.6. Influence of high methoxylated pectin (HMP), medium methoxylated pectin (MMP) isolated

from banana passion fruit (Passiflora tripartita var. mollisima), and low methoxylated pectin (LMP) on

the creaming stability of emulsions under simulated gastrointestinal conditions consisting of initial (a),

oral (b), gastric (c), and intestinal (d) phases. Control (C) corresponds to the emulsions without addition of

pectin.

was quite different from those present in the oral phase (Figure 5.5). In the gastric phase, there

appeared to be a greater number of smaller flocs than in the oral phase. This effect may have

occurred because of the dilution of the emulsions or because of changes in environmental

conditions that changed the nature of the colloidal interactions, such as pH and ionic strength

(Hur, Lim, Decker, & McClements, 2011; Singh, Ye, & Horne, 2009). After exposure to small

intestine conditions, there was evidence of a broad range of different sized particles in both the

particle size distributions (Figure 5.4d) and confocal microscopy images (Figure 5.5). Lipid

digestion may result in numerous different kinds of colloidal particles being present in the

intestinal fluids, including undigested fat droplets, mixed micelles assembled from FFAs, DAGs,

a. Initial b. Oral

c. Gastric d. Intestinal

C HMP MMP LMP C HMP MMP LMP

C HMP MMP LMP C HMP MMP LMP

Page 148: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

139

bile salts, phospholipids, and insoluble calcium soaps formed from long chain FFAs and calcium

(Golding & Wooster, 2010; McClements & Li, 2010). However, it is not possible to discern the

exact nature of these different particles from the light scattering or confocal images.

5.3.2.2. Rheological properties of emulsions

The presence of dietary fibers in the emulsions is likely to alter the rheological properties of the

gastrointestinal fluids, which may impact the rate and extent of lipid digestion by altering mixing

and mass transport processes (Langhout, Schutte, Van Leeuwen, Wiebenga, & Tamminga, 1999).

The apparent viscosity of the gastrointestinal fluids was therefore measured after the emulsions

were exposed to each stage of the simulated GIT (Figure 5.7). Initially, all of the emulsions

containing pectin had a higher viscosity than the control emulsions due to the ability of pectin

molecules to increase the effective volume fraction of the dispersed phase (Mohnen, 2008;

Ridley, O'Neill, & Mohnen, 2001).

Figure 5.7. Influence of high methoxylated pectin (HMP), medium methoxylated pectin (MMP) isolated

from banana passion fruit (Passiflora tripartita var. mollisima), and low methoxylated pectin (LMP) on

the apparent viscosity () of emulsions under simulated gastrointestinal conditions consisting of initial,

oral, gastric, and intestinal phases. Control corresponds to the emulsions without addition of pectin.

0.0

0.2

0.4

0.6

0.8

1.0

Initial Oral Gastric Intestinal

(P

a s

)

Phase

Control

HMP

MMP

LMP

Page 149: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

140

Table 5.3. Molecular characteristics of the high methoxylated pectin (HMP), medium methoxylated pectin

(MMP) isolated from banana passion fruit (Passiflora tripartita var. mollisima), and low methoxylated

pectin (LMP) molecules used in the theoretical calculations of the depletion interactions.

Parameter HMP MMP LMP

Average molecular weight (kDa)1 181 148 130

n2 1006 822 722

rg (nm)3 6.8 4.3 3.1

RV4 12 10 10

1Obtained from the multi-angle laser light scattering detector. (Reported according to the signal at 90°). 2Average number of monomers per molecule (n = MW/MW0). MW is the average molecular weight of the pectin

molecules, and MW0 is the molecular weight of a galacturonic acid monomer unit ( 180 g mol-1). 3Effective radius of the pectin molecules in solution (gyration radius) obtained from the multi-angle laser light scattering

detector (reported according to the signal at 90°). 4Volume ratio (dimensionless). It was assumed that pectin molecules were random coil in conformation.

The extent of the increase in viscosity decreased in the following order HMP > MMP > LMP.

This effect may have been due to differences in the average molecular weight of the different

pectins: HMP > MMP > LMP (Table 5.2), which led to corresponding differences in the radius

of gyration (Table 5.3). Extended polymers with higher molecular weights tend to have higher

effective volume fractions in aqueous solutions, and therefore cause larger increases in viscosity

(McClements, 2000). As the emulsions passed through the successive stages of the simulated

gastrointestinal system there was a progressive decrease in the apparent viscosities of the

emulsions, which can be attributed to the dilution of the systems leading to a lower effective

disperse phase volume fraction. However, in each phase the viscosities still decreased in the same

order for the different pectins: HMP > MMP > LMP. The relatively high viscosities of the

emulsions containing pectin may also have been due to some flocculation of the emulsion

droplets promoted by the biopolymer (e.g. depletion or bridging flocculation) or due to formation

of hydrogel particles (e.g., calcium pectinate) (Jenkins & Snowden, 1996; McClements, 2000).

Interestingly, the viscosities of all the samples were relatively low once they reached the small

intestine phase, which can be attributed to the fact that the emulsions had undergone appreciable

dilution. Consequently, one might not expect a large influence of viscosity on the lipid digestion

process in the small intestine (Golding & Wooster, 2010).

Page 150: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

141

5.3.2.3. Electrical characteristics of emulsions

Initially, the electrical charge on the control emulsions (containing no pectin) was around -7 mV

(Figure 5.8b), which can be attributed to the presence of some anionic impurities in the corn oil

phase (such as DAGs, MAGs, FFAs or phospholipids) as well as surfactant ingredients, to the

ionization of the hydroxyl groups of Tween 80 (Tween 80 is a nonionic surfactant with the

presence of some weakly ionizable hydroxyl groups), specific ion adsorption from bulk solution

onto lipid droplets surfaces, or due to the preferential adsorption of hydroxyl ions (OH⊝), rather

than hydronium ions (H3O⊕) from water by the lipid droplet surfaces (Nikiforidis &

Kiosseoglou, 2011). The addition of pectin to the emulsions led to an increase in the measured

negative charge, with the magnitude of the effect increasing with decreasing degree of

methoxylation. This effect can be attributed to the contribution of the anionic pectin molecules to

the measured signal used to calculate the -potential.

Figure 5.8. Influence of high methoxylated pectin (HMP), medium methoxylated pectin (MMP) isolated

from banana passion fruit (Passiflora tripartita var. mollisima), and low methoxylated pectin (LMP) on

the volume-surface mean diameter (d32, a) and the electrical charge (-potential, b) of emulsions under

simulated gastrointestinal conditions consisting of initial, oral, gastric, and intestinal phases. Control

corresponds to the emulsions without addition of pectin.

-40

-30

-20

-10

0

Initial Oral Gastric Intestinal

-P

ote

nti

al

(mV

)

Phase

Control

HMP

MMP

LMP0.0

0.3

0.6

0.9

1.2

Initial Oral Gastric Intestinal

d3

2(

m)

Phase

Control

HMP

MMP

LMP

a. b.

Page 151: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

142

The electrical charge became slightly more negative in all of the emulsions after exposure to the

oral conditions, which can be attributed to the presence of anionic mucin molecules in the SSF

(Espinal-Ruiz, Parada-Alfonso, Restrepo-Sanchez, Narvaez-Cuenca, & McClements, 2014;

Vingerhoeds, Blijdenstein, Zoet, & van Aken, 2005).

The magnitude of the negative charge on the particles decreased appreciably when exposed to

simulated stomach phase, which may be due to the relatively low pH and high ionic strength of

the gastric fluids (Singh, Ye, & Horne, 2009). The acidic pH of the gastric fluids reduced the

negative charge on the pectin molecules, as well as on the non-ionic surfactant coated oil

droplets. Finally, the negative charged increased appreciably after exposure to the simulated

small intestine conditions, which can be attributed to the anionic nature of the molecules that

assemble the colloidal particles in this phase, i.e., FFAs, bile salts, and phospholipids (Hur, Lim,

Decker, & McClements, 2011; McClements & Li, 2010). The electrical properties of the

emulsions were affected by pectin samples according to their methoxylation degree (related to

electrical charge). For all gastrointestinal phases, the measured negative charge of the emulsions

increased with decreasing the degree of pectin methoxylation, which can be attributed to the

higher negative charge density of the pectin molecules (Figure 5.8b). In addition, the pectin

molecules are not digested within the upper gastrointestinal tract, and therefore, they should

remain in the gastrointestinal fluids of each phase, thereby contributing to the measured electrical

properties (Ridley, O'Neill, & Mohnen, 2001).

5.3.2.4. Digestion of emulsified lipids

In this section, the influence of the different kinds of pectin on the rate and extent of lipid

digestion was determined. In general, there was a rapid increase in the amount of FFAs released

within the first few minutes, followed by a more gradual increase at later times (Figure 5.9a). For

the control sample, the amount of FFAs formed eventually reached around 100% (w/w)

indicating that all of the TAGs were hydrolyzed by the lipase. For the emulsions containing

pectin samples, the lipid digestion profile depended on the nature of the pectin molecules in the

system. The final extent of lipid digestion decreased in the following order:

Page 152: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

143

Figure 5.9. Influence of high methoxylated pectin (HMP), medium methoxylated pectin (MMP) isolated

from banana passion fruit (Passiflora tripartita var. mollisima), and low methoxylated pectin (LMP) on

free fatty acids (FFA) released after the digestion process. Kinetic profile of intestinal release of FFA (a),

and FFA released after 2 hours of intestinal digestion (b). Control corresponds to the emulsions without

addition of pectin.

100, 92, 70 and 47% (w/w) for the control, LMP, MMP, and HMP samples, respectively (Figure

5.9b). These results suggested that the extent of lipid digestion decreased as the degree of

methoxylation and the molecular weight of the pectin molecules increased. It should be stressed

that the results obtained using simple simulated GIT models should be treated with caution, since

they cannot represent the compositional, structural, and dynamic complexity of the processes

occurring within the human GIT. Nevertheless, they may provide some useful insights into the

potential physicochemical mechanisms occurring within the GIT.

In principle, there are numerous ways that pectin methoxylation can alter the lipid digestion

process. An increase in methoxylation leads to an increase in the number of non-polar

(hydrophobic) groups on the molecules, and a decrease in the number of negative groups

(Dongowski, Lorenz, & Proll, 2002). An increase in the number of non-polar groups may lead to

increased binding of bile salts through hydrophobic attraction (Dongowski, 1995; Espinal-Ruiz,

Parada-Alfonso, Restrepo-Sánchez, Narváez-Cuenca, & McClements, 2014; Wilde & Chu,

0

20

40

60

80

100

0 20 40 60 80 100 120

FF

A R

elea

sed

(%

w/w

)

Digestion Time (min)

Control

LMP

MMP

HMP

0

20

40

60

80

100

Control LMP MMP HMP

Fin

al

Dig

esti

on

(F

FA

% w

/w)

Sample

a. b.

Page 153: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

144

2011). The binding of bile salts upon addition of pectin may inhibit lipid digestion because they

can no longer interact with the fat droplet surfaces (thereby inhibiting lipase adsorption) or

because they can no longer solubilize FFAs generated at the fat droplet surfaces and thereby,

inhibiting lipase activity (Espinal-Ruiz, Parada-Alfonso, Restrepo-Sánchez, Narváez-Cuenca, &

McClements, 2014). An increase in methoxylation would therefore be expected to decrease the

extent of lipid digestion through this effect (Reis, Holmberg, Watzke, Leser, & Miller, 2009). A

decrease in the number of anionic carboxyl groups on the pectin chains (e.g., increased

methoxylation) may lead to decreased binding with cations, such as calcium ions (Willats, Knox,

& Mikkelsen, 2006). Calcium ions play an number of important roles in the lipid digestion

process: (i) a minimum level is required for proper lipase functioning; (ii) they precipitate long

chain FFAs, thereby removing them from the fat droplet surfaces and avoiding the adsorption of

lipase; and (iii) they form insoluble soaps with long chain FFAs thereby decreasing their

absorption (Wilde & Chu, 2011). An increase in methoxylation degree (decreased negative

charge) might therefore be expected to alter the extent of lipid digestion. Differences in the

ability of pectin molecules to promote lipid droplet flocculation may have altered the digestion

extent (Espinal-Ruiz, Parada-Alfonso, Restrepo-Sanchez, Narvaez-Cuenca, & McClements,

2014). Flocculated fat droplets may be digested more slowly than non-flocculated ones, because

the surface area of lipids exposed to the lipase in the aqueous phase is reduced (Reis, Holmberg,

Watzke, Leser, & Miller, 2009). In this study, all of the pectins used promoted flocculation in the

mouth, stomach, and small intestine and therefore had potential to inhibit digestion through this

mechanism. Nevertheless, there may have been differences in the nature of the flocs formed, e.g.,

the packing of the fat droplets within the flocs (Figure 5.5). There are a number of

physicochemical phenomena that might account for the observed decrease in lipid digestion with

increasing methoxylation of the pectin molecules, such as binding of bile salts to the non-polar

groups. However, further studies would be required to characterize the importance of this

mechanism.

Besides the contribution of the methoxylation degree, the molecular weight (as stated above in

Section 5.3.2.2) is also an important parameter that contributes to the overall inhibition of lipid

digestion. The viscosity of the gastrointestinal phases increased with increasing molecular weight

Page 154: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

145

of pectin samples (HMP > MMP > LMP, Figure 5.7). It is well known that the higher the

molecular weight of pectin, the greater its capacity to form complex structures (e.g., gels and

hydrated networks) which are able to trap water and other components such as lipids in their

inner structures (Willats, Knox, & Mikkelsen, 2006). An increase in the viscosity of the gel

causes a restriction on the diffusive processes of lipids and lipases, inhibiting their capacity to

interact to each other and consequently, reducing the lipolytic reaction extent. Furthermore, as the

lipids are trapped inside pectin gels, lipases will not be able to access the lipid surfaces and thus,

the lipolytic reaction extent will be also reduced (Espinal-Ruiz, Parada-Alfonso, Restrepo-

Sanchez, Narvaez-Cuenca, & McClements, 2014). Although both the methoxylation degree and

molecular weight are important parameters determining the physicochemical properties of pectin

molecules, other structural parameters such as monosaccharide composition and degree of

branching may also influence their functionality, and therefore their influence on the

gastrointestinal fate of lipids. Further studies are therefore needed to clarify the importance of

specific molecular characteristics on pectin functionality.

5.3.3. Calculation of the depletion attraction between the lipid droplets

In this section, we provide a theoretical rationalization for the influence of pectin type (HMP,

MMP, and LMP) on the depletion flocculation of the emulsions in terms of the characteristics of

the different pectin molecules (Klinkesorn, Sophanodora, Chinachoti, & McClements, 2004;

McClements, 2000). The presence of non-adsorbed pectin molecules in the aqueous phase (bulk

solution) of an emulsion is known to increase the osmotic attraction between the lipid droplets

through a depletion mechanism (McClements, 2000). The magnitude of this attractive interaction

can be calculated using the following equations (Klinkesorn, Sophanodora, Chinachoti, &

McClements, 2004):

[ (

)

(

)

(

)

(

)] (5.2)

(

) (5.3)

Page 155: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

146

Here, wdepletion(h) is the inter-droplet pair potential due to depletion interactions at a surface-to-

surface droplet separation of h, r is the lipid droplet radius, POsm is the osmotic pressure arising

from the exclusion of non-adsorbed pectin molecules from a narrow region ( rg) surrounding the

lipid droplets, and rg is the effective radius of the pectin molecules in solution (gyration radius).

In addition, c, MW, and m are the concentration, molecular weight, and molecular density of the

pectin molecules in the aqueous solution, respectively. NA is the Avogadro´s number, k is the

Boltzmann´s constant, and T is the absolute temperature. The parameter, RV, is referred to as the

volume ratio, which is equal to the effective volume of a pectin molecule in solution divided by

the actual volume of the constituent atoms making up the molecule (McClements, 2000). If a

pectin molecule adopts a compact spherical conformation, like a globular protein, then RV 1.

However, pectin molecules entrain large quantities of water as they rotate in solution, then RV ⪢

1. The effective volume of pectin in solution should be considerably greater than the volume

occupied by the atoms that make up the pectin chain because it sweeps out a large volume of

solvent as it rapidly rotates due to Brownian motion (Jenkins & Snowden, 1996; McClements,

2000). In this model, we have assumed that pectin molecules behave like random coil in solution,

so that:

(5.4)

Here, n is the number of monomer units per molecule (n = MW/MW0), MW is the molecular

weight of the whole pectin molecules, MW0 is the molecular weight of a GalA monomer unit (

180 g mol-1

), l is the length of the monomer unit ( 0.47 nm), and m is the density of the pectin

chain ( 2000 kg m-3

). The molecular characteristics of the pectin molecules used in our

calculations are shown in Table 5.3. It should be noted that w(h)/kT = 0 for h 2rg and that the

strongest interaction between the fat droplets occurs when they come into contact (h = 0). So that,

Equation 5.2 is applicable for the boundary condition h < 2rg and when the separation between

the lipid droplets is small compared to their size (h ⪡ r).

Page 156: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

147

Equations 5.2, 5.3, and 5.4 were used to calculate the influence of pectin type (HMP, MMP, and

LMP) on the depletion attraction between lipid droplets (r = 100 nm), assuming that pectin

molecules behave as random coil in aqueous solution. The variation of the droplet attraction

potential (w(h)/kT) with the droplet separation (h) for different pectin types, but the same overall

aqueous concentrations (1.8% (w/w) equivalent to the initial phase) is shown in Figure 5.10a.

Both the magnitude (depletion attraction w(h)/kT) and the range (lipid droplet separation h) of the

attractive depletion attraction between lipid droplets increased with increasing molecular weight

and methoxylation degree: HMP > MMP > LMP > Control. An estimate of the overall strength of

the depletion attraction in a particular system can be obtained by calculating the magnitude of

wdepletion (h = 0) when the droplets are in contact:

*

+ (5.5)

The dependence of wdepletion(h = 0)/kT on pectin type was calculated (Figure 5.10b). The strength

of the depletion attraction increases progressively with the simultaneous increase of the

molecular weight and methoxylation degree (HMP > MMP > LMP > Control). In the absence of

pectin (control), the lipid droplets are prevented from flocculating because the repulsive droplet-

droplet interactions (e.g., steric, electrostatic, and hydration repulsion) dominate the attractive

interactions (e.g., van der Waals and electrostatic) (McClements, 2000). Addition of the pectin

molecules to the emulsion increases the depletion attraction between the lipid droplets, until

eventually the overall attractive interactions overcome the repulsive interactions and the droplets

flocculate. As the molecular weight and the methoxylation degree of the pectin molecules

increased simultaneously, a smaller amount of pectin needs to be added to the emulsion in order

to generate the additional attraction (depletion force) required to promote droplet flocculation

(Jenkins & Snowden, 1996). These calculations suggest that each of the pectins tested can

promote depletion flocculation in the emulsions used in this work. In particular, Equation 5.2

suggests that the strength of the depletion interaction is highly dependent on the molecular weight

of the pectin molecules as well as the surface electrical charge (methoxylation degree), which can

be represented by the gyration radius (rg).

Page 157: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

148

Figure 5.10. Inter-droplet pair potential attraction due to depletion interactions of lipid droplets containing

high methoxylated pectin (HMP), medium methoxylated pectin (MMP) isolated from banana passion fruit

(Passiflora tripartita var. mollisima), and low methoxylated pectin (LMP), related to the thermal energy

(kT) of the system. Inter-droplet pair potential (w(h)/kT) due to depletion interactions at a surface-to-

surface droplet separation of h (a), and inter-droplet pair potential (w(h=0)/kT) when the droplets are in

contact (b). The model corresponds to the depletion interactions of lipid droplets in the initial phase (1.8%

w/w pectins) at 37 °C, prior to subjection to the static in vitro digestion model. Control corresponds to the

emulsions without addition of pectin.

In addition, it should be stressed that the electrical properties of pectin molecules may also

indirectly influence the depletion interaction by altering the effective size (gyration radius) of the

colloidal particles and the depletion zone. For example, increasing the number of negative

charges on a pectin molecule by either decreasing the methoxylation degree or increasing the pH,

can increase its effective size. In the one hand, increasing the number of negative groups usually

causes the pectin molecules to become more extended because of electrostatic repulsion between

negatively charged groups (–COO⊝). On the other hand, decreasing the number of negative

charges on a pectin molecule by either increasing the methoxylation degree or reducing the pH,

can decrease its effective size. Thus, the strength of the depletion interaction may depend on the

electrical properties of the pectin molecules as well as the environmental conditions such as pH,

temperature, and ionic strength (Furusawa, Ueda, & Nashima, 1999). Additional calculations

were conducted to evaluate the relative contribution of the molecular weight and the

-2.5

-2.0

-1.5

-1.0

-0.5

0.0

0.01 0.1 1 10

w(h

)/k

T

h (nm)

Control

LMP

MMP

HMP-2.5

-2.0

-1.5

-1.0

-0.5

0.0

Control LMP MMP HMP

w(h

=0

)/k

TSample

a. b.h

Page 158: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

149

Figure 5.11. Schematic representation of the inhibition of lipid digestion by pectin. High methoxylated

pectin (HMP) and medium methoxylated pectin (MMP) isolated from banana passion fruit (Passiflora

tripartita var. mollisima) form a closed structure around the lipid droplets (depletion flocculation) which

restricts the access of lipase to their surfaces and therefore, preventing the lipid digestion (a), whereas low

methoxylated pectin (LMP) forms an open structure due to its electrostatic repulsion with negatively

charged lipid droplets, allowing the access of lipase to their surfaces and promoting lipid digestion (b).

methoxylation degree to the overall magnitude of the depletion attraction. These calculations

were performed by fixing the rg value (rg = 4.7 nm) and changing the molecular weight

(according to Table 5.3), and by fixing the molecular weight (MW = 153 kDa) and changing the

rg value (according to Table 5.3 as well). These calculations allowed us to establish that the

molecular weight of pectin molecules had a significant impact on the magnitude of the depletion

attraction (w(h=0)/kT of -0.3, -1.1, and -1.7 for LMP, MMP, and HMP, respectively) while the

degree of methoxylation (represented by rg) had a smaller contribution to the overall magnitude

of the depletion attraction (w(h=0)/kT of -0.04, -0.07, and -0.09 for LMP, MMP, and HMP,

respectively). Finally, we can suggest that both HMP and MMP form a closed structure around

the lipid droplets which restricts the access of lipase to their surfaces and therefore, preventing

Lipase Lipid dropletHMP

MMPLMP

a. b.

Page 159: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

150

lipid digestion (Figure 5.11a), whereas LMP forms an open structure due to its repulsion with

negatively charged lipid droplets, allowing the access of lipase to their surfaces and promoting

lipid digestion (Figure 5.11b).

5.4. Conclusions

The objective of this work was to study the impact of different types of pectin on the

physicochemical characteristics and microstructure of emulsified lipids during passage through a

simulated gastrointestinal model. Three pectins with different molecular characteristics were

studied: LMP and HMP from citrus fruit and MMP from banana passion fruit. These pectins

differ in their molecular weights and degrees of methoxylation, which led to differences in their

molecular dimensions (radius of gyration) and electrical characteristics (-potential). All three

pectins promoted flocculation of the fat droplets in the emulsions, which was attributed to a

depletion flocculation mechanism, associated with exclusion of the biopolymers from the fat

droplet surfaces. The pectin molecules decreased the extent of lipid digestion with increasing

degree of methoxylation and molecular weight: HMP > MMP > LMP. These effects may have

been due to the impact of the pectin molecules on the rheological properties of the

gastrointestinal fluids, binding of key digestive components (such as calcium, free fatty acids,

and bile), alteration in the droplet aggregation state, or entrapment of the lipid droplets by pectin

microgels. Further studies are clearly required to establish the relative contribution of the

methoxylation degree and the molecular weight to the overall inhibitory effect, as well as to

identify the precise molecular origin of this inhibition. This information may be useful for the

design of emulsion-based functional foods that give healthier lipid profiles and thereby promote

health and wellness.

Acknowledgments

We are grateful to Departamento Administrativo de Ciencias, Tecnología e Innovación

(COLCIENCIAS) and Vicerrectoría Académica of Universidad Nacional de Colombia for

Page 160: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

151

providing a fellowship to Mauricio Espinal-Ruiz supporting this work. We also thank the United

States Department of Agriculture (USDA), NRI Grants (2011-03539, 2013-03795, 2011-67021,

and 2014-67021); and Red Nacional para la Bioprospección de Frutas Tropicales

COLCIENCIAS-RITFRUBIO (Contrato 0459-2013) for supporting this research. We are grateful

to student Mayra Alejandra Quintero from Departamento de Química, Universidad Nacional de

Colombia, for supporting both the extraction and characterization of Passiflora tripartita var.

mollissima pectin (MMP) sample.

References

Beysseriat, M., Decker, E. A., & McClements, D. J. (2006). Preliminary study of the influence of dietary

fiber on the properties of oil-in-water emulsions passing through an in vitro human digestion

model. Food Hydrocolloids, 20(6), 800-809.

Blijdenstein, T. B. J., Winden, A. J. M. v., Vliet, T. v., Linden, E. v. d., & van Aken, G. A. (2004). Serum

separation and structure of depletion- and bridging-flocculated emulsions: a comparison. Colloids

and Surfaces A: Physicochemical and Engineering Aspects, 245(1–3), 41-48.

Bray, G. A., & Popkin, B. M. (1998). Dietary fat intake does affect obesity! The American Journal of

Clinical Nutrition, 68(6), 1157-1173.

Dongowski, G. (1995). Influence of pectin structure on the interaction with bile acids under in vitro

conditions. Zeitschrift für Lebensmittel-Untersuchung und Forschung, 201(4), 390-398.

Dongowski, G., Lorenz, A., & Proll, J. (2002). The Degree of Methylation Influences the Degradation of

Pectin in the Intestinal Tract of Rats and In Vitro. The Journal of Nutrition, 132(7), 1935-1944.

Edashige, Y., Murakami, N., & Tsujita, T. (2008). Inhibitory Effect of Pectin from the Segment

Membrane of Citrus Fruits on Lipase Activity. Journal of Nutritional Science and Vitaminology,

54(5), 409-415.

Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sanchez, L.-P., Narvaez-Cuenca, C.-E., & McClements,

D. J. (2014). Impact of dietary fibers [methyl cellulose, chitosan, and pectin] on digestion of lipids

under simulated gastrointestinal conditions. Food & Function, 5, 3083-3095.

Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L.-P., Narváez-Cuenca, C.-E., & McClements,

D. J. (2014). Interaction of a Dietary Fiber (Pectin) with Gastrointestinal Components (Bile Salts,

Calcium, and Lipase): A Calorimetry, Electrophoresis, and Turbidity Study. Journal of

Agricultural and Food Chemistry, 62(52), 12620-12630.

Funami, T., Nakauma, M., Ishihara, S., Tanaka, R., Inoue, T., & Phillips, G. O. (2011). Structural

modifications of sugar beet pectin and the relationship of structure to functionality. Food

Hydrocolloids, 25(2), 221-229.

Furusawa, K., Ueda, M., & Nashima, T. (1999). Bridging and depletion flocculation of synthetic latices

induced by polyelectrolytes. Colloids and Surfaces, A: Physicochemical and Engineering Aspects,

153(1–3), 575-581.

Gil, M., Restrepo, A., Millán, L., Alzate, L., & Rojano, B. (2014). Microencapsulation of Banana Passion

Fruit (Passiflora tripartita Var. mollissima): A New Alternative as a Natural Additive as

Antioxidant. Food and Nutrition Sciences, 5, 671-682.

Page 161: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

152

Golding, M., & Wooster, T. J. (2010). The influence of emulsion structure and stability on lipid digestion.

Current Opinion in Colloid & Interface Science, 15(1–2), 90-101.

Grabitske, H. A., & Slavin, J. L. (2009). Gastrointestinal Effects of Low-Digestible Carbohydrates.

Critical Reviews in Food Science and Nutrition, 49(4), 327-360.

Hur, S. J., Lim, B. O., Decker, E. A., & McClements, D. J. (2011). In vitro human digestion models for

food applications. Food Chemistry, 125(1), 1-12.

Inman, M. (2011). How Bacteria Turn Fiber into Food. PLoS Biology, 9(12), e1001227.

Jenkins, P., & Snowden, M. (1996). Depletion flocculation in colloidal dispersions. Advances in Colloid

and Interface Science, 68(0), 57-96.

ac ur ov , ., Cape , ., asin ov , ., ellner, N., & Ebringerová, A. (2000). FT-IR study of plant

cell wall model compounds: pectic polysaccharides and hemicelluloses. Carbohydrate Polymers,

43(2), 195-203.

Klinkesorn, U., Sophanodora, P., Chinachoti, P., & McClements, D. J. (2004). Stability and rheology of

corn oil-in-water emulsions containing maltodextrin. Food Research International, 37(9), 851-

859.

Kumar, V., Sinha, A. K., Makkar, H. P. S., de Boeck, G., & Becker, K. (2011). Dietary Roles of Non-

Starch Polysachharides in Human Nutrition: A Review. Critical Reviews in Food Science and

Nutrition, 52(10), 899-935.

Langhout, D. J., Schutte, J. B., Van Leeuwen, P., Wiebenga, J., & Tamminga, S. (1999). Effect of dietary

high-and low-methylated citrus pectin on the activity of the ileal microflora and morphology of

the small intestinal wall of broiler chicks. British Poultry Science, 40(3), 340-347.

Lattimer, J. M., & Haub, M. D. (2010). Effects of Dietary Fiber and Its Components on Metabolic Health.

Nutrients, 2(12), 1266-1289.

Li, Y., Kim, J., Park, Y., & McClements, D. J. (2012). Modulation of lipid digestibility using structured

emulsion-based delivery systems: Comparison of in vivo and in vitro measurements. Food &

Function, 3(5), 528-536.

Manrique, G. D., & Lajolo, F. M. (2002). FT-IR spectroscopy as a tool for measuring degree of methyl

esterification in pectins isolated from ripening papaya fruit. Postharvest Biology and Technology,

25(1), 99-107.

McClements, D. J. (2000). Comments on viscosity enhancement and depletion flocculation by

polysaccharides. Food Hydrocolloids, 14(2), 173-177.

McClements, D. J., Decker, E. A., & Park, Y. (2009). Controlling Lipid Bioavailability through

Physicochemical and Structural Approaches. Critical Reviews in Food Science and Nutrition,

49(1), 48-67.

McClements, D. J., & Li, Y. (2010). Review of in vitro digestion models for rapid screening of emulsion-

based systems. Food & Function, 1(1), 32-59.

Mesbahi, G., Jamalian, J., & Farahnaky, A. (2005). A comparative study on functional properties of beet

and citrus pectins in food systems. Food Hydrocolloids, 19(4), 731-738.

Minekus, M., Alminger, M., Alvito, P., Ballance, S., Bohn, T., Bourlieu, C., Carriere, F., Boutrou, R.,

Corredig, M., Dupont, D., Dufour, C., Egger, L., Golding, M., Karakaya, S., Kirkhus, B., Le

Feunteun, S., Lesmes, U., Macierzanka, A., Mackie, A., Marze, S., McClements, D. J., Menard,

O., Recio, I., Santos, C. N., Singh, R. P., Vegarud, G. E., Wickham, M. S. J., Weitschies, W., &

Brodkorb, A. (2014). A standardised static in vitro digestion method suitable for food - an

international consensus. Food & Function, 5(6), 1113-1124.

Mohnen, D. (2008). Pectin structure and biosynthesis. Current Opinion in Plant Biology, 11(3), 266-277.

Nikiforidis, C. V., & Kiosseoglou, V. (2011). Competitive displacement of oil body surface proteins by

Tween 80 – Effect on physical stability. Food Hydrocolloids, 25(5), 1063-1068.

Pal, R. (2011). Rheology of simple and multiple emulsions. Current Opinion in Colloid & Interface

Science, 16(1), 41-60.

Page 162: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 5

153

Reis, P., Holmberg, K., Watzke, H., Leser, M. E., & Miller, R. (2009). Lipases at interfaces: A review.

Advances in Colloid and Interface Science, 147–148(0), 237-250.

Ridley, B. L., O'Neill, M. A., & Mohnen, D. (2001). Pectins: structure, biosynthesis, and

oligogalacturonide-related signaling. Phytochemistry, 57(6), 929-967.

Rufino, M. d. S. M., Alves, R. E., de Brito, E. S., Pérez-Jiménez, J., Saura-Calixto, F., & Mancini-Filho, J.

(2010). Bioactive compounds and antioxidant capacities of 18 non-traditional tropical fruits from

Brazil. Food Chemistry, 121(4), 996-1002.

Schieber, A., Stintzing, F. C., & Carle, R. (2001). By-products of plant food processing as a source of

functional compounds — recent developments. Trends in Food Science & Technology, 12(11),

401-413.

Singh, H., & Sarkar, A. (2011). Behaviour of protein-stabilised emulsions under various physiological

conditions. Advances in Colloid and Interface Science, 165(1), 47-57.

Singh, H., Ye, A., & Horne, D. (2009). Structuring food emulsions in the gastrointestinal tract to modify

lipid digestion. Progress in Lipid Research, 48(2), 92-100.

Torcello-Gomez, A., Maldonado-Valderrama, J., Martin-Rodriguez, A., & McClements, D. J. (2011).

Physicochemical properties and digestibility of emulsified lipids in simulated intestinal fluids:

influence of interfacial characteristics. Soft Matter, 7(13), 6167-6177.

Tsujita, T., Sumiyoshi, M., Han, L. K., Fujiwara, T., Tsujita, J., & Okuda, H. (2003). Inhibition of lipase

activities by citrus pectin. Journal of Nutritional Science and Vitaminology, 49(5), 340-345.

van den Hoogen, B. M., van Weeren, P. R., Lopes-Cardozo, M., van Golde, L. M. G., Barneveld, A., &

van de Lest, C. H. A. (1998). A Microtiter Plate Assay for the Determination of Uronic Acids.

Analytical Biochemistry, 257(2), 107-111.

Vingerhoeds, M. H., Blijdenstein, T. B. J., Zoet, F. D., & van Aken, G. A. (2005). Emulsion flocculation

induced by saliva and mucin. Food Hydrocolloids, 19(5), 915-922.

Voragen, A. G. J., Schols, H. A., & Pilnik, W. (1986). Determination of the degree of methylation and

acetylation of pectins by h.p.l.c. Food Hydrocolloids, 1(1), 65-70.

Voragen, F. J., Timmers, J. J., Linssen, J. H., Schols, H., & Pilnik, W. (1983). Methods of analysis for

cell-wall polysaccharides of fruit and vegetables. Zeitschrift für Lebensmittel-Untersuchung und

Forschung, 177(4), 251-256.

Wilde, P. J., & Chu, B. S. (2011). Interfacial & colloidal aspects of lipid digestion. Advances in Colloid

and Interface Science, 165(1), 14-22.

Willats, W. G. T., Knox, J. P., & Mikkelsen, J. D. (2006). Pectin: new insights into an old polymer are

starting to gel. Trends in Food Science & Technology, 17(3), 97-104.

Wootton-Beard, P. C., & Ryan, L. (2011). Improving public health?: The role of antioxidant-rich fruit and

vegetable beverages. Food Research International, 44(10), 3135-3148.

Yao, M., Xiao, H., & McClements, D. J. (2014). Delivery of Lipophilic Bioactives: Assembly,

Disassembly, and Reassembly of Lipid Nanoparticles. Annual Review of Food Science and

Technology, 5(1), 53-81.

Yonekura, L., & Nagao, A. (2009). Soluble Fibers Inhibit Carotenoid Micellization in Vitro and Uptake

by Caco-2 Cells. Bioscience Biotechnology and Biochemistry, 73(1), 196-199.

Yoo, S.-H., & Jane, J.-l. (2002). Molecular weights and gyration radii of amylopectins determined by

high-performance size-exclusion chromatography equipped with multi-angle laser-light scattering

and refractive index detectors. Carbohydrate Polymers, 49(3), 307-314.

Page 163: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

Interaction of a dietary fiber (pectin) with gastrointestinal components (bile

salts, calcium, and lipase): A calorimetry, electrophoresis, and turbidity study

Published as:

Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L. P., Narváez-Cuenca, C. E.,

& McClements, D. J. Journal of Agricultural and Food Chemistry. 62 (2014): 12620 – 12630.

Page 164: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

155

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. The interaction of pectin with

gastrointestinal components may be related to the formation of microgel structures which are able

to promote the flocculation of lipid droplets by depletion interactions and thereby, reducing their

digestibility. It was found that pectin reduced the extent of lipid digestion, which was attributed

to its interactions with 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. This research provides valuable insights into

the physicochemical and molecular mechanisms of the interaction of pectin with gastrointestinal

components that may impact the rate and extent of lipid digestion.

Keywords: Pectin, gastrointestinal tract, lipid digestion, isothermal titration calorimetry,

turbidity, flocculation.

Page 165: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

156

6.1. Introduction

There has been an appreciable increase in the total amount of calories consumed by humans

during the past few decades, which is believed to be an important contributing factor to increases

in obesity, diabetes, and cardiovascular diseases (Bray & Popkin, 1998). Fat has the highest

calorie density of the major food components (fat, protein, and carbohydrates) and so there has

been a major focus on the identification of effective strategies to reduce the fat content of foods,

while maintaining their desirable quality attributes. Several studies have suggested that certain

types of dietary fiber can inhibit lipid digestion and absorption in the small intestine (Beysseriat,

Decker, & McClements, 2006; Gunness & Gidley, 2010; Howarth, Saltzman, & Roberts, 2001;

Hur, Lim, Decker, & McClements, 2011; Li & McClements, 2014; McClements & Li, 2010a,

2010b; Mun, Decker, Park, Weiss, & McClements, 2006). Increased consumption of dietary fiber

may therefore be one approach for reducing some of the adverse affects associated with eating

high-fat food products.

Dietary fibers can be classified as either water-soluble or water-insoluble (Anderson, Baird,

Davis Jr, Ferreri, Knudtson, Koraym, et al., 2009). Water-insoluble fibers, such as lignins,

celluloses, and some hemicelluloses, play an important role in regulating intestinal peristalsis

(Kritchevsky, 1988). Water-soluble fibers, such as pectin, carageenan, xanthan gum, and alginate,

influence the gastrointestinal fate of ingested foods due to their ability to bind water, thicken or

gel intestinal fluids, and interact with specific food and gastrointestinal components (Theuwissen

& Mensink, 2008). Pectin is a water-soluble dietary fiber that is used widely in the

pharmaceutical, biotechnology, and food industries as a functional ingredient (Thakur, Singh,

Handa, & Rao, 1997). Pectin is a polysaccharide that has linear anionic regions formed by D-

galacturonic acid (GalA) monomers linked by -(1,4) glycosidic bonds, and branched regions

primarily formed by various types of neutral monosaccharides (mainly rhamnose, xylose,

mannose, and arabinose) linked together. The GalA units have carboxyl groups, which may be

present as free carboxyl groups or methyl esterified groups depending on the orign, isolation, and

processing of pectin (Caffall & Mohnen, 2009; Mohnen, 2008; Ridley, O'Neill, & Mohnen, 2001;

Page 166: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

157

Thakur, Singh, Handa, & Rao, 1997; Yapo, 2011). Pectin has been successfully used for many

years in the food and beverage industries as a thickening and gelling agent, as well as a colloidal

stabilizer (Maxwell, Belshaw, Waldron, & Morris, 2012; Thakur, Singh, Handa, & Rao, 1997).

The gelling characteristics of pectin have been utilized to form hydrogel delivery systems for a

range of pharmaceutical and food bioactive compounds (Maxwell, Belshaw, Waldron, & Morris,

2012; Munarin, Tanzi, & Petrini, 2012).

Previous studies have shown that pectin reduces the rate and extent of lipid digestion by

presumably interacting with various food and intestinal components (Beysseriat, Decker, &

McClements, 2006; Gunness & Gidley, 2010; Li & McClements, 2014). A number of different

physicochemical and physiological mechanisms have been proposed to account for this effect.

Pectin can interact with bile salts and phospholipids in the small intestine, which may alter lipid

digestion by reducing the amount of surface-active components available to stabilize the lipid

droplets (triacylglycerols), or to solubilize and transport lipid digestion products (free fatty acids

and monoacylglycerols) from the droplet surfaces to the epithelium cells (Hur, Lim, Decker, &

McClements, 2011; McClements & Li, 2010a). The binding of bile salts to pectin in the small

intestine has also been proposed as one of the major mechanisms responsible for the ability of

pectin to reduce cholesterol levels (Pfeffer, Doner, Hoagland, & McDonald, 1981). Pectin

molecules may interfere with the re-absorption of bile salts in the small intestine, thereby

reducing the amount of cholesterol absorbed and transported to the blood (Pfeffer, Doner,

Hoagland, & McDonald, 1981). Furthermore, the conformation and aggregation state of pectin in

aqueous solutions may be altered due to its interactions with certain gastrointestinal components,

which leads to changes in solution rheology that might impact lipid digestion, e.g., by altering

gastric emptying times or the mass transport of digestive enzymes (Beysseriat, Decker, &

McClements, 2006).

Pectin may also directly interact with other components in the gastrointestinal tract such as

CaCl2, NaCl, and digestive enzymes (Espinal-Ruiz, Parada-Alfonso, Restrepo-Sánchez, &

Narváez-Cuenca, 2014; McClements, 2000b; McClements & Li, 2010a; Simo, Mao, Tokle,

Decker, & McClements, 2012). Finally, pectin is known to promote depletion flocculation of

Page 167: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

158

lipid droplets, which may reduce lipid digestion by protecting them from attack by lipolytic

enzymes (Jenkins & Snowden, 1996b; McClements, 2000a).

Pectin ingredients obtained from different natural sources using different isolation and processing

methods may vary widely in their molecular and physicochemical characteristics, such as

backbone length, electrical charge, hydrophobicity, surface-activity, conformation, self-

association, viscosity, and binding capacity (Yapo, 2011). In addition, the molecular and

functional characteristics of pectins may be altered appreciably in response to changes in solution

and environmental conditions such as pH, composition, ionic strength, and temperature

(Maxwell, Belshaw, Waldron, & Morris, 2012; Mohnen, 2008; Munarin, Tanzi, & Petrini, 2012;

Ridley, O'Neill, & Mohnen, 2001). For this reason, the study of the nature of the interactions that

pectin may have with the components of the gastrointestinal tract is important in understanding

the effect of pectin on digestive processes. An improved understanding of the origin and nature of

these interactions would lead to the design of functional foods with improved nutritional,

physicochemical, and sensory properties.

In this study, isothermal titration calorimetry (ITC) was used to study the interactions between

pectin and gastrointestinal components (bile salts, pancreatic lipase, CaCl2, and NaCl). ITC

measures the heat absorbed or evolved when one solution is titrated into another solution (Doyle,

1997; Freire, Mayorga, & Straume, 1990; Leavitt & Freire, 2001). Previous studies have shown

that ITC is an extremely valuable tool for studying chemical interactions of polysaccharides with

other molecules because it provides valuable data concerning binding enthalpies, critical

aggregation concentrations, and binding stoichiometries (Chang, McLandsborough, &

McClements, 2011; Wangsakan, Chinachoti, & McClements, 2004). In addition, electrophoresis,

turbidity, and microstructural observations were used to provide additional information about the

nature of the interactions between pectin and gastrointestinal components. The results of this

study may aid the rational design of functional foods designed to improve human health and

wellness by controlling lipid digestion within the gastrointestinal tract.

Page 168: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

159

6.2. Materials and methods

6.2.1. Chemicals

Corn oil was purchased from a commercial food supplier (Mazola, ACH Food Companies Inc.,

Memphis, TN) and stored at 4 °C until use. The manufacturer reported that this oil contained

approximately 14, 29, and 57% (w/w) of saturated, monounsaturated, and polyunsaturated fatty

acids, respectively. Commercial powdered high methoxylated pectin (Genu Pectin (Citrus),

USP/100) was kindly donated by CP Kelco (Lille Skensved, Denmark) and was used without

further purification. The manufacturer reported that the powdered pectin ingredient contained

6.9% moisture, 89.0% galacturonic acid, and 8.6% methoxyl groups, with a degree of

methoxylation of approximately 62%. The average molecular weight was reported to be 200 kDa.

Lipase from porcine pancreas (Type II, L3126, triacylglycerol hydrolase E.C. 3.1.1.3), bile

extract (porcine, B8631), mucin from porcine stomach (Type II, M2378, bound sialic acid ≤

1.2%), and pepsin A from porcine gastric mucose (P7000, endopeptidase E.C. 3.4.23.1, activity ≥

250 units/mg solid) were purchased from Sigma-Aldrich Chemical Company (St Louis, MO,

USA). The supplier reported that the lipase activity was 100-400 units mg-1

protein (using olive

oil) and 30-90 units mg-1

protein (using triacetin) for 30 min incubation (one unit of lipase

activity was defined as the amount of enzyme required for the release of 1 eq of fatty acid from

either triacetin (pH 7.4) or olive oil (pH 7.7) in 1 h at 37 °C). The composition of the bile extract

was reported as 49% (w/w) total bile salt (BS), containing 10-15% glycodeoxycholic acid, 3-9%

taurodeoxycholic acid, 0.5-7% deoxycholic acid, 1-5% hydrodeoxycholic acid, and 0.5-2% cholic

acid; 5% (w/w) phosphatidyl choline (PC); Ca2+ ≤ 0.06% (w/w); critical micelle concentration of

bile extract 0.07 0.04 mM; and mole ratio of BS to PC being around 15:1. All other chemicals

were purchased from Sigma-Aldrich Chemical Company (St Louis, MO, USA). Double distilled

water was used to prepare all solutions.

Page 169: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

160

6.2.2. Simulated gastrointestinal studies

6.2.2.1. Solutions and emulsions preparation

Pectin stock solution (4.0% w/w) was prepared by dispersing 4 g of powdered pectin into 96 g of

5 mM phosphate buffer solution (pH 7.0). The resulting solution was stirred at 800 rpm for 12 h

(overnight) at room temperature to ensure complete dispersion and dissolution. Pectin stock

solution was then adjusted to pH 7.0 and equilibrated for 10 min before analysis.

A stock emulsion was prepared by mixing together 20% (w/w) corn oil and 80% (w/w) buffered

emulsifier solution (5 mM phosphate buffer, pH 7.0; containing 2.5% (w/w) Tween 80) for 5 min

using a bio-homogenizer (Speed 2, Model MW140/2009-5, Biospec Products Inc., ESGC,

Switzerland). The resulting coarse emulsion was then passed 5 times through a high-pressure

homogenizer (Microfluidizer M-110L processor, Microfluidics Inc., Newton, MA, USA)

operating at 11,000 psi (75.8 MPa) to reduce the particle size further.

Pectin-emulsion mixtures were then prepared by mixing stock emulsion (containing 20% (w/w)

corn oil) with pectin stock solution (containing 4.0% (w/w) pectin), to obtain systems of varying

composition: 2.0% (w/w) corn oil and 0.2-3.6% (w/w) pectin. The pectin-emulsion mixtures were

then stirred with a high-shear mixer (Fisher Steadfast Stirrer, Model SL-1200, Fisher Scientific,

Pittsburgh, PA) at 1000 rpm and stored overnight at room temperature. Pectin-emulsion mixtures

were then characterized to obtain the initial phase, prior to subjection to the in vitro

gastrointestinal model.

6.2.2.2. Simulated gastrointestinal tract model

Each emulsion sample (initial phase) was passed through a simulated in vitro gastrointestinal

tract that consisted of oral, gastric, and intestinal phases. This method is based on a static model

utilized in previous studies (Hur, Decker, & McClements, 2009; Salvia-Trujillo, Qian, Martin-

Belloso, & McClements, 2013).

Page 170: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

161

6.2.2.2.1. Oral phase

Simulated saliva fluid (SSF, pH 6.8) containing 3% (w/w) mucin was prepared according to the

composition shown in Table 6.1. Each emulsion (initial phase) was mixed with SSF (ratio 1:1

w/w) to obtain a mixture containing 1% (w/w) corn oil and 0.1-1.8% (w/w) pectin. The simulated

oral phase consisted of a conical flask containing emulsion-SSF mixture incubated at 37 °C with

continuous shaking at 100 rpm for 10 min in a temperature controlled air incubator (Excella E24

Incubator Shaker, New Brunswick Scientific, NJ, USA). The mixture resulting from processing

of the initial emulsions in the oral phase (bolus) was used in the gastric phase.

6.2.2.2.2. Gastric phase

Simulated gastric fluid (SGF) was prepared by adding 2.0 g NaCl, 7.0 mL concentrated HCl

(37% w/w), and 3.2 g pepsin A (from porcine gastric mucose, 250 units mg-1

) to a flask, and then

diluting with double distilled water to a volume of 1.0 L, and finally adjusting to pH 1.2 using 1.0

M HCl. Samples from the oral phase (bolus) were mixed with SGF (ratio 1:1 w/w) so that the

final mixture contained 0.5% (w/w) corn oil, and 0.05-0.90% (w/w) pectin.

Table 6.1. Chemical composition of simulated saliva fluid (SSF) used to simulate oral conditions.

Compound Chemical formula Concentration (g L-1

)1

Sodium chloride NaCl 1.594

Ammonium nitrate NH4NO3 0.328

Potassium dihydrogen phosphate KH2PO4 0.636

Potassium chloride KCl 0.202

Potassium citrate K3C6H5O7•H2O 0.308

Uric acid sodium salt C5H3N4O3Na 0.021

Urea H2NCONH2 0.198

Lactic acid sodium salt C3H5O3Na 0.146

Porcine gastric mucin (Type II) ---- 30

1The SSF was prepared in double distilled water and then pH 6.8 was adjusted using 0.1 M NaOH.

Page 171: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

162

This mixture was then adjusted to pH 2.5 using 1.0 M NaOH and incubated at 37 °C with

continuous shaking at 100 rpm for 2 h. The mixture resulting from processing of the emulsions in

the gastric phase (chyme) was used in the intestinal phase.

6.2.2.2.3. Intestinal phase

Samples obtained from the gastric phase (20.0 mL) were incubated for 2 h at 37 °C in a simulated

small intestine fluid (SIF) consisting of 2.5 mL pancreatic lipase (24 mg mL-1

), 3.5 mL bile

extract solution (54 mg mL-1

), and 1.5 mL salt solution containing 0.25 M CaCl2 and 3.0 M

NaCl, to obtain a final composition in the reaction vessel of 0.36% (w/w) corn oil, and 0.05-

0.65% (w/w) pectin. The free fatty acids (FFAs) released were monitored by determining the

amount of 0.10 M NaOH needed to maintain a constant pH 7.0 within the reaction vessel using

an automatic titration unit (pH stat titrator, 835 Titrando, Metrohm USA, Inc., Riverview, FL,

USA). All components were dissolved in 5 mM phosphate buffer solution (pH 7.0) before use.

Lipase addition and initialization of the titration program were carried out only after the addition

of all pre-dissolved ingredients and balancing the pH to 7.0. The volume of 0.10 M NaOH added

to the emulsion for balancing the pH to 7.0 was recorded over time and used to calculate the

FFAs generated by lipolysis. The amount of FFAs released over time were calculated using the

following equation:

(

) (6.1)

Where, CNaOH is the concentration of the sodium hydroxide (0.10 M), MWLipid is the average

molecular weight of corn oil (872 g mol-1

), WLipid is the initial weight of corn oil in the intestinal

phase (0.10 g), and VNaOH is the volume of NaOH (L) titrated into the reaction vessel to

neutralize the FFA released, assuming that all triacylglycerols are hydrolyzed in two molecules of

FFA and one molecule of monoacylglycerol. Titration blanks were performed by inactivating

lipase in boiling water for 15 min prior to initialization of the titration program.

Page 172: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

163

6.2.2.3. Creaming stability measurements of initial emulsions

Ten milliliters of emulsions (initial phase) were transferred into a test tube (internal diameter 15

mm, height 125 mm), tightly sealed with a plastic cap, and then stored at room temperature for 24

h, after which appreciable phase separation into an opaque layer at the top, a turbid layer in the

middle, and a transparent layer at the bottom was observed in some of the systems. We defined

the serum layer to be the sum of the turbid and transparent layers. The total height of the

emulsion (HE) and the height of the serum layer (HS) were measured using a laser vertical

profiling system (Turbiscan Classic MA2000, Formulaction, Wynnewood, PA). The extent of

creaming was then characterized by the creaming index (CI), defined as CI = 100 × (HS / HE).

The creaming index provided indirect information about droplet aggregation, since an increase in

particle size (e.g., due to flocculation) leads to faster creaming (provided the droplet

concentration is not too high).

6.2.3. Interaction of pectin with gastrointestinal components

6.2.3.1. Solutions preparation

Stock solutions of 0.9% (w/w) pectin, 2.7% (w/w) bile salts, and 1.1% (w/w) lipase were

individually prepared by dispersing them in 5 mM phosphate buffer solution (pH 7.0) at room

temperature for 12 h (overnight), to ensure their complete dispersion and dissolution. Salt

solutions consisting of 55 mM CaCl2 or 545 mM NaCl were prepared freshly before each

experiment. A mixture containing the above components (at the same initial concentrations of the

individual solutions), was also prepared. The initial concentration of this mixture was referred as

100% relative concentration of each component.

To establish the interaction of pectin with specific gastrointestinal components, individual

solutions of CaCl2, NaCl, lipase, bile salts, and the mixture of all components, were titrated into

either buffer solution (blank) or pectin at 37 °C. Twenty-nine 100 L aliquots of individual

solutions containing 2.7% (w/w) bile salts, 1.1% (w/w) lipase, 55 mM CaCl2, 545 mM NaCl, or

Page 173: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

164

the mixture of all components (100% relative concentration), were added sequentially into a test

tube containing 14.5 mL of either buffer (5 mM phosphate buffer, pH 7.0) or 0.9% (w/w) pectin

solutions. Turbidity, -potential, and microstructure (observed by optical microscopy) of these

solutions were then characterized throughout the titration process.

6.2.3.2. Solutions characterization

6.2.3.2.1. Optical turbidity

Optical turbidity of the solutions was determined by measuring their absorbance at 600 nm

(=A600 nm) using a UV-Visible spectrophotometer (Ultrospec 3000 pro Pharmacia Biotech,

Biochrom Ltd., Cambridge, UK) at 37 °C. The samples were contained within 1 cm path length

optical cells, and buffer solution was used as a control. The change in turbidity was defined as

= Pectin Buffer. Triplicate measurements of turbidity were carried out on each sample.

6.2.3.2.2. Electrical surface charge

Electrical surface charge (-potential) of the solutions was determined using a particle

microelectrophoresis instrument (Zetasizer NanoSeries, Malvern Instruments Ltd.,

Worcestershire, UK). Solutions were injected into the measurement chamber, equilibrated for

120 s and then the -potential was determined by measuring the direction and velocity that the

particles moved in the applied electric field. Each individual -potential measurement was

calculated from the average of 20 continuous readings made per sample. The -potential was

recorded at each pH after 60 s equilibrium.

6.2.3.2.3. Microstructure

Microstructure of the solutions was characterized by optical microscopy. An optical microscope

(C1 Digital Eclipse, Niko, Tokyo, Japan) with a 60× objective lens was used to capture images of

Page 174: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

165

the solutions. Solutions were gently stirred to form a homogeneous mixture without introducing

any air bubbles. A small aliquot of the solutions (5 L) was then transferred to a glass

microscope slide and covered with a glass cover slip. The cover slip was fixed to the slide using

nail polish to avoid evaporation. A small amount of immersion oil (Type A, Nikon, Melville, NY,

USA) was placed on the top of the cover slip. All optical images were taken using a digital

camera, and then characterized using the instrument software (EZ CS1 version 3.8, Niko,

Melville, NY, USA).

6.2.3.3. Isothermal titration calorimetry measurements

An isothermal titration calorimeter (ITC) instrument (Microcalorimeter VP-ITC, MicroCal Inc.,

Northampton, MA, USA) was used to measure the enthalpy (H) resulting from titration of

individual solutions of CaCl2, NaCl, lipase, bile salts, and a mixture of all components into either

buffer solution (blank) or pectin. Twenty-nine 10 L aliquots of individual solutions of 2.7%

(w/w) bile salts, 1.1% (w/w) lipase, 55 mM calcium chloride, 545 mM sodium chloride, or a

mixture of all components (100% relative concentration) were injected sequentially into a 1450

L titration cell initially containing either buffer (5 mM phosphate buffer, pH 7.0) or 0.9% (w/w)

pectin solutions (pectin and gastrointestinal component concentrations mimic approximately the

concentrations in the intestinal phase of the simulated gastrointestinal model). Each injection

lasted 24 s and there was an interval of 240 s between successive injections. The temperature of

the solution in the titration cell was 37 °C and the solution was stirred at a speed of 315 rpm

throughout the experiments. Triplicate measurements of enthalpy were carried out on each

sample and they were reproducible to better than 5%.

6.2.4. Data analysis

All measurements were performed at least three times using freshly prepared samples. Averages

and standard deviations were calculated from these triplet measurements.

Page 175: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

166

6.3. Results and discussion

6.3.1. Properties of initial emulsion – pectin mixtures

Initially, the properties of the initial emulsions used to prepare the emulsion-pectin mixtures were

characterized (5 mM phosphate buffer, pH 7.0). The initial emulsions had a relatively small mean

particle diameter immediately after preparation (d32 = 200 10 nm), which did not change during

the course of the experiments (Espinal-Ruiz, Parada, Restrepo-Sanchez, Narvaez-Cuenca, &

McClements, 2014). This can be attributed to the relatively strong steric repulsion between the

non-ionic surfactant-coated lipid droplets resulting from the polyoxyethylene headgroups on the

Tween 80 molecules (Almgren, 2000). The lipid droplets initially had a small negative charge (

= -5.8 0.2 mV), even though they were stabilized by a non-ionic surfactant, which can be

attributed to anionic impurities in the oil or surfactant ingredients (such as free fatty acids and

phospholipids) or to preferential adsorption of hydroxyl ions from water (OH⊝) (Singh, Ye, &

Horne, 2009; Thiam, Farese Jr, & Walther, 2013). The droplets in the initial emulsions

containing no pectin were stable to creaming throughout the experimental time frame (Figure

6.1), which is due to the fact that the droplets were small and did not associate with each other

(Weiss, Takhistov, & McClements, 2006).

Emulsions containing relatively low levels of added pectin (≤ 0.4% w/w) were still observed to be

stable to gravitational separation, i.e., they maintained a uniform white appearance (Figure 6.1).

The stability of the emulsions in presence of low levels of pectin can be evidenced according to

the slight increase of the particle size (d32 = 210 nm) of the lipid droplets in the initial phase

(containing 0.2% (w/w) pectin), compared to emulsions without pectin (d32 = 200 nm). However,

emulsions containing higher levels of pectin clearly separated into a creamed layer and a serum

layer, indicating that the droplets had moved upwards due to gravity. The origin of this effect is

depletion flocculation induced by the presence of non-adsorbed pectin molecules within the

aqueous phase (Dickinson, Semenova, Antipova, & Pelan, 1998; Jenkins & Snowden, 1996b).

Page 176: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

167

Figure 6.1. Influence of the concentration of pectin on creaming stability of 2% (w/w) corn oil-in-water

emulsions stabilized with 0.2% (w/w) Tween 80 before the digestion process (initial phase). Pectin

concentrations are referred to the initial phase.

The depletion flocculation of the lipid droplets in presence of high levels of pectin can be

evidenced according to the increase of the particle size (d32 = 350 nm) in the initial phase

(containing 2.4% (w/w) pectin), as compared to emulsions without pectin (d32 = 200 nm). Pectin

molecules should not be attracted to the surfaces of the non-ionic surfactant-coated lipid droplets

because of electrostatic and steric repulsion effects (the carboxyl groups of pectin are completely

dissociated at pH 7.0 since 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 (Jenkins & Snowden, 1996a). 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) since when they come into close contact, the volume of the

thermodynamically unfavorable exclusion zone is reduced (McClements, 2000a). 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 repulsion), and the emulsion remains stable to

droplet aggregation. However, once a critical pectin concentration is exceeded, the attractive

0

20

40

60

80

100

0 1 2 3 4

Cre

am

ing

In

dex

(%

)

Pectin (% w/w)

Pectin Concentration (% w/w)

0.0 0.4 1.2 2.4 3.6

Page 177: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

168

forces exceed the repulsive forces and droplet flocculation occurs (Jenkins & Snowden, 1996b).

It should be noted that the main thermodynamic driving force for depletion flocculation is

entropy (Méndez-Alcaraz & Klein, 2000; Reiffers-Magnani, Cuq, & Watzke, 2000). The reason

that the thickness of the creamed layer increases (CI decreases) as the pectin concentration

increases above the critical flocculation concentration (Figure 6.1) is that the depletion attraction

is higher and so the droplets are held more strongly into a particle gel network (Dickinson, 1995).

In summary, these measurements showed that the degree of droplet flocculation in the initial

emulsion-pectin mixtures depended on the amount of pectin present in the system.

6.3.2. Influence of pectin on lipid digestion

In a recent study, we examined the influence of pectin addition on the gastrointestinal fate of

emulisified lipids (Espinal-Ruiz, Parada, Restrepo-Sanchez, Narvaez-Cuenca, & McClements,

2014). 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. 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 was

reached after 25 min of digestion (Figure 6.2a). The final amount of FFAs produced at the end of

the two hour digestion period decreased with increasing pectin concentration (Figure 6.2b),

which is in good agreement with our earlier study (Espinal-Ruiz, Parada, Restrepo-Sanchez,

Narvaez-Cuenca, & McClements, 2014). The ability of pectin to decrease the rate and extent of

lipid digestion (principally the extent rather than rate) in corn oil-in-water emulsions could be due

to several physicochemical and physiological mechanisms.

Page 178: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

169

Figure 6.2. Influence of the concentration of pectin on free fatty acids (FFA) released after the digestion

process. Kinetic profile of intestinal release of FFA (a), and FFA released after 2 hours of intestinal

digestion (b). Pectin concentrations are referred to the intestinal phase.

Pectin may have interacted with calcium ions and formed a highly viscous solution or gel

network that impeded the diffusion of GIT components (such as lipase or bile salts), thereby

retarding the lipid digestion process (Braccini & Pérez, 2001; Leroux, Langendorff, Schick,

Vaishnav, & Mazoyer, 2003; Willats, Knox, & Mikkelsen, 2006). Pectin-calcium interactions

may also have inhibited lipid digestion due to the important role that calcium ions play in

removing FFAs from lipid droplet surfaces (Devraj, Williams, Warren, Mullertz, Porter, &

Pouton, 2013). Calcium ions normally form insoluble soaps with long chain FFAs that help to

remove them from the lipid droplet surfaces and thereby allow the lipase to keep working (Ye,

Cui, Zhu, & Singh, 2013). If the calcium ions are tightly bound to pectin molecules, then the

FFAs may accumulate at the lipid droplet surfaces, thereby inhibiting further digestion. Pectin

may also have bound to bile salts or phospholipids, and so prevented them from from adsorbing

to lipid droplet surfaces or from solubilizing lipid digestion products (Anderson, et al., 2009;

Leroux, Langendorff, Schick, Vaishnav, & Mazoyer, 2003; Maxwell, Belshaw, Waldron, &

Morris, 2012). The solubilization of long chain FFAs in mixed micelles is another important

means of removing them from the lipid droplet surfaces and thereby allowing lipase to keep

60

70

80

90

100

0.00 0.07 0.15 0.55 0.65

Fin

al

FF

A R

elea

sed

(%

w/w

)Pectin (% w/w)

0

20

40

60

80

100

0 50 100 150

FF

A R

elea

sed

(%

w/w

)

Time (min)

0.00%

0.07%

0.15%

0.55%

0.65%

a. b.

Page 179: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

170

functioning (Almgren, 2000). The adsorption of bile salts and phospholipids onto lipid droplet

surfaces often facilitates the subsequent adsorption of lipase molecules. Pectin may also be able

to alter lipid digestion by interacting directly with lipase molecules (de Roos & Walstra, 1996;

Espinal-Ruiz, Parada-Alfonso, Restrepo-Sánchez, & Narváez-Cuenca, 2014). Finally, the

presence of pectin in the system may alter the aggregation state of the lipid droplets through

either depletion or bridging flocculation (Biggs, Habgood, Jameson, & Yan, 2000).

The digestion of highly flocculated lipid droplets is often less than that of non-flocculated ones

because the lipase molecules have to diffuse through the flocs before they can reach the lipid

droplets in their interiors (Li, Hu, & McClements, 2011; Simo, Mao, Tokle, Decker, &

McClements, 2012). Clearly, the potential influence of pectin on the lipid digestion process

within the GIT is complex since pectin is able to interact with many constitiuents within the GIT

fluids. The purpose of this study was therefore to provide further insights into the potential origin

of these effects by studying the interactions between pectin and gastrointestinal components.

6.3.3. Interactions of pectin with mixed gastrointestinal components

In this series of experiments, we used ITC, electrophoresis, turbidity, and microstructure

measurements to examine the interactions of pectin with a solution containing a mixture of all the

major GIT components: sodium, calcium, bile salts, and lipase. The microstructure of the

emulsion-pectin mixtures changed appreciably after exposure to the mixed gastrointestinal

components (Figure 6.3). In the absence of pectin (control), the individual lipid droplets were too

small to be observed by optical microscopy, but confocal microscopy images have shown that

they are evenly distributed throughout the system (Espinal-Ruiz, Parada, Restrepo-Sanchez,

Narvaez-Cuenca, & McClements, 2014). After digestion, there were clearly some large

aggregates in the control samples, which may have been lipid digestion products or remnants

from the various GIT components (such as calcium, sodium, bile salts, and lipase). In the

presence of pectin, the lipid droplets were trapped within large clusters prior to digestion (lipid

droplets size of d32 = 350 nm for an emulsion containing 2.4% (w/w) pectin, compared to lipid

droplets size of d32 = 200 nm for an emulsion without pectin), which can be attributed to

Page 180: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

171

Figure 6.3. Influence of 2.4% (w/w) pectin on the microstructure (observed by optical microscopy) of 2%

(w/w) corn oil-in-water emulsions stabilized with 0.2% (w/w) Tween 80 before (initial phase) and after

(intestinal phase) of the digestion process. The scale bar corresponds to 20 m.

depletion flocculation induced by the non-adsorbed pectin molecules (Figure 6.3) (McClements,

2000a). After digestion, there were a relatively high number of large aggregates (undigested lipid

droplets) remaining in the samples containing pectin, which could have been due to interactions

of the pectin molecules with various components in the simulated GIT fluids. Turbidity (), -

potential, and enthalpy change (H) were measured when increasing amounts of mixed GIT

fluids (0 to 100% of mixture) were titrated into a pectin (0.9% w/w) solution at pH 7.0 (Figure

6.4). The 100% GIT fluids contained the levels of the NaCl, CaCl2, bile salts, and lipase in the

final simulated small intestinal fluids. There was a progressive increase in the turbidity when the

amount of mixed GIT fluids titrated into the pectin solution increased (Figure 6.4a). The -

potential of the mixed system moved from highly negative (-19.0 mV) to less negative (-2.0 mV)

as increasing amounts of GIT fluids were added (Figure 6.4a), which may have occurred due to

electrostatic screening effects induced by the presence of salts or due to binding of cations in the

GIT fluids to anionic pectin molecules (Israelachvili, 2011).

Pectin

Before

Digestion

After

Digestion

Control

Page 181: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

172

Figure 6.4. Influence of the relative concentration (0 to 100%) of simulated gastrointestinal fluids on

optical turbidity (=A600 nm) and -potential (a), and interaction enthalpy (H) (b) 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 () was defined as = Pectin - Buffer.

There was also an appreciable difference in the enthalpy profiles of samples with either buffer or

pectin solutions (Figure 6.4b). In the absence of pectin, the enthalpy change was relatively low,

and may be attributed to heat of dilution effects associated with the various molecules and/or

particles moving further apart when the GIT fluids were injected into the reaction cell (Chang,

McLandsborough, & McClements, 2011; Doyle, 1997; McClements, 2000b).

In the presence of pectin, there was a large exothermic enthalpy change when the first few

aliquots of the mixed GIT fluids were injected into the reaction cell containing the pectin

solution. The enthalpy change became progressively less exothermic when 0 to 20% of the GIT

fluids were added, until it eventually reached a relatively constant endothermic value at higher

GIT levels. These results suggest that the pectin molecules interacted with some of the

components in the GIT fluids, possibly through electrostatic interactions, and formed aggregates

that were large enough to scatter light strongly.

-20

-15

-10

-5

0

0.0

0.5

1.0

1.5

2.0

0 20 40 60 80 100

-P

oten

tial (m

V)

(A6

00

nm

)

Relative concentration (%)

-200

-100

0

100

200

0 20 40 60 80 100

H

(k

cal/

inje

ctio

n)

Relative concentration (%)

Pectin

Buffer

a. b.

Page 182: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

173

6.3.4. Interactions of pectin with specific gastrointestinal components

In this series of experiments, ITC, electrophoresis, turbidity, and microstructure measurements

were used to characterize the interactions of pectin with specific GIT components (NaCl, CaCl2,

bile salts, or lipase). Measurements were made as increasing amounts of GIT components were

individually added to a pectin solution (0.9% w/w).

Figure 6.5. Influence of the concentration of NaCl (a), CaCl2 (b), bile salts (c), and lipase (d) on optical

turbidity (=A600 nm) and -potential of solutions containing an initial concentration of 0.9% (w/w) pectin.

Optical turbidity () was defined as = Pectin - Buffer.

-20

-15

-10

-5

0

0.0

0.5

1.0

1.5

2.0

0 2 4 6 8 10

-P

oten

tial (m

V)

(A6

00

nm

)

CaCl2 (mM)

-20

-15

-10

-5

0

0.0

0.5

1.0

1.5

2.0

0 20 40 60 80 100

-P

oten

tial (m

V)

(A6

00

nm

)

NaCl (mM)

-20

-15

-10

-5

0

0.0

0.5

1.0

1.5

2.0

0.0 0.1 0.2 0.3 0.4 0.5

-P

oten

tial (m

V)

(A6

00

nm

)

Bile salts (% w/w)

a.

-20

-15

-10

-5

0

0.0

0.5

1.0

1.5

2.0

0.00 0.05 0.10 0.15 0.20

-P

oten

tial (m

V)

(A6

00

nm

)

Lipase (% w/w)

b.

c. d.

Page 183: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

174

In addition, the turbidities and -potentials (Figure 6.5) reached at the end of these titrations

(after the injection of twenty-nine aliquouts) are reported, so that the different samples can easily

be compared (Figure 6.6).

6.3.4.1. NaCl

Addition of increasing amounts of NaCl to the pectin solutions caused an appreciable decrease in

the magnitude of the -potential of the pectin solutions (Figure 6.5a), which can be attributed to

electrostatic screening effects, i.e., accumulation of sodium ions around the negative groups on

the pectin molecules (Israelachvili, 2011). There was no evidence of an increase in aggregate

formation within the microstructure images (Figure 6.8) and there was little change in the

general appearance (Figure 6.7) and optical turbidity (Figure 6.5a) of these samples upon NaCl

addition. We did observe a slight increase in the turbidity at the highest NaCl levels used, which

suggested that the NaCl may have caused a slight amount of pectin aggregation, presumably

through screening of the electrostatic interactions (Furusawa, Ueda, & Nashima, 1999). The ITC

measurements indicated that there was not a strong interaction (H = -102 kcal/injection)

between the pectin and the NaCl (Figure 6.9a), since the enthalpy versus salt concentration

profiles were fairly similar. Overall, these results suggest that pectin did not have a major

interaction with the NaCl in the GIT fluids.

6.3.4.2. CaCl2

Addition of CaCl2 to the pectin solutions had a more pronounced influence on their

physicochemical properties than NaCl addition. Addition of increasing amounts of CaCl2 caused

the -potential of the pectin solutions to go from around -19.0 mV to -6.0 mV (Figure 6.5b),

which can be attributed to electrostatic screening and ion binding effects (Israelachvili, 2011).

Multivalent counter-ions are more effective at screening electrostatic interactions than

monovalent ones, as they have a greater tendency to bind strongly to oppositely charged groups

(Winzor, Carrington, Deszczynski, & Harding, 2004).

Page 184: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

175

Figure 6.6. 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 (=A600 nm) (a)

and -potential (b) of solutions containing an initial concentration of 0.9% (w/w) pectin (P).

The electrical change in the -potential with calcium addition was fairly similar to that observed

with addition of the mixed GIT solution (Figure 6.6b), which suggests that calcium ions play an

important role in determing the overall electrical interactions. The ITC measurements also

indicated that there was a strong interaction (H = -241 kcal/injection) between the calcium ions

and pectin (Figure 6.9b). When calcium ions were injected into buffer solution there was an

exothermic enthalpy change that decreased in magnitude with increasing CaCl2 concentration,

which can be attributed to heat of dilution effects (McClements, 2000b). However, when calcium

ions were injected into pectin solution there was initially a strong exothermic reaction (from 0 to

4 mM), then a relatively strong endothermic reaction (from 4 to 6 mM), followed by an enthalpy

change close to zero after 6 mM CaCl2 (Figure 6.9b). The shape of the ITC curve suggests that

there may have been several events occurring sequentially, such as calcium binding, pectin

conformational changes, and/or pectin aggregation. Nevertheless, it is not possible to determine

the molecular origin of these events from ITC measurements alone, since they just provide the

overall enthalpy change.

0.0

0.5

1.0

1.5

2.0

P P+NaCl P+L P+BS P+CaCl2 P+M

(A6

00

nm

)

-20

-15

-10

-5

0P P+BS P+L P+NaCl P+CaCl2 P+M

-P

ote

nti

al

(mV

)

a. b.

Page 185: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

176

Figure 6.7. Influence of the concentration of NaCl, CaCl2, bile salts, lipase, and a mixture of all

components (% relative concentrations) on turbidity of solutions containing an initial concentration of

0.9% (w/w) pectin.

The addition of calcium ions to the pectin solutions caused a large increase in their optical

turbidity (Figures 6.5b and 6.6a), which can be attributed to the appearance of large aggregates,

as observed within the optical microscopy images (Figure 6.8). We postulate that these

aggregates were microgels formed by pectin in the presence of calcium (Braccini & Pérez, 2001;

Willats, Knox, & Mikkelsen, 2006). The calcium ions may have cross-linked the anionic pectin

molecules through electrostatic bridging (Thakur, Singh, Handa, & Rao, 1997). As mentioned

earlier, the binding of calcium ions to pectin may be important for a number of reasons.

NaCl (mM)

Lipase (% w/w)

Mixture (%)

Bile salts (% w/w)

CaCl2 (mM)

0 27 49 70 91 0 2.7 4.9 7.0 9.0

0 0.15 0.25 0.35 0.45

0 23 45 65 100

0 0.04 0.09 0.14 0.18

Page 186: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

177

Figure 6.8. Influence of 91 mM NaCl, 9 mM CaCl2, 0.45% (w/w) bile salts, 0.18% (w/w) lipase, 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 m.

Buffer

Bile salts

NaCl

Lipase

Mixture

CaCl2

Pectin

Page 187: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

178

Figure 6.9. Influence of the concentration of NaCl (a), CaCl2 (b), bile salts (c), and lipase (d) on

interaction enthalpy (H) of solutions containing either buffer solution or an initial concentration of 0.9%

(w/w) pectin.

Firstly, pectin-calcium microgels may trap lipid droplets inside, thereby restricting the access of

lipase to the lipid substrates (Simo, Mao, Tokle, Decker, & McClements, 2012). Second, the

formation of these electrostatic complexes may prevent the calcium ions from forming insoluble

soaps with long-chain FFAs at the lipid droplet surfaces, thereby preventing the lipase from

functioning properly (Ye, Cui, Zhu, & Singh, 2013). Although the interaction of pectin with

NaCl had not a significant impact, the presence of NaCl may influence the interaction of pectin

-200

-100

0

100

200

0.0 0.1 0.2 0.3 0.4 0.5

Hea

t (k

cal/

inje

ctio

n)

Bile salts (% w/w)

Pectin

Buffer

-200

-100

0

100

200

0.00 0.05 0.10 0.15 0.20

Hea

t (k

cal/

inje

ctio

n)

Lipase (% w/w)

Pectin

Buffer

-200

-100

0

100

200

0 2 4 6 8 10

Hea

t (k

cal/

inje

ctio

n)

CaCl2 (mM)

Pectin

Buffer

-200

-100

0

100

200

0 20 40 60 80 100

Hea

t (k

cal/

inje

ctio

n)

NaCl (mM)

Pectin

Buffer

a. b.

c. d.

Page 188: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

179

with other gastrointestinal components such as CaCl2. In general, salts are known to screen

electrostatic interactions in aqueous solutions and would therefore be expected to influence the

formation and properties of electrostatic complexes. For instance, it can be postulated that

electrostatic interaction between pectin and CaCl2 can be weakened in the presence of NaCl,

which means that more CaCl2 had to be added to reach charge neutralization. The addition of

NaCl also would reduce the high turbidity reached with CaCl2, since the formation of calcium-

pectin complexes would be partially suppressed by electrostatic screening interactions with NaCl.

There may be considerable variations in the type and concentration of ions surrounding the lipids

droplets, which may impact the electrostatic interactions in the system through electrostatic

screening or binding effects. For example, long chain FFAs and bile salts may precipitate in the

presence of calcium ions, thereby removing them from the lipid droplet surface facilitating their

further digestion, but which may also reduce their subsequent absorption due to calcium soap

formation. Sufficiently high concentrations of monovalent (NaCl) and multivalent ions (CaCl2)

can promote extensive flocculation of emulsions containing electrically charged droplets, which

may restrict the access of lipase to the oil-water interface and slow down digestion. Certain types

of ions are capable of promoting the gelation of biopolymers, which would affect the ability of

digestive enzymes to reach any entrapped lipid droplets. For example, pectin form strong gels if

there are sufficiently high levels of calcium ions present in solution (McClements & Li, 2010a).

6.3.4.3. Bile salts

The addition of bile salts to the pectin solutions had a limited influence on their physicochemical

properties. Addition of increasing amounts of bile salts only caused a relatively small decrease in

the negative charge on the pectin molecules (Figure 6.5c), which can probably be attributed to

electrostatic screening effects (Hur, Lim, Decker, & McClements, 2011; McClements & Li,

2010a). There was an appreciable increase in the optical turbidity of the pectin solutions when

bile salts were added (Figures 6.5c and 6.6a) and the solutions appeared visibly cloudier (Figure

6.7). These effects can be attributed to the formation of large aggregates within the bile salts-

pectin solutions as observed by optical microscopy (Figure 6.8). The ITC measurements also

Page 189: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

180

indicated that the bile salts interacted with the pectin molecules. In the absence of pectin, there

was a relatively large endothermic enthalpy change (H = +204 kcal/injection), which

progressively decreased with increasing bile salt concentration (Figure 6.9c). This effect may

have been due to the breakdown of bile salt micelles within the reaction chamber, resulting in the

exposure of a greater number of non-polar groups to water (Thongngam & McClements, 2005).

In the presence of pectin, there was a large endothermic peak (from 0.05 to 0.15 % w/w) followed

by a value that was close to zero at higher bile salt concentrations (after 0.2% (w/w), Figure

6.9c). The large difference between the two curves suggests that there was an interaction between

pectin and bile salts, e.g., bile salt micelles may have bound to pectin molecules and/or promoted

their aggregation (Wangsakan, Chinachoti, & McClements, 2006).

One might not expect a strong interaction between bile salts and pectin molecules because they

are both usually negatively charged at neutral pH. However, bile salts may be able to interact

with pectin molecules through hydrophobic interactions (Dongowski, 1997). In general, the

hydrophobic interactions of pectin with bile salts are known to increase with the degree of

methoxylation of pectin molecules (by increasing its hydrophobicity) and the molecular weight

(related to viscosity) of the pectin molecules (Pfeffer, Doner, Hoagland, & McDonald, 1981).

These experiments clearly show that bile salts can interact with pectin molecules, which may

have important implications for the effects of pectin on lipid digestion observed in Figure 6.2. If

the bile salts form microgel particles with pectin, then they may not be available to absorb onto

the surfaces of the lipid droplets or to solubilize lipid digestion products, as discussed earlier. In

addition, if any lipid droplets are trapped inside the microgels then the rate of digestion may be

decreased because the lipase molecules cannot easily reach the lipid droplet surfaces

(McClements & Li, 2010a).

6.3.4.4. Lipase

Finally, we examined the influence of lipase addition to the pectin solutions on their overall

physicochemical properties. The addition of increasing amounts of lipase caused a slight change

in the -potential of the pectin (Figure 6.5d), which may have been due to some binding of lipase

Page 190: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

181

molecules to the pectin molecules. In addition, there was an appreciable difference between the

ITC curves of the control and system containing pectin (Figure 6.9d), which also suggests that

some form of interaction occurred (H = -129 kcal/injection) (Espinal-Ruiz, Parada-Alfonso,

Restrepo-Sánchez, & Narváez-Cuenca, 2014). In the absence of pectin, there was a relatively

large exothermic change observed when the lipase solution was titrated into the buffer solution.

This effect can probably be attributed to the heat of dilution associated with salts in the lipase

solution. In the presence of pectin, there was a large exothermic change observed at low lipase

levels (0.0 to 0.1% w/w), followed by a much smaller exothermic change at higher lipase levels.

The large difference between the curves in the presence and absence of pectin suggests that there

was a strong interaction between the lipase and the pectin, although again the molecular origin of

this effect cannot be established from the ITC measurements. There was a slight increase in the

optical turbidity of the lipase-pectin solutions when high levels of lipase were added (Figure

6.5d), but there was little change in their overall visual appearance (Figure 6.7) or their

microstructure (Figure 6.8). These results suggest that either any interactions between the lipase

and pectin did not lead to extensive aggregation or that the lipase-pectin complexes formed were

soluble in buffer solution.

Finally, the molecular mechanisms governing the interaction between pectin and the

gastrointestinal compounds evaluated in this study (NaCl, CaCl2, bile salts, and pancreatic lipase)

are represented schematically in the Figure 6.10. The interaction between pectin and NaCl is

mainly due to electrostatic screening effects, whereas the interaction between pectin and CaCl2 is

mainly due to electrostatic crosslinking and the further formation of egg-box structures. The

interaction between pectin and bile salts can be attributed to hydrophobic interactions. In

addition, we suggested that the interaction between pectin and pancreatic lipase is mainly due to

both electrostatic and hydrophobic interactions. Furthermore, the molecular complexes formed

after the molecular interaction between pectin and the gastrointestinal compounds was observed

to be soluble for NaCl and pancreatic lipase, and insoluble for CaCl2 and bile salts, leading to the

formation of the microgel-like structures observed in the Figure 6.8.

Page 191: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

182

Figure 6.10. Molecular mechanisms governing the interaction between pectin and gastrointestinal tract

(GIT) compounds (NaCl, CaCl2, bile salts, and pancreatic lipase). The solubility of the complexes formed

after the interaction of GIT compounds with pectin is also showed (highly schematic).

ITC is an advantageous technique in terms of thermodynamic studies of biomolecular systems.

The magnitude (H) and the sign (endothermic or exothermic) of the interaction allows to

improve the understanding of the molecular effects exerted by pectin on the digestion of lipids.

The determination of the relative contribution of the individual components to the overall

interaction may lead to the design of highly structured pectin-based food systems that selectively

contribute to the control of the lipid digestion. In addition, ITC is a convenient technique for

these studies since it allows to obtain reliable information from non-covalent interactions such as

hydrophobic, hydrogen bonding, electrostatic, and van der Waals interactions, which are

responsible of the interactions between pectin and gastrointestinal components. Other analytical

techniques to evaluate interactions are currently available.

Bile salts

NaCl

Lipase

CaCl2

⊝⊝

2⊕

2⊕

2⊕

Electrostatic screeningSoluble

Electrostatic crosslinking

―Egg box structures‖Insoluble

Hydrophobic interactionsInsoluble

Both electrostatic and

hydrophobic interactionsSoluble

Page 192: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

183

High Performance Size Exclusion Chromatography (HPSEC) allows to evaluate the formation of

molecular species arising from the interaction between pectin and the components of the

gastrointestinal system. However, this technique only allows to obtain qualitative information of

the interactions, disregarding the affinities which leads to stablish the relative contribution of

each component to the overall interaction.

6.4. Conclusions

The purpose of this study was to investigate the interaction of pectin molecules with some of the

major constituents within gastrointestinal fluids, i.e., NaCl, CaCl2, bile salts, and lipase. Our

measurements have shown that a number of these components interact strongly with pectin,

which may have important implications for understanding the influence of pectin on lipid

digestion. In particular, calcium ions and bile salts appear to promote the formation of pectin

microgels that can lead to flocculation of lipid droplets in the gastrointestinal tract, decreasing the

ability of lipids to be reached by gastrointestinal lipases and thereby, inhibiting their digestibility.

If appreciable amounts of calcium and bile salts are trapped within these microgels, then they will

not be able to remove long chain FFAs from lipid droplet surfaces, which would inhibit lipid

digestion. In addition, if lipid droplets are trapped within pectin microgels, then it may be more

difficult for lipase molecules to access the surfaces of the lipid substrate, again inhibiting lipid

digestion. These results have important implications for the rational design of dietary fiber-based

functional foods that may modulate lipid digestion within the gastrointestinal tract.

Acknowledgments

The authors are grateful to COLCIENCIAS and Universidad Nacional de Colombia for providing

a fellowship to Mauricio Espinal-Ruiz supporting this work. This material is partly based upon

work supported by United States Department of Agriculture, NRI Grants (2011-03539, 2013-

03795, 2011-67021, and 2014-67021).

Page 193: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

184

References

Almgren, M. (2000). Mixed micelles and other structures in the solubilization of bilayer lipid membranes

by surfactants. Biochimica et Biophysica Acta (BBA) - Biomembranes, 1508(1–2), 146-163.

Anderson, J. W., Baird, P., Davis Jr, R. H., Ferreri, S., Knudtson, M., Koraym, A., Waters, V., &

Williams, C. L. (2009). Health benefits of dietary fiber. Nutrition Reviews, 67(4), 188-205.

Beysseriat, M., Decker, E. A., & McClements, D. J. (2006). Preliminary study of the influence of dietary

fiber on the properties of oil-in-water emulsions passing through an in vitro human digestion

model. Food Hydrocolloids, 20(6), 800-809.

Biggs, S., Habgood, M., Jameson, G. J., & Yan, Y.-d. (2000). Aggregate structures formed via a bridging

flocculation mechanism. Chemical Engineering Journal, 80(1–3), 13-22.

Braccini, I., & Pérez, S. (2001). Molecular Basis of Ca2+-Induced Gelation in Alginates and Pectins:  The

Egg-Box Model Revisited. Biomacromolecules, 2(4), 1089-1096.

Bray, G. A., & Popkin, B. M. (1998). Dietary fat intake does affect obesity! The American Journal of

Clinical Nutrition, 68(6), 1157-1173.

Caffall, K. H., & Mohnen, D. (2009). The structure, function, and biosynthesis of plant cell wall pectic

polysaccharides. Carbohydrate Research, 344(14), 1879-1900.

Chang, Y., McLandsborough, L., & McClements, D. J. (2011). Interactions of a Cationic Antimicrobial

(ε-Polylysine) with an Anionic Biopolymer (Pectin): An Isothermal Titration Calorimetry,

Microelectrophoresis, and Turbidity Study. Journal of Agricultural and Food Chemistry, 59(10),

5579-5588.

de Roos, A., & Walstra, P. (1996). Loss of enzyme activity due to adsorption onto emulsion droplets.

Colloids and Surfaces B: Biointerfaces, 6(3), 201-208.

Devraj, R., Williams, H. D., Warren, D. B., Mullertz, A., Porter, C. J. H., & Pouton, C. W. (2013). In vitro

digestion testing of lipid-based delivery systems: Calcium ions combine with fatty acids liberated

from triglyceride rich lipid solutions to form soaps and reduce the solubilization capacity of

colloidal digestion products. International Journal of Pharmaceutics, 441(1-2), 323-333.

Dickinson, E. (1995). On flocculation and gelation in concentrated particulate systems containing added

polymer. Journal of the Chemical Society-Faraday Transactions, 91(24), 4413-4417.

Dickinson, E., Semenova, M. G., Antipova, A. S., & Pelan, E. G. (1998). Effect of high-methoxy pectin

on properties of casein-stabilized emulsions. Food Hydrocolloids, 12(4), 425-432.

Dongowski, G. (1997). Effect of pH on the in vitro interactions between bile acids and pectin. Zeitschrift

für Lebensmitteluntersuchung und -Forschung A, 205(3), 185-192.

Doyle, M. L. (1997). Characterization of binding interactions by isothermal titration calorimetry. Current

Opinion in Biotechnology, 8(1), 31-35.

Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L.-P., & Narváez-Cuenca, C.-E. (2014).

Inhibition of digestive enzyme activities by pectic polysaccharides in model solutions. Bioactive

Carbohydrates and Dietary Fibre, 4(1), 27-38.

Espinal-Ruiz, M., Parada, F., Restrepo-Sanchez, L., Narvaez-Cuenca, C., & McClements, D. J. (2014).

Impact of dietary fibers [methyl cellulose, chitosan, and pectin] on digestion of lipids under

simulated gastrointestinal conditions. Food & Function.

Freire, E., Mayorga, O. L., & Straume, M. (1990). Isothermal Titration Calorimetry. Analytical Chemistry,

62(18), 950A-959A.

Furusawa, K., Ueda, M., & Nashima, T. (1999). Bridging and depletion flocculation of synthetic latices

induced by polyelectrolytes. Colloids and Surfaces, A: Physicochemical and Engineering Aspects,

153(1–3), 575-581.

Page 194: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

185

Gunness, P., & Gidley, M. J. (2010). Mechanisms underlying the cholesterol-lowering properties of

soluble dietary fibre polysaccharides. Food & Function, 1(2), 149-155.

Howarth, N. C., Saltzman, E., & Roberts, S. B. (2001). Dietary Fiber and Weight Regulation. Nutrition

Reviews, 59(5), 129-139.

Hur, S. J., Decker, E. A., & McClements, D. J. (2009). Influence of initial emulsifier type on

microstructural changes occurring in emulsified lipids during in vitro digestion. Food Chemistry,

114(1), 253-262.

Hur, S. J., Lim, B. O., Decker, E. A., & McClements, D. J. (2011). In vitro human digestion models for

food applications. Food Chemistry, 125(1), 1-12.

Israelachvili, J. (2011). Intermolecular and Surface Forces, Third Edition (Third Edition ed.). London,

UK: Academic Press.

Jenkins, P., & Snowden, M. (1996a). Depletion flocculation in colloidal dispersions. Advances in Colloid

and Interface Science, 68, 57-96.

Jenkins, P., & Snowden, M. (1996b). Depletion flocculation in colloidal dispersions. Advances in Colloid

and Interface Science, 68(0), 57-96.

Kritchevsky, D. (1988). Dietary Fiber. Annual Review of Nutrition, 8(1), 301-328.

Leavitt, S., & Freire, E. (2001). Direct measurement of protein binding energetics by isothermal titration

calorimetry. Current Opinion in Structural Biology, 11(5), 560-566.

Leroux, J., Langendorff, V., Schick, G., Vaishnav, V., & Mazoyer, J. (2003). Emulsion stabilizing

properties of pectin. Food Hydrocolloids, 17(4), 455-462.

Li, Y., Hu, M., & McClements, D. J. (2011). Factors affecting lipase digestibility of emulsified lipids

using an in vitro digestion model: Proposal for a standardised pH-stat method. Food Chemistry,

126(2), 498-505.

Li, Y., & McClements, D. J. (2014). Modulating lipid droplet intestinal lipolysis by electrostatic

complexation with anionic polysaccharides: Influence of cosurfactants. Food Hydrocolloids,

35(0), 367-374.

Maxwell, E. G., Belshaw, N. J., Waldron, K. W., & Morris, V. J. (2012). Pectin – An emerging new

bioactive food polysaccharide. Trends in Food Science & Technology, 24(2), 64-73.

McClements, D. J. (2000a). Comments on viscosity enhancement and depletion flocculation by

polysaccharides. Food Hydrocolloids, 14(2), 173-177.

McClements, D. J. (2000b). Isothermal Titration Calorimetry Study of Pectin−Ionic Surfactant

Interactions. Journal of Agricultural and Food Chemistry, 48(11), 5604-5611.

McClements, D. J., & Li, Y. (2010a). Review of in vitro digestion models for rapid screening of

emulsion-based systems. Food & Function, 1(1), 32-59.

McClements, D. J., & Li, Y. (2010b). Structured emulsion-based delivery systems: Controlling the

digestion and release of lipophilic food components. Advances in Colloid and Interface Science,

159(2), 213-228.

Méndez-Alcaraz, J. M., & Klein, R. (2000). Depletion forces in colloidal mixtures. Physical Review E,

61(4), 4095-4099.

Mohnen, D. (2008). Pectin structure and biosynthesis. Current Opinion in Plant Biology, 11(3), 266-277.

Mun, S., Decker, E., Park, Y., Weiss, J., & McClements, D. J. (2006). Influence of Interfacial

Composition on in Vitro Digestibility of Emulsified Lipids: Potential Mechanism for Chitosan's

Ability to Inhibit Fat Digestion. Food Biophysics, 1(1), 21-29.

Munarin, F., Tanzi, M. C., & Petrini, P. (2012). Advances in biomedical applications of pectin gels.

International Journal of Biological Macromolecules, 51(4), 681-689.

Pfeffer, P. E., Doner, L. W., Hoagland, P. D., & McDonald, G. G. (1981). Molecular interactions with

dietary fiber components. Investigation of the possible association of pectin and bile acids.

Journal of Agricultural and Food Chemistry, 29(3), 455-461.

Page 195: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 6

186

Reiffers-Magnani, C. K., Cuq, J. L., & Watzke, H. J. (2000). Depletion flocculation and thermodynamic

incompatibility in whey protein stabilised O/W emulsions. Food Hydrocolloids, 14(6), 521-530.

Ridley, B. L., O'Neill, M. A., & Mohnen, D. (2001). Pectins: structure, biosynthesis, and

oligogalacturonide-related signaling. Phytochemistry, 57(6), 929-967.

Salvia-Trujillo, L., Qian, C., Martin-Belloso, O., & McClements, D. J. (2013). Influence of particle size

on lipid digestion and beta-carotene bioaccessibility in emulsions and nanoemulsions. Food

Chemistry, 141(2), 1472-1480.

Simo, O. K., Mao, Y., Tokle, T., Decker, E. A., & McClements, D. J. (2012). Novel strategies for

fabricating reduced fat foods: Heteroaggregation of lipid droplets with polysaccharides. Food

Research International, 48(2), 337-345.

Singh, H., Ye, A., & Horne, D. (2009). Structuring food emulsions in the gastrointestinal tract to modify

lipid digestion. Progress in Lipid Research, 48(2), 92-100.

Thakur, B. R., Singh, R. K., Handa, A. K., & Rao, M. A. (1997). Chemistry and uses of pectin — A

review. Critical Reviews in Food Science and Nutrition, 37(1), 47-73.

Theuwissen, E., & Mensink, R. P. (2008). Water-soluble dietary fibers and cardiovascular disease.

Physiology & Behavior, 94(2), 285-292.

Thiam, A. R., Farese Jr, R. V., & Walther, T. C. (2013). The biophysics and cell biology of lipid droplets.

Nat Rev Mol Cell Biol, 14(12), 775-786.

Thongngam, M., & McClements, D. J. (2005). Isothermal titration calorimetry study of the interactions

between chitosan and a bile salt (sodium taurocholate). Food Hydrocolloids, 19(5), 813-819.

Wangsakan, A., Chinachoti, P., & McClements, D. J. (2004). Effect of Surfactant Type on

Surfactant−Maltodextrin Interactions:  Isothermal Titration Calorimetry, Surface Tensiometry,

and Ultrasonic Velocimetry Study. Langmuir, 20(10), 3913-3919.

Wangsakan, A., Chinachoti, P., & McClements, D. J. (2006). Isothermal titration calorimetry study of the

influence of temperature, pH and salt on maltodextrin–anionic surfactant interactions. Food

Hydrocolloids, 20(4), 461-467.

Weiss, J., Takhistov, P., & McClements, D. J. (2006). Functional Materials in Food Nanotechnology.

Journal of Food Science, 71(9), R107-R116.

Willats, W. G. T., Knox, J. P., & Mikkelsen, J. D. (2006). Pectin: new insights into an old polymer are

starting to gel. Trends in Food Science & Technology, 17(3), 97-104.

Winzor, D. J., Carrington, L. E., Deszczynski, M., & Harding, S. E. (2004). Extent of Charge Screening in

Aqueous Polysaccharide Solutions. Biomacromolecules, 5(6), 2456-2460.

Yapo, B. M. (2011). Pectic substances: From simple pectic polysaccharides to complex pectins—A new

hypothetical model. Carbohydrate Polymers, 86(2), 373-385.

Ye, A., Cui, J., Zhu, X., & Singh, H. (2013). Effect of calcium on the kinetics of free fatty acid release

during in vitro lipid digestion in model emulsions. Food Chemistry, 139(1–4), 681-688.

Page 196: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

Effect of pectins on the mass transfer kinetics of monosaccharides,

amino acids, and a corn oil-in-water emulsion in a Franz diffusion cell

Published as:

Espinal-Ruiz, M., Restrepo-Sánchez, L. P., & Narváez-Cuenca, C. E.

Food Chemistry. 209 (2016): 144 – 153.

Page 197: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

188

Abstract

The effect of high (HMP) and low (LMP) methoxylated pectins (2% w/w) on the rate and extent

of the mass transfer of monosaccharides, amino acids, and a corn oil-in-water emulsion across a

cellulose membrane was evaluated. A sigmoidal response kinetic analysis was used to calculate

both the diffusion coefficients (rate) and the amount of nutrients transferred through the

membrane (extent). In all cases, except for lysine, HMP was more effective than LMP in

inhibiting both the rate and extent of the mass transfer of nutrients through the membrane. LMP

and HMP, e.g., reduced 1.3 and 3.0 times, respectively, the mass transfer rate of glucose, as

compared to control (containing no pectin), and 1.3 and 1.5 times, respectively, the amount of

glucose transferred through the membrane. Viscosity, molecular interactions, and flocculation

were the most important parameters controlling the mass transfer of electrically neutral nutrients,

electrically charged nutrients, and emulsified lipids, respectively.

Keywords: Pectin, methoxylation degree, diffusion, viscosity, Franz diffusion cell.

Page 198: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

189

7.1. Introduction

There is a growing interest among consumers about the nutritional, therapeutic, and functional

properties of foods consumed in the diet (Mohamed, 2014). The current trend of consumers is

based on the consumption of foods with functional components beneficial for human health such

as phytosterols, polyphenolics, anthocyanins, carotenoids, and dietary fibers (Bigliardi & Galati,

2013). In recent years, dietary fiber has received an important attention from the food industry

and consumers due to the health benefits associated with the consumption of dietary fiber-rich

foods (Brownlee, 2014). Dietary fiber components possess distinctive physicochemical

characteristics determining their functionality, and they can be categorized as either insoluble or

soluble dietary fibers (Phillips, 2013). The consumption of insoluble dietary fiber has been

associated with increasing the bulkiness of the digesta content and improving the gastrointestinal

tract (GIT) transit (Edwards, Johnson, & Read, 1988), whereas the consumption of soluble

dietary fiber has been associated with a wider variety of physiological functions such as

controlling postprandial glycemic and lipid response (Pasquier, Armand, Guillon, Castelain,

Borel, Barry, et al., 1996), decreasing blood lipid and glucose levels (Ye, Arumugam,

Haugabrooks, Williamson, & Hendrich, 2015), inhibiting the GIT enzyme activities (Espinal-

Ruiz, Parada-Alfonso, Restrepo-Sánchez, & Narváez-Cuenca, 2014), controlling the rate and

extent of lipid digestion (Espinal-Ruiz, Parada-Alfonso, Restrepo-Sanchez, Narvaez-Cuenca, &

McClements, 2014; Espinal-Ruiz, Restrepo-Sánchez, Narváez-Cuenca, & McClements, 2016),

increasing the viscosity of the GIT content (Edwards, Johnson, & Read, 1988; Elleuch, Bedigian,

Roiseux, Besbes, Blecker, & Attia, 2011), and retarding the mass transfer process of nutritional

compounds to be absorbed by enterocytes (Fabek, Messerschmidt, Brulport, & Goff, 2014;

Srichamroen & Chavasit, 2011). Among the aforementioned properties of soluble dietary fiber,

the ability to modulate the viscosity of the GIT content stands out because this feature is related

to the control of the mobility of nutrients and their further digestion and absorption (Fabek,

Messerschmidt, Brulport, & Goff, 2014).

The functional properties of soluble dietary fibers (e.g., gums, mucilages, and pectins) rely on

their ability to thicken into swollen hydrated networks and their subsequent increasing of

Page 199: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

190

viscosity, determining their potential to exert physiological effects along the GIT, specifically

through the stomach and small intestine (Brownlee, 2014). The proposed mechanism by which

viscosity may induce physiological responses includes increasing of luminal bulk (Ye,

Arumugam, Haugabrooks, Williamson, & Hendrich, 2015) and inhibiting nutrient diffusion

across the unstirred water layer of the mucosal membrane (Fabek, Messerschmidt, Brulport, &

Goff, 2014; Gunness, Flanagan, Shelat, Gilbert, & Gidley, 2012). Viscous soluble fibers may

hinder the mass transfer process of nutritional compounds from the lumen to the unstirred water

layer coating the small intestine by reducing the mixing among them and reducing the time

available for intestinal absorption by enterocytes (Fabek, Messerschmidt, Brulport, & Goff, 2014;

Gunness, Flanagan, Shelat, Gilbert, & Gidley, 2012). Therefore, an increase in digesta viscosity

arising from the soluble dietary fiber consumption might influence the processes occurring during

the nutrient digestion and adsorption (Espinal-Ruiz, Parada-Alfonso, Restrepo-Sanchez, Narvaez-

Cuenca, & McClements, 2014).

The effect of different sources of dietary fiber [e.g. barley (Hordeum vulgare) -glucans (Gs)

and wheat (Triticum aestivum) arabinoxylans (AXs)] was evaluated in in vitro models (Fabek,

Messerschmidt, Brulport, & Goff, 2014; Gunness, Flanagan, Shelat, Gilbert, & Gidley, 2012;

Srichamroen & Chavasit, 2011). Both Gs and AXs proved to be effective in retarding both the

rate and extent of the mass transfer process of bile salts through a dialysis membrane. The

inhibition of the mass transfer process of bile salts through a dialysis membrane by increasing the

viscosity of the solution upon addition of Gs and AXs was suggested to be a possible

mechanism controlling the mass transfer process (Gunness, Flanagan, Shelat, Gilbert, & Gidley,

2012). In other study it was shown that the alkaline-extracted malva nut (Sterculia lychnophora)

gum reduced significantly the amount of glucose transferred through a dialysis membrane in an

in vitro experiment, as compared to control (containing no dietary fiber) (Srichamroen &

Chavasit, 2011).

Among the different sources of soluble dietary fiber available, pectin is a polysaccharide obtained

from fruits and vegetables and it is widely used in the pharmaceutical and food industries for its

thickening, gelling, and texturing properties (Maxwell, Belshaw, Waldron, & Morris, 2012).

Page 200: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

191

Pectin is a biopolymer mainly formed by galacturonic acid (GalA) units joined in chains by -D-

(1,4) glycosidic linkages. Three pectic structures [homogalacturan (HG), rhamnogalacturonan-I

(RG-I), and rhamnogalacturonan-II (RG-II)] have been isolated and structurally characterized

(Mohnen, 2008). HG corresponds to the linear chain of -(1,4) GalA units in which the carboxyl

group (–COO⊝) of GalA moieties can be partially esterified with methanol (methoxylated),

forming the carbomethoxyl group (–COOCH3). An important characteristic of HG, present in

pectins, is the methoxylation degree, defined as the percentage of carboxyl groups which have

been methoxylated. If more than 50% of the carboxyl groups are methoxylated, the pectin is

called high methoxylated pectin (HMP), and less than that methoxylation degree are called low

methoxylated pectin (LMP). RG-I is a pectic structure containing a linear backbone of the

disaccharide [4)--D-GalA-(12)--L-Rha-(1], where Rha corresponds to rhamnose, and

RG-II consists of a HG backbone substituted with side branches consisting of twelve different

types of monosaccharides in up to twenty different linkages (Maxwell, Belshaw, Waldron, &

Morris, 2012). The most abundant pectic structure is HG that comprises 65% (mol/mol) of

pectin, whereas RG-I and RG-II comprise 25 and 10% (mol/mol), respectively (Yapo, 2011).

Many functional properties of pectin (e.g., viscosity, solubility, and gelation capacity) are

dependent on its structural parameters such as molecular weight, methoxylation degree and the

distribution pattern of methoxylation within the GalA chains (Mohnen, 2008; Ryden,

MacDougall, Tibbits, & Ring, 2000; Yapo, 2011). In previous studies, we have demonstrated that

pectin inhibits the activity of some digestive enzymes (Espinal-Ruiz, Parada-Alfonso, Restrepo-

Sánchez, & Narváez-Cuenca, 2014) and interferes with the digestion of emulsified lipids

(Espinal-Ruiz, Parada-Alfonso, Restrepo-Sanchez, Narvaez-Cuenca, & McClements, 2014;

Espinal-Ruiz, Restrepo-Sánchez, Narváez-Cuenca, & McClements, 2016) by interacting with the

GIT components (Espinal-Ruiz, Parada-Alfonso, Restrepo-Sánchez, Narváez-Cuenca, &

McClements, 2014).

Although studies regarding the mass transfer control of some nutritional compounds have been

carried out, no information is currently available concerning the influence of pectins on the mass

transfer process of the primary nutritional compounds consumed in the diet or produced after

digestion such as monosaccharides, amino acids, and lipids. The objective of this study was,

Page 201: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

192

therefore, to determine the effect of HMP and LMP on the rate and extent of the mass transfer

process of the main nutritional compounds consumed in the diet or obtained after digestion, such

as monosaccharides, amino acids, and lipids (represented by a corn oil-in-water emulsion). The

mass transfer profiles of monosaccharides, amino acids, and a corn oil-in-water emulsion in the

absence (control) or presence of pectins (LMP or HMP) was reported and analyzed to suggest a

possible interaction mechanism between pectin and the evaluated nutritional compounds. These

results might lead to the design, formulation, and fabrication of functional foods developed to

control the postprandial blood concentration of the monosaccharides, amino acids, and lipids

consumed in the diet.

7.2. Materials and methods

7.2.1. Chemicals

D-(+)-glucose, D-(+)-galactose, D-(–)-fructose, and D-(–)-ribose were purchased from Panreac

Química SLU (Barcelona, Spain). L-lysine, L-glycine, L-aspartic acid, L-tyrosine, ninhydrin

monohydrate, Tween 80, and a dialysis tubing cellulose membrane (cut-off 14 kDa) were

purchased from Sigma-Aldrich Chemical Company (St Louis, MO, USA). Corn oil was

purchased from a commercial food supplier (Grasco LTDA, Bogotá DC, Colombia) and stored in

darkness at room temperature until use. The manufacturer reported that the corn oil contained

approximately 14, 35, and 51% (w/w) of saturated, monounsaturated, and polyunsaturated fatty

acids, respectively. Commercial powdered LMP was donated by TIC Gums Inc. (Belcamp, MA,

USA) and was used without further purification, with methoxylation degree and average

molecular weight previously reported as 30% (mol/mol) and 130 kDa, respectively (Espinal-Ruiz,

Restrepo-Sánchez, Narváez-Cuenca, & McClements, 2016). Commercial powdered HMP (Genu

Citrus Pectin USP/100) was donated by CP Kelco Co. (Lille Skensved, Denmark) and was also

used without further purification, with methoxylation degree and average molecular weight

previously reported as 71% (mol/mol) and 181 kDa, respectively (Espinal-Ruiz, Restrepo-

Sánchez, Narváez-Cuenca, & McClements, 2016). All other chemicals were purchased from

Merck KGaS (Darmstadt, Germany). Deionized water was used to prepare all solutions. Franz

Page 202: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

193

diffusion cells (Figure 7.1) were fabricated by Siliser LTDA (Bogotá DC, Colombia). The

volumes of the donor and receptor compartments were 3 and 17 mL, respectively. The area of

mass transfer (A) was 1.33 cm2.

7.2.2. Sample preparation

LMP and HMP stock solutions (4% w/w) were prepared separately by dispersing 2 g of powdered

pectins into 48 g of deionized water. These solutions were stirred at 1,000 rpm overnight at room

temperature to ensure complete dispersion and dissolution. LMP and HMP solutions were then

adjusted to pH 7.0 by using 0.1 M NaOH.

A monosaccharide stock solution (280 mM of each compound) was prepared by dissolving

together glucose, galactose, fructose, and ribose in deionized water. The monosaccharide stock

solution (2.5 mL) was then mixed with either 2.5 mL deionized water (control)

Figure 7.1. Structural design of a Franz diffusion cell. Donor compartment with a nominal volume of 3

mL (a), receptor compartment with a nominal volume of 17 mL (b), position for a semipermeable

cellulose membrane (c) with an effective area of mass transfer (A) of 1.33 cm2 (d = 1.30 cm), sampling

port (d), magnetic stir bar (e), thermostated chamber at 37 °C (f), water inlet (g), and water outlet (h).

a

b

e

f

g

h

c d

Page 203: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

194

or 2.5 mL of 4% (w/w) pectin stock solutions (LMP or HMP), to obtain monosaccharide working

solutions containing 140 mM of each monosaccharide and 2% (w/w) pectin. Monosaccharide

working solutions were then adjusted to pH 7.0 by using 0.1 M NaOH.

Amino acid stock solutions (4 mM of each compound) were prepared by dissolving separately

lysine, glycine, aspartic acid, or tyrosine in deionized water. Each amino acid stock solution (2.5

mL) was then mixed separately with either 2.5 mL deionized water (control) or 2.5 mL of 4%

(w/w) pectin stock solutions (LMP or HMP), to obtain amino acid working solutions containing 2

mM of each amino acid and 2% (w/w) pectin. Each amino acid working solution was then

adjusted to pH 7.0 by using 0.1 M NaOH.

A stock emulsion (20% w/w) was prepared by mixing together 20 g corn oil and 80 g buffered

emulsifier aqueous solution (5 mM phosphate buffer pH 7.0, containing 2.5% (w/w) Tween 80)

for 5 min using an Ultra-Turrax homogenizer (Speed 35,000 min-1

; Model Miccra D-9; ART

Prozess & Labortechnik GmbH & Co. KG; Müllheim; Germany). A dilution was prepared by

mixing 10 g stock emulsion with 90 g buffer solution pH 7.0, to obtain a diluted emulsion

containing 2% (w/w) corn oil. Finally, the diluted emulsion (2.5 mL) was mixed with either 2.5

mL deionized water (control) or 2.5 mL of 4% (w/w) pectin stock solutions (LMP or HMP), to

obtain working emulsions containing 1% (w/w) corn oil and 2% (w/w) pectin samples. Each

working emulsion was then adjusted to pH 7.0 by using 0.1 M NaOH.

7.2.3. Effect of pectins on the mass transfer kinetics of monosaccharides, amino acids and a

corn oil-in-water emulsion in a Franz diffusion cell

Seventeen milliliters of deionized water (adjusted to pH 7.0) were transferred to the receptor

compartment of the Franz diffusion cell (Figure 7.1), and incubated at 37 °C for 15 min. Then, 3

mL of the thermostated working solutions (37 °C) containing monosaccharides, amino acids, or

the corn oil-in-water emulsion, with the absence (control) or presence of pectin samples (LMP or

HMP), were transferred separately to the donor compartment. The Franz diffusion cell was

thermostated at 37 ° C throughout the experiment. At regular times intervals (0, 2, 4, 6, 8, 10, 24,

Page 204: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

195

26, 28, 30, 32, 34, and 48 h) an aliquot of 500 μL was collected from the receptor compartment

and then, 500 μL of deionized water (adjusted to pH 7.0 and thermostated to 37 °C) were added

immediately to reconstitute the total volume of the cell. Samples collected from the Franz

diffusion cell over time were analyzed for their monosaccharide, amino acid or emulsion

contents, as described below. Contents of monosaccharides, amino acids, and the corn oil-in-

water emulsion were corrected by the dilution factor introduced at each sampling time.

7.2.4. Analysis of monosaccharides, amino acids, and the corn oil-in-water emulsion

7.2.4.1. Analysis of monosaccharides

Monosaccharide samples collected from the receptor compartment were analyzed by high

performance liquid chromatography coupled to a refractive index detector. An UHPLC+ Focused

Dionex Ultimate 3000 liquid chromatograph (Thermo Scientific Inc., Waltham, MA, USA)

equipped with an ion exchanger resin column [MetaCarb Ca Plus column (300 × 7.8 mm × 9 μm,

Agilent Technologies, Santa Clara, CA, USA)] was used. The column was operated at 70 °C and

at a flow rate of 0.5 mL min-1

with deionized water (adjusted to pH 7.0) as eluent.

Monosaccharide samples were centrifuged (18,000 g; 20 min; 4 °C) and then 10 μL of the clear

supernatant was injected into the column. Components eluting from the column were detected

using a Shodex RI-101 refractive index detector (Showa Denko Corporation, Tokyo, Japan)

thermostated at 35 °C. The amount of monosaccharides transported through the cell was

determined using an external standard method. Calibration lines were obtained at concentrations

ranging from 2 to 20 mM for each monosaccharide (n=10; r2=0.995, 0.998, 0.999, and 0.990 for

glucose, galactose, fructose, and ribose, respectively).

7.2.4.2. Analysis of amino acids

Amino acid samples collected from the receptor compartment (500 μL) were mixed with 500 μL

of ninhydrine solution [2% (w/v) prepared in 100 mM citrate buffer pH 4.0] and then incubated at

92 °C (boiling water) for 10 min to generate the ninhydrine chromophore (Ruhemann´s purple)

Page 205: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

196

(Friedman, 2004). The mixtures were cool down to room temperature and then the absorbance

was recorded at 570 nm with a Genesys 10uv Spectrophotometer (Thermo Scientific Inc.,

Waltham, MA, USA). The amount of amino acids transported through the cell was determined

using an external standard method. Calibration lines were obtained at concentrations ranging

from 35 to 350 μM for each amino acid (n=10; r2=0.997, 0.989, 0.992, and 0.995 for glycine,

aspartic acid, lysine, and tyrosine, respectively).

7.2.4.3. Analysis of the optical turbidity of the corn oil-in-water emulsion

The optical turbidity (at 600 nm) of the emulsion samples collected from the receptor

compartment was measured using a Genesys 10uv Spectrophotometer (Thermo Scientific Inc.,

Waltham, MA, USA) at room temperature. The amount of the emulsion transported through the

cell was determined using an external standard method. A calibration line was obtained at

concentrations of the corn oil-in-water emulsion ranging from 0.02 to 0.20% (w/w) (n=10;

r2=0.998).

7.2.5. Kinetic model

A sigmoidal response kinetic model (Equation 7.1) was used to fit the experimental data of the

mass transfer profiles of monosaccharides, amino acids, and a corn oil-in-water emulsion through

the Franz diffusion cell. Both the effective diffusion coefficients (the rate of the mass transfer

process) and the maximum amount of the compound transferred through the Franz diffusion cell

(the extent of the mass transfer process), in the absence (control) or presence of 2% (w/w) HMP

or LMP were calculated. This model represented by Equation 7.1 assumes that the rate of the

mass transfer process is directly proportional to the concentration of the nutritional compound to

be transferred across the membrane (Edwards, Johnson, & Read, 1988), as follows:

(

) (7.1)

Page 206: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

197

Where C is the concentration of the nutritional compound transferred at time t, k is the steepness

of the curve, and CMax is the maximum observable value for C. The relative transfer percentage

[RT (%)] of the nutritional compound can be conveniently expressed as the ratio of C and CMax.

The integration of this first-order differential equation leads to Equation 7.2, as follows:

( ( ( )))

(7.2)

Where RT (%) corresponds to the relative transfer percentage of the nutritional compound, and

t1/2 represents the time in which the 50% of the total concentration of the nutritional compound

has been transferred across the membrane. In addition, the effective diffusion coefficient (DEff)

can be defined as follows:

A (7.3)

Where A corresponds to the effective area of the mass transfer process.

7.2.6. Apparent viscosity of pectin solutions

Pectin solutions (2% w/w) for apparent viscosity measurements were prepared by dissolving 400

mg pectin samples (LMP or HMP) in 19.6 g of 25 mM NaCl aqueous solution with gentle stirring

for 18 h at room temperature. Then, pectin solutions were centrifuged (3,000 g; 15 min; 4 °C) to

devoid the entrapped air bubbles. Afterward, pectin solutions were stored at 4 °C overnight

before measurements. The apparent viscosity measurements of pectin solutions were determined

using an Ares Discovery HR-1 Rheometer (TA Instruments, New Castle, DE, USA) with a

parallel plate geometry with a diameter of 40 mm and a gap of 1 mm. Samples were placed in a

temperature-controlled Peltier plate and allowed to equilibrate at 25 °C for 5 min prior to

conducting the measurements. All apparent viscosity measurements were performed using a

series of fixed shear rates that consecutively increased from 0.002 to 10 s-1

and recorded at 25 °C.

Page 207: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

198

The Power Law Model (Equation 7.4) was used to analyze the rheological profiles of pectin

samples (Marcotte, Taherian Hoshahili, & Ramaswamy, 2001) by calculating both the K and n

values, as follows:

(7.4)

Where is the apparent viscosity, is the shear rate, K is the consistency index (which

corresponds to the apparent viscosity of a fluid behaving as Newtonian), and n is the flow index

(which indicates the degree of deviation from Newtonian behavior).

7.2.7. Emulsion characterization

7.2.7.1. Microstructure

The microstructure of the emulsions was characterized by optical microscopy. An optical

microscopy (C1 Digital Eclipse, Nikon Co., Tokyo, Japan) with a 60x objective lens was used to

capture images of the emulsions. A small aliquot of the emulsions (5 μL) was transferred to a

glass microscope slide and covered with a glass cover slip. Then, the cover slip was fixed to the

slide using sealing resin (Ted Pella Inc., Redding, CA, USA) to avoid evaporation. Next, a small

amount of immersion oil (Type A, Nikon Co., Melville, NY, USA) was placed on the top of

cover slip. All optical microscopy images were recorded by using the instrument software (EZ

CS1 version 3.8, Nikon Co., Melville, NY, USA).

7.2.7.2. Particle size distribution

The particle size distribution of emulsions was measured using a dynamic light scattering

instrument (Zetasizer Nano ZSP, Malvern Instruments Ltd., Worcestershire, United Kingdom).

Refractive indices of 1.467 (corn oil) and 1.333 (water) were used for the calculations of the

particle size distribution. Particle sizes were reported as particle size distribution profiles [volume

fraction (% v/v) vs. particle diameter (nm)].

Page 208: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

199

7.2.8. Data analysis

All experiments were performed at least three times using freshly prepared solutions. Apparent

viscosity measurements of pectin solutions were performed by quadruplicate. Averages and

standard deviations were calculated from these replicated measurements. Statistical analysis was

performed by using STATGRAPHICS Centurion XVI version 16.1.11 for Windows. A Fisher´s

Least Significance Difference (LSD) test was conducted to detect any significant differences

among the pectin samples. A p-value<0.05 was considered as statistical significance.

7.3. Results

7.3.1. Mass transfer profiles of monosaccharides

The functional properties of soluble dietary fibers (e.g., gums, mucilages, and pectins) rely on

their ability to thicken into swollen hydrated networks and their subsequent increasing of

viscosity, determining their potential to exert physiological effects along the GIT, specifically

through the stomach and small intestine. Pectins may hinder the mass transfer process of

nutritional compounds from the lumen to the unstirred water layer coating the small intestine by

reducing the mixing among them and reducing the time available for intestinal absorption by

enterocytes. Therefore, an increase in digesta viscosity arising from the pectin consumption

might influence the processes occurring during the nutrient digestion and adsorption. The results

obtained in these experiments showed that pectins, especially HMP, were able to inhibit both the

rate and extent of the mass transfer process of monosaccharides through the Franz diffusion cell

(especially the extent rather than rate). Figure 7.2 shows the mass transfer profiles of

monosaccharides [glucose (a), galactose (b), fructose (c), and ribose (d)]. Both the rate and

extent of the mass transfer process of monosaccharides through a Franz diffusion cell in the

absence (control) or presence of pectin (LMP or HMP) were characterized by calculating DEff and

maximum RT (%) values, respectively (Table 7.1). The maximum RT (%) value is related to the

amount of nutrients which can be effectively transferred through the cellulose membrane

Page 209: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

200

Figure 7.2. Effect of low (LMP) and high (HMP) methoxylated pectins on the mass transfer kinetics of

glucose (a), galactose (b), fructose (c), and ribose (d) through a Franz diffusion cell. Control corresponds

to the sample without addition of pectin. Points correspond to experimental data and lines correspond to

fitted data according to Equation 7.2.

(the extent of the mass transfer process), whereas the DEff value is related to the rate at which the

mass transfer process occurs. The RT (%) values increased with time and tended towards a

plateau after 36 h for each of the tested monosaccharides (Figure 7.2). When a pectin sample

(LMP or HMP) was added, all of the tested monosaccharides diffused slower as compared to

control (Table 7.1, DEff), reaching a lower maximum RT (%) value [Table 7.1, maximum RT

(%)]. The inhibitory effect of the mass transfer process was higher with HMP than LMP.

0

20

40

60

80

100

0 12 24 36 48

Rel

ati

ve

Tra

nsf

er (

%)

Time (h)

Control

LMP

HMP

0

20

40

60

80

100

0 12 24 36 48

Rel

ati

ve

Tra

nsf

er (

%)

Time (h)

Control

LMP

HMP

0

20

40

60

80

100

0 12 24 36 48

Rel

ati

ve

Tra

nsf

er (

%)

Time (h)

Control

LMP

HMP

0

20

40

60

80

100

0 12 24 36 48

Rel

ati

ve

Tra

nsf

er (

%)

Time (h)

Control

LMP

HMP

a. b.

c. d.

Page 210: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

201

Table 7.1. Effect of low (LMP) and high (HMP) methoxylated pectins on the effective diffusion coefficient (DEff) and the maximum relative

transfer percentage [maximum RT (%)] of monosaccharides, amino acids, and a corn oil-in-water emulsion through a Franz diffusion cell. Control

corresponds to the sample without addition of pectin. Different letters indicate significant differences. Lowercase letters indicate significant

differences of the same compound among treatments (Control, LMP, and HMP; comparison between columns), and uppercase letters indicate

significant differences of the same treatment among the type of compound (comparison within columns) at the p<0.05 level.

Compound DEff x10

5 (cm

2 s

-1)1 Maximum RT (%)

Control LMP HMP Control LMP HMP

Monosaccharides

Glucose 1.44 0.04a, B

1.14 0.04b, A

0.48 0.04c, A

94.7 0.9a, A

71.5 2.4b, B

63.3 1.6c, B

Galactose 1.44 0.04a, B

1.18 0.04b, A

0.48 0.04c, A

94.5 0.7a, A

68.5 1.6b, B

62.4 3.7c, B

Fructose 1.40 0.04a, B

1.14 0.04b, A

0.48 0.04c, A

96.0 1.6a, A

67.4 1.5b, B

60.6 2.1c, B

Ribose 3.91 0.41a, A

1.11 0.11b, A

0.52 0.04c, A

80.4 0.8b, B

86.1 3.8a, A

86.0 1.7a, A

Amino acids

Glycine 3.83 0.04c, B

4.65 0.15b, B

5.24 0.04a, B

87.8 3.3a, B

17.2 1.3b, C

10.6 0.7c, B

Aspartic acid 3.87 0.18a, B

3.91 0.26a, C

4.20 0.04a, C

96.0 1.8a, A

81.6 2.6b, A

39.3 1.9c, A

Lysine 4.06 0.04a, A

0.59 0.07c, D

3.54 0.15b, D

90.8 1.2a, B

7.0 0.7c, D

40.2 2.7b, A

Tyrosine 3.87 0.04b, B

6.75 0.29a, A

6.45 0.04a, A

95.2 1.7a, A

69.4 3.4b, B

36.7 1.8c, A

Lipids

Emulsion 3.47 0.26a 3.43 0.15

a 3.24 0.15

a 82.9 1.9

a 43.1 0.3

b 6.8 0.1

c

1Example: for Glucose (Control): DEff x 105 (cm2 s-1) = 1.44 should be read as DEff = 1.44 x 10-5 cm2 s-1.

Page 211: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

202

For example, for the diffusion of glucose (Figure 7.2a and Table 7.1), DEff values were 1.44 x

10-5

, 1.14 x 10-5

, and 0.48 x 10-5

cm2 s

-1 for control, LMP, and HMP, respectively; and the

maximum RT (%) values were 94.7, 71.5, and 63.3% for control, LMP, and HMP, respectively.

Similar trends were observed for galactose (Figure 7.2b), fructose (Figure 7.2c), and ribose

(Figure 7.2d).

7.3.2. Mass transfer profiles of amino acids

As observed with the mass transfer of monosaccharides, the rate and the extent of amino acids

transported towards the Franz diffusion cell were reduced by the presence of pectins, with a

higher effect of HMP towards glycine, aspartic acid, and tyrosine, and a higher effect of LMP

towards lysine. Figure 7.3 shows the mass transfer profiles of amino acids [glycine (a), aspartic

acid (b), lysine (c), and tyrosine (d)]. Although the mass transfer profiles obtained for the amino

acids were fairly similar to those obtained for monosaccharides, a very important difference in

the mass transfer profiles of amino acids was observed: a sigmoidal kinetic behavior in the

control and pectin containing experiments was obtained. A lag time of approximately 10 h was

observed for the mass transfer process of each of the tested amino acids in both absence (control)

and presence of pectins (LMP and HMP). Before the lag time (t<10 h), no significant differences

in the RT (%) values were observed among pectin samples for each of the tested amino acids, as

compared to control (the amount of amino acids transferred through the membrane was less than

3% after 10 h in both absence and presence of pectins). The RT (%) values increased with time

and tended towards a plateau after 36 h for each of the tested amino acids (Figure 7.3). When a

pectin sample (LMP or HMP) was added, different results for glycine, aspartic acid, and tyrosine

were observed, in comparison to lysine.

Upon addition of pectins, glycine, aspartic acid, and tyrosine diffused faster (Table 7.1, DEff), but

they reached a lower maximum RT (%) value, as compared to control [Table 7.1, maximum RT

(%)]. For example, for the diffusion of glycine (Figure 7.3a and Table 7.1), DEff values were

3.83 x 10-5

, 4.65 x 10-5

, and 5.24 x 10-5

cm2 s

-1 for control, LMP, and HMP, respectively; and the

maximum RT (%) values were 87.8, 17.2, and 10.6% for control, LMP, and HMP, respectively.

Page 212: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

203

Figure 7.3. Effect of low (LMP) and high (HMP) methoxylated pectins on the mass transfer kinetics of

glycine (a), aspartic acid (b), lysine (c), and tyrosine (d) through a Franz diffusion cell. Control

corresponds to the sample without addition of pectin. Points correspond to experimental data and lines

correspond to fitted data according to Equation 7.2.

Similar trends were observed for aspartic acid (Figure 7.3b and Table 7.1) and tyrosine (Figure

7.3d and Table 1). These results indicate that pectins, especially HMP, were able to inhibit the

amounts of glycine, aspartic acid, and tyrosine that can be effectively transferred through the

Franz diffusion cell (the extent of the mass transfer process), but they were not able to reduce the

rate at which the mass transfer process occurs.

0

20

40

60

80

100

0 12 24 36 48

Rel

ati

ve

Tra

nsf

er (

%)

Time (h)

Control

LMP

HMP

a.

0

20

40

60

80

100

0 12 24 36 48

Rel

ati

ve

Tra

nsf

er (

%)

Time (h)

Control

LMP

HMP

b.

0

20

40

60

80

100

0 12 24 36 48

Rel

ati

ve

Tra

nsf

er (

%)

Time (h)

Control

LMP

HMP

c.

0

20

40

60

80

100

0 12 24 36 48

Rel

ati

ve

Tra

nsf

er (

%)

Time (h)

Control

LMP

HMP

d.

Page 213: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

204

Figure 7.4. Effect of low (LMP) and high (HMP) methoxylated pectins on the mass transfer kinetics of a

corn oil-in-water emulsion through a Franz diffusion cell (a). Control corresponds to the sample without

addition of pectin. Points correspond to experimental data and lines correspond to fitted data according to

Equation 7.2. Particle size distribution of a corn oil-in-water emulsion in presence of LMP and HMP (b).

Microstructure of a corn oil-in-water emulsion observed by optical microscopy in presence of LMP and

HMP (c). The scale bars corresponds to 20 μm. Control corresponds to the emulsion without addition of

pectin.

In addition, it was observed that the inhibitory effect of the extent of the mass transfer process of

glycine, aspartic acid, and tyrosine through the Franz diffusion cell was higher with HMP than

LMP. Opposite to the behavior observed for glycine, aspartic acid, and tyrosine, the effect of

pectins towards the diffusion of lysine (Figure 7.3c and Table 7.1) was quite different. For this

amino acid, DEff were 4.06 x 10-5

, 0.59 x 10-5

, and 3.54 x 10-5

cm2 s

-1 for control, LMP, and HMP,

0

5

10

15

1 10 100 1000 10000V

olu

me

fra

ctio

n (

%)

Particle diameter (nm)

Control

LMP

HMP

0

20

40

60

80

100

0 12 24 36 48

Rel

ati

ve

tra

nsf

er (

%)

Time (h)

Control

LMP

HMP

a. b.

Control LMP HMP

c.

Page 214: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

205

respectively; and the maximum RT (%) values were 90.8, 7.0, and 40.2% for control, LMP, and

HMP, respectively. These results indicate that pectins, especially LMP, were able to inhibit both

the amount of lysine that can be effectively transferred through the Franz diffusion cell and the

rate at which the mass transfer process of lysine occurs. In addition, it was observed that the

inhibitory effect of the rate and extent of the mass transfer process of lysine through the Franz

diffusion cell was higher with LMP than HMP.

7.3.3. Mass transfer profile of the corn oil-in-water emulsion

Pectins, especially HMP, were able to inhibit both the rate and extent of the mass transfer process

of the corn oil-in-water emulsion through the Franz diffusion cell. Figure 7.4 shows the mass

transfer profile of the corn oil-in-water emulsion through the Franz diffusion cell. The mass

transfer profile obtained for the emulsion was fairly similar to that obtained for the

monosaccharides: the sigmoidal behavior was not observed and the effect of pectins on the rate

and extent of the mass transfer process followed the same order: Control>LMP>HMP. It was

observed that both the rate and extent of the mass transfer process of the emulsion was inhibited

upon addition of pectins (DEff were 3.47 x 10-5

, 3.43 x 10-5

, and 3.24 x 10-5

cm2 s

-1 for control,

LMP, and HMP, respectively; and the maximum RT (%) values were 82.9, 43.1, and 6.8 for

control, LMP, and HMP, respectively), being the inhibitory effect of the rate and extent of the

mass transfer process of the emulsion through the Franz diffusion cell higher with HMP than

LMP.

7.4. Discussion

7.4.1. Mass transfer profiles of electrically neutral nutrients

The inhibition of both the rate and extent of the mass transfer process of electrically neutral

nutrients (monosaccharides, glycine, and tyrosine) observed upon addition of pectins might be

due to the high viscosity of both LMP and HMP as compared to that of control. The apparent

viscosity profiles of 2% (w/w) LMP and HMP solutions are shown in Figure 7.5.

Page 215: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

206

Figure 7.5. Rheological profiles (apparent viscosity versus shear rate) measured at 25 °C of 2% (w/w) low

(LMP) and high (HMP) methoxylated pectin solutions prepared in 25 mM NaCl aqueous solution. Pectin

solutions were thermally stabilized at 25 °C for 5 min prior to analysis. Points correspond to experimental

data and lines correspond to fitted data according to Equation 7.4.

Both LMP and HMP solutions behaved as pseudoplastic non-Newtonian fluids because the

viscosity of the solutions decreased as the shear rate increased (Sato, Oliveira, & Cunha, 2008).

The apparent viscosity of HMP was significantly higher than the one of LMP, at low shear rates

(below to 0.01 s-1

). In addition, the flow behavior indexes for both LMP and HMP solutions were

less than 1 (n=0.292 for HMP and n=0.183 for LMP), demonstrating that pectin solutions

behaved as non-Newtonian fluids [it has been defined for Newtonian fluids, n=1 (Cross, 1965)].

Due to the high molecular weight of HMP (181 kDa) as compared to that of LMP (130 kDa), the

consistency index of HMP (281 mPa s) solution was significantly higher than that of LMP (62

mPa s). The non-Newtonian behavior of pectin solutions aggress with the presence of

macroscopic network structures which are characteristic of the hydration process of pectin

molecules (Löfgren, Walkenström, & Hermansson, 2002). Viscous fibers with hydrophilic nature

such as pectins have been reported to create complex networks by the entanglements of fully

hydrated chains of the polymer producing viscous solutions (Ryden, MacDougall, Tibbits, &

0

5000

10000

15000

0.001 0.01 0.1 1 10

Ap

pa

ren

t V

isco

sity

(m

Pa

s)

Shear Rate (s-1)

HMP

LMP

Page 216: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

207

Ring, 2000). The formation of these complex hydrated networks is commonly related to the

capacity of pectin molecules to embed nutritional compounds, thereby limiting their further

mobility (Zsivanovits, MacDougall, Smith, & Ring, 2004). Therefore, the retardation of the rate

and extent of the mass transfer process of electrically neutral nutrients (monosaccharides,

glycine, and tyrosine) through the Franz diffusion cell was higher with HMP than LMP probably

due to the high viscosity of HMP as compared to that of LMP.

It is important to consider that amino acids can be electrically charged depending on the pH, and

this electrical charge will determine their potential to interact with other charged species (such as

pectin). The net electrical charge for glycine and tyrosine was calculated by using the relative

distribution of their charged species at pH 7.0. For glycine, the neutral specie (charge 0,

⊕NH3CH(H)COO⊝) has a relative abundance of 99.8%; and for tyrosine, the neutral specie

(charge 0, ⊕NH3CH(CH2OH)COO⊝) has a relative abundance of 99.4%. These calculations

confirm that both glycine and tyrosine are neutral amino acids at pH 7.0. Because of the neutral

structure of glycine and tyrosine at pH 7.0, the electrostatic interactions between them and pectin

were not significant. Therefore, the viscosity of the pectin solutions might be the parameter

governing the retardation of the mass transfer process of all the studied electrically neutral

nutrients. For electrically neutral nutrients, however, other structural parameters such as

hydrophobicity and molecular size might probably be involved in the interactions with pectins.

Nevertheless, the effect of these parameters on the mass transfer process of electrically neutral

nutrients was unclear.

7.4.2. Mass transfer profiles of electrically charged amino acids

We suggest that the mass transfer process of electrically charged amino acids (aspartic acid and

lysine) can be influenced by both the viscosity of the pectin solutions and the electrostatic

interactions between each charged amino acid and the surface electrical charge of LMP and

HMP. It is well known that pectin molecules are anionic in nature (Mohnen, 2008) because of the

spontaneous ionization of the carboxyl group in aqueous solution (–COOH + H2O –COO⊝ +

H3O⊕). The surface electrical charge of the pectins tested in this study were previously

Page 217: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

208

characterized (Espinal-Ruiz, Restrepo-Sánchez, Narváez-Cuenca, & McClements, 2016). LMP

possess a higher surface negative charge ( = -47.5 mV at pH 7.0) as compared to that of HMP

( = -28.2 mV at pH 7.0) because a higher fraction of the carboxyl groups (–COO⊝) of LMP

remain free than in HMP (Espinal-Ruiz, Restrepo-Sánchez, Narváez-Cuenca, & McClements,

2016). Furthermore, the electrical net charge of aspartic acid and lysine at pH 7.0 is -1 and +1,

respectively (Moore, 1985).

We suggest, therefore, that the repulsive electrostatic interactions between the carboxyl group of

aspartic acid (Figure 7.3b) and the carboxyl groups of LMP prevented aspartic acid molecules to

be trapped in the structural network of LMP, and therefore, the extent of the mass transfer

process of aspartic acid through the Franz diffusion cell was not significantly hindered upon

addition of LMP. In contrast, because of the lower surface negative charges of HMP as compared

to that of LMP, the electrostatic interactions between aspartic acid and HMP were not significant,

and therefore, the inhibition of the extent of the mass transfer process observed for aspartic acid

upon addition of HMP can be attributed to viscosity effects. Conversely, the attractive

electrostatic interactions between the -amino group of lysine (–NH3⊕) and the carboxyl group of

LMP (–COO⊝) promoted significantly the entrapment of lysine molecules in the structural

network of LMP, and therefore, both the rate and extent of lysine through the Franz diffusion cell

were significantly hindered upon addition of LMP. Moreover, our results suggest that the

electrostatic attractive interactions overcomed the viscosity effects and consequently, the capacity

of LMP to inhibit the mass transfer process of lysine through the Franz diffusion cell was higher

as compared to that of HMP (Table 7.1).

We hypothesize that positive charges of lysine molecules play an important role controlling the

interaction of them with pectins because lysine was the only amino acid evaluated where the

effect of LMP was higher than HMP (Figure 7.3c). In addition, we suggest that the lag phase and

the sigmoidal behavior observed in the diffusion profiles of all of the tested amino acids (Figure

7.3) were attributed to the strong electrostatic interactions between the ionizable groups of the

amino acids (–NH3⊕, –COO⊝, and side chains) and the cellulose molecules composing the

membrane (Mahrenholz, Tapia, Stigler, & Volkmer, 2010).

Page 218: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

209

7.4.3. Diffusion profile of the corn oil-in-water emulsion

The mechanism by which pectin is able to retard the rate and extent of the mass transfer process

of the emulsion might be related to the high capacity of HMP to flocculate the lipid droplets of

the emulsion as compared to LMP (Espinal-Ruiz, Restrepo-Sánchez, Narváez-Cuenca, &

McClements, 2016).

Figure 7.6. Molecular mechanisms governing the interaction between pectin and nutritional compounds

(monosaccharides, amino acids, and emulsified lipids) and its effect on the mass transfer process of them

through the Franz diffusion cell (highly schematic). The mass transfer process of monosaccharides and

neutral amino acids (glycine and tyrosine) is controlled by the viscosity of the pectin solutions, whereas

the mass transfer process of charged amino acids is controlled by both viscosity and repulsive (aspartic

acid) and attractive (lysine) electrostatic interactions between them and pectin. Finally, pectin induces

flocculation of lipid droplets by depletion attraction and therefore, their mobility is reduced.

⊝⊝

⊝⊝

⊝⊝⊝

Monosaccharides,

glycine, and tyrosine

Viscosity

Lipids

Flocculation

Lysine

Electrostatic attraction

Aspartic acid

Electrostatic repulsion

Viscosity

Flow

Pectin

Membrane

Compound

Direction and magnitude

of the mass transfer

Page 219: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

210

The depletion interactions might be the driving force which promotes the flocculation of the

emulsified lipids, and it has been established that this force induced by polysaccharides increases

as the molecular weight of the polysaccharide increases (McClements, 2000). Consequently,

because the molecular weight of HMP (181 kDa) is higher than that of LMP (130 kDa), the

flocculation capacity of HMP was greater than that of LMP. Therefore, the inhibition of the rate

and extent of the mass transfer process of the emulsion through the Franz diffusion cell was

greater with HMP (6.8% emulsion was transferred) as compared to LMP (43.1% emulsion was

transferred), because of the higher capacity of HMP to flocculate the lipid droplets composing the

emulsion. The particle size distribution analysis (Figure 7.4b) indicated that the control emulsion

(without addition of pectin) contained small lipid droplets (the lipid droplet diameter was around

300 nm) with a monomodal distribution (only one peak was observed), suggesting that the lipid

droplets in the control emulsion were stable to aggregation (Figure 7.4c, Control). However, the

multimodal distribution of the emulsions containing LMP and HMP (Figure 7.4b) indicated that

pectins were able to induce aggregation of the lipid droplets (Figures 7.4c, LMP and HMP)

through a mechanism of flocculation [lipid droplet (flocs) diameters were 1100 and 1500 nm for

LMP and HMP, respectively]. Therefore, it can be suggested that the inhibitory effect of the rate

and extent of the mass transfer process of the corn oil-in-water emulsion through the Franz

diffusion cell was higher with HMP than LMP probably due to the high capacity of HMP to

flocculate the lipid droplets of the emulsion as compared to LMP (Figures 7.4b and 7.4c).

All in all, it can be suggested that the chemical nature of the nutritional compound

(monosaccharides, amino acids, and a corn oil-in-water emulsion) to be transferred through the

Franz diffusion cell will determine the mechanism by which the nutritional compound is able to

interact with pectin molecules (Figure 7.6). For example, it was observed that viscosity was the

most important parameter governing both the rate and extent of the mass transfer process of

monosaccharides and neutral amino acids (glycine and tyrosine), whereas electrostatic

interactions played an important role controlling the extent of the mass transfer process of

charged amino acids (repulsive for the interaction between pectin and aspartic acid, and attractive

for the interaction between pectin and lysine). Furthermore, the complex aggregated structures

formed by the flocculation of emulsified lipids leads to complex interactions between them and

Page 220: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

211

pectin molecules, being flocculation the most important parameter governing both the rate and

extent of the mass transfer process of the corn oil-in-water emulsion.

7.5. Conclusions

Viscosity of pectin solutions and electrostatic interactions between pectins and nutrients are

proposed to be the main factors influencing both the rate and extent of the mass transfer process

of nutrients across a cellulose membrane in a Franz diffusion cell. Our results suggest that the

higher the viscosity of the pectin solution, the higher the inhibition of the rate and extent of the

mass transfer of nutrients through the Franz diffusion cell. In addition, it was established that

HMP was more effective than LMP to inhibit the extent of the mass transfer process of all

evaluated nutritional compounds, except for lysine, where its strong attractive electrostatic

interaction with LMP was the dominant factor governing its mass transfer process. Our results

suggest that the inclusion of pectin in food formulations might be an effective strategy for

controlling the calorie intake by limiting the digestion and absorption of nutritional compounds.

However, the formulation of foodstuffs with high pectin contents must be done carefully in

countries whose inhabitants are deficient in minerals such as iron, calcium, and zinc, since the

consumption of pectin has been associated with a decreased bioavailability of these minerals.

Acknowledgments

We are grateful to Departamento Administrativo de Ciencia, Tecnología e Innovación de

Colombia (COLCIENCIAS) and Vicerrectoría Académica of Universidad Nacional de Colombia

for providing a fellowship to Mauricio Espinal-Ruiz supporting this work. We are also grateful to

Vicerrectoría de Investigaciones of Universidad Nacional de Colombia for fundig the project

Efectos moleculares de la pectina sobre el metabolismo de lípidos y carbohidratos (Código

Hermes 20610).

Page 221: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

212

References

Bigliardi, B., & Galati, F. (2013). Innovation trends in the food industry: The case of functional foods.

Trends in Food Science & Technology, 31(2), 118-129.

Brownlee, I. (2014). The impact of dietary fibre intake on the physiology and health of the stomach and

upper gastrointestinal tract. Bioactive Carbohydrates and Dietary Fibre, 4(2), 155-169.

Cross, M. M. (1965). Rheology of non-Newtonian fluids: A new flow equation for pseudoplastic systems.

Journal of Colloid Science, 20(5), 417-437.

Edwards, C. A., Johnson, I. T., & Read, N. W. (1988). Do viscous polysaccharides slow absorption by

inhibiting diffusion or convection? European journal of clinical nutrition, 42(4), 307-312.

Elleuch, M., Bedigian, D., Roiseux, O., Besbes, S., Blecker, C., & Attia, H. (2011). Dietary fibre and

fibre-rich by-products of food processing: Characterisation, technological functionality and

commercial applications: A review. Food Chemistry, 124(2), 411-421.

Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L.-P., & Narváez-Cuenca, C.-E. (2014).

Inhibition of digestive enzyme activities by pectic polysaccharides in model solutions. Bioactive

Carbohydrates and Dietary Fibre, 4(1), 27-38.

Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sanchez, L.-P., Narvaez-Cuenca, C.-E., & McClements,

D. J. (2014). Impact of dietary fibers [methyl cellulose, chitosan, and pectin] on digestion of lipids

under simulated gastrointestinal conditions. Food & Function, 5(12), 3083-3095.

Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L.-P., Narváez-Cuenca, C.-E., & McClements,

D. J. (2014). Interaction of a Dietary Fiber (Pectin) with Gastrointestinal Components (Bile Salts,

Calcium, and Lipase): A Calorimetry, Electrophoresis, and Turbidity Study. Journal of

Agricultural and Food Chemistry, 62(52), 12620-12630.

Espinal-Ruiz, M., Restrepo-Sánchez, L.-P., Narváez-Cuenca, C.-E., & McClements, D. J. (2016). Impact

of pectin properties on lipid digestion under simulated gastrointestinal conditions: Comparison of

citrus and banana passion fruit (Passiflora tripartita var. mollissima) pectins. Food Hydrocolloids,

52, 329-342.

Fabek, H., Messerschmidt, S., Brulport, V., & Goff, H. D. (2014). The effect of in vitro digestive

processes on the viscosity of dietary fibres and their influence on glucose diffusion. Food

Hydrocolloids, 35, 718-726.

Friedman, M. (2004). Applications of the Ninhydrin Reaction for Analysis of Amino Acids, Peptides, and

Proteins to Agricultural and Biomedical Sciences. Journal of Agricultural and Food Chemistry,

52(3), 385-406.

Gunness, P., Flanagan, B. M., Shelat, K., Gilbert, R. G., & Gidley, M. J. (2012). Kinetic analysis of bile

salt passage across a dialysis membrane in the presence of cereal soluble dietary fibre polymers.

Food Chemistry, 134(4), 2007-2013.

Löfgren, C., Walkenström, P., & Hermansson, A.-M. (2002). Microstructure and Rheological Behavior of

Pure and Mixed Pectin Gels. Biomacromolecules, 3(6), 1144-1153.

Mahrenholz, C. C., Tapia, V., Stigler, R. D., & Volkmer, R. (2010). A study to assess the cross-reactivity

of cellulose membrane-bound peptides with detection systems: an analysis at the amino acid level.

Journal of Peptide Science, 16(6), 297-302.

Marcotte, M., Taherian Hoshahili, A. R., & Ramaswamy, H. S. (2001). Rheological properties of selected

hydrocolloids as a function of concentration and temperature. Food Research International, 34(8),

695-703.

Maxwell, E. G., Belshaw, N. J., Waldron, K. W., & Morris, V. J. (2012). Pectin – An emerging new

bioactive food polysaccharide. Trends in Food Science & Technology, 24(2), 64-73.

Page 222: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 7

213

McClements, D. J. (2000). Comments on viscosity enhancement and depletion flocculation by

polysaccharides. Food Hydrocolloids, 14(2), 173-177.

Mohamed, S. (2014). Functional foods against metabolic syndrome (obesity, diabetes, hypertension and

dyslipidemia) and cardiovasular disease. Trends in Food Science & Technology, 35(2), 114-128.

Mohnen, D. (2008). Pectin structure and biosynthesis. Current Opinion in Plant Biology, 11(3), 266-277.

Moore, D. S. (1985). Amino acid and peptide net charges: A simple calculational procedure. Biochemical

Education, 13(1), 10-11.

Pasquier, B., Armand, M., Guillon, F., Castelain, C., Borel, P., Barry, J.-L., Pleroni, G., & Lairon, D.

(1996). Viscous soluble dietary fibers alter emulsification and lipolysis of triacylglycerols in

duodenal medium in vitro. The Journal of Nutritional Biochemistry, 7(5), 293-302.

Phillips, G. O. (2013). Dietary fibre: A chemical category or a health ingredient? Bioactive Carbohydrates

and Dietary Fibre, 1(1), 3-9.

Ryden, P., MacDougall, A. J., Tibbits, C. W., & Ring, S. G. (2000). Hydration of pectic polysaccharides.

Biopolymers, 54(6), 398-405.

Sato, A. K., Oliveira, P., & Cunha, R. (2008). Rheology of Mixed Pectin Solutions. Food Biophysics,

3(1), 100-109.

Srichamroen, A., & Chavasit, V. (2011). In vitro retardation of glucose diffusion with gum extracted from

malva nut seeds produced in Thailand. Food Chemistry, 127(2), 455-460.

Yapo, B. M. (2011). Pectic substances: From simple pectic polysaccharides to complex pectins—A new

hypothetical model. Carbohydrate Polymers, 86(2), 373-385.

Ye, Z., Arumugam, V., Haugabrooks, E., Williamson, P., & Hendrich, S. (2015). Soluble dietary fiber

(Fibersol-2) decreased hunger and increased satiety hormones in humans when ingested with a

meal. Nutrition Research, 35(5), 393-400.

Zsivanovits, G., MacDougall, A. J., Smith, A. C., & Ring, S. G. (2004). Material properties of

concentrated pectin networks. Carbohydrate Research, 339(7), 1317-1322.

Page 223: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

General discussion

Page 224: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

215

Although it has been previously demonstrated that pectin has several physiological functions

beneficial to human health (Cerda, 1988), studies explaining the mechanisms by which pectin is

capable to exert its physiological functions have not been conducted. The research described in

this thesis was focused on the evaluation of the mechanisms by which pectin is capable to exert

its physiological functions.

The mechanisms evaluated in this thesis included the effect of pectin on i) the activities of the

major digestive enzyme (pancreatic lipase, -amylase, alkaline phosphatase, and protease), ii) the

rate and extent of the digestion process of emulsified lipids, and iii) the rate and extent of the

mass transfer process of some of the most important nutritional compounds (monosaccharides,

amino acids, and lipids). Because pancreatic lipase was strongly inhibited by pectin, a special

emphasis was made on the effect of pectin on the gastrointestinal fate of emulsified lipids. The

structural characteristics of pectin (e.g., molecular weight and methoxylation degree) in relation

to their ability to inhibit the digestion process of emulsified lipids were studied. The effect of the

molecular interactions of pectin with the different gastrointestinal components involved in the

lipid digestion process (e.g., pancreatic lipase, bile salts, CaCl2, and NaCl) was evaluated as well.

In this chapter, we discuss the relevance of the mechanisms involved on the physiological

functions of pectin including the effect of pectin on i) the digestive enzyme activities, ii) the

gastrointestinal fate of emulsified lipids, and iii) the mass transfer kinetics of nutritional

compounds. We also discuss some important aspects of the experimental models used to evaluate

the mechanisms by which pectin is capable to exert its physiological functions including i) the

use of artificial chromogenic substrates as a model to evaluate the in vitro activity of the major

digestive enzymes, ii) the advantages and disadvantages of using emulsions as an experimental

model, iii) the advantages and disadvantages of the simulated in vitro digestion model used to

evaluate the effect of pectin on the gastrointestinal fate of emulsified lipids, and iv) the use of a

Franz diffusion cell as an experimental model to evaluate the effect of pectin on the mass transfer

kinetics of some of the most important nutritional compounds.

Page 225: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

216

8.1. Mechanisms involved on the physiological functions of pectin

8.1.1. Effect of pectin on the digestive enzyme activities

In chapter 3, we evaluated the effect of the methoxylation degree (MD) of pectin on the activity

of the digestive enzymes. Alkaline hydrolysis of high methoxylated pectin was performed to

obtain pectins with different MD. The MD of the pectins obtained after alkaline hydrolysis

ranged from 7.1% to 87.4% (mol/mol) (Table 3.1). This allowed us to evaluate the effect of both

low methoxylated (LMP) and high methoxylated (HMP) pectins on the activities of the digestive

enzymes. We found that increasing the concentration of pectin decreased the activity of all of the

tested digestive enzymes by means of a non-competitive inhibition mechanism (Figures 3.2 and

3.3). We also found that HMP was more efficient in inhibiting the activity of all of the tested

digestive enzymes than LMP. Interestingly, each enzyme was inhibited with different efficiencies

(Figure 3.5). Pancreatic lipase was the most likely to be inhibited by pectin, followed by -

amylase, alkaline phosphatase, and protease. Therefore, because pectins, especially HMP,

exhibited a high capacity to inhibit the activity of pancreatic lipase, it was expected that the

digestion process of lipids was also affected by the addition of pectin. Therefore, the following

research was focused in evaluating the effect of pectin on the gastrointestinal fate of lipids. A

simulated in vitro digestion model was used to evaluate the effect of pectin on the gastrointestinal

fate of lipids by using a corn oil-in-water emulsion as the experimental model.

8.1.2. Effect of pectin on the gastrointestinal fate of emulsified lipids

The effect of pectin on the gastrointestinal fate of emulsified lipids was evaluated by using a

simulated in vitro digestion model designed to mimic the oral, gastric, and small intestine phases

of the human gastrointestinal tract (GIT). In chapter 4, the influence on the dietary fiber type

(chitosan, methyl cellulose, and pectin at different concentrations) on the gastrointestinal fate of

emulsified lipids was examined. The aim of that chapter was to compare the effect of pectin on

the rate and extent of the digestion process of emulsified lipids with other sources of dietary fiber

Page 226: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

217

(methyl cellulose and chitosan) with different structural characteristics, especially, the electrical

charge. We found that both the rate and extent (principally the extent of the lipid digestion rather

than rate) of the lipid digestion process were inhibited with increasing chitosan, methyl cellulose,

and pectin concentrations (Figures 4.7 and 4.8). However, the magnitude of the inhibition was

different with each dietary fiber: methyl cellulose had the highest capacity to inhibit the lipid

digestion process, followed by pectin and chitosan. We suggested that the physicochemical

mechanisms that may account for the observed influence of dietary fiber on the rate and extent of

the lipid digestion process are i) the modification of the viscosity of the GIT fluids, ii) the

flocculation of the emulsified lipids and iii) the electrostatic interactions between dietary fiber

and emulsified lipids.

As observed, several mechanisms by which dietary fiber, especially pectin, is able to inhibit the

lipid digestion process were proposed. However, a relationship between chemical structure and

functional properties of dietary fiber could not be established in that chapter because the majority

of the structural characteristics of the different sources of dietary fiber used in chapter 4 were

unknown. Therefore, we focused on studying the effect of the structural characteristics of pectin

(e.g., molecular weight and methoxylation degree) on the rate and extent of digestion process of

emulsified lipids. In chapter 5, the effect of pectin properties (e.g., methoxylation degree and

molecular weight) on the gastrointestinal fate of emulsified lipids was evaluated. Medium

methoxylated pectin (MMP) was isolated from banana passion fruit (Passiflora tripartita var.

mollisima) and then, the impact of MMP on the rate and extent of the digestion process of

emulsified lipids was compared to that of commercial LMP and HMP. We found that all three

pectins promoted flocculation of the lipid droplets, which can be attributed to a depletion

flocculation mechanism. Calculations of the inter-droplet pair potential due to depletion attraction

of lipid droplets (Figure 5.10) in emulsions containing LMP, MMP, and HMP revealed that the

strength of the depletion attraction increases with the simultaneous increase of both molecular

weight and methoxylation degree in the following order: HMP>MMP>LMP. Therefore, it can be

suggested that HMP exhibited the highest ability to inhibit the lipid digestion process probably

due to its higher capacity to flocculate the lipid droplets, followed MMP and LMP (Figure 5.9).

Page 227: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

218

Although flocculation was identified as a mechanism controlling the rate and extent of the lipid

digestion process, the results obtained so far did not give any information concerning the

physicochemical nature of the interactions occurring between pectin and the various GIT

components associated with the digestion process of emulsified lipids. Therefore, in the next

section we provide a discussion on the effect of the molecular interactions occurring between

pectin and the various GIT components and their relationship with the rate and extent of the

digestion process of emulsified lipids.

8.1.3. Molecular interactions between pectin and the GIT components

We showed that pectin is able to reduce the rate and extent of the digestion process by interacting

with emulsified lipids. However, the interaction of pectin with the different components of the

GIT involved in the digestion process of emulsified lipids (e.g., pancreatic lipase, bile salts, and

electrolytes) may also contribute to the overall inhibitory effect. In chapter 6, isothermal titration

calorimetry (ITC) was used to study the interactions of pectin with the major GIT components

involved in the digestion process of emulsified lipids (e.g., pancreatic lipase, bile salts, CaCl2,

and NaCl). In addition, microelectrophoresis, turbidity, and microstructural observations were

used to provide additional information about the nature of the interactions between pectin and the

aforementioned GIT components.

We found that the addition of NaCl (Figure 6.5a), bile salts (Figure 6.5c), and pancreatic lipase

(Figure 6.5d) to pectin solutions caused a small decrease in the magnitude of the zeta-potential,

as compared to the large decrease caused upon addition of CaCl2 (Figure 6.5b). Consequently,

the addition of CaCl2 to pectin solutions caused the highest increase of the optical turbidity

because of the formation of large aggregates (Figure 6.8), followed by bile salts, pancreatic

lipase, and NaCl. This observation allowed us to suggest that both CaCl2 and bile salts play a

determinant role in the overall interaction of pectin with the GIT components. In particular,

CaCl2 and bile salts promoted the formation of pectin gels (Figure 6.8) that can lead to the

flocculation of lipid droplets and decrease the ability of emulsified lipids to be reached by

pancreatic lipase, thereby inhibiting the digestion process.

Page 228: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

219

The formation of gel-like structures resulting from the interaction between pectin and the GIT

components can lead to the flocculation of lipid droplets and thereby inhibiting the digestion

process of emulsified lipids. However, the increase of the viscosity upon addition of pectin may

also affect the mobility of both the GIT components and the nutritional compounds to be

digested. Therefore, the effect of pectin on the rate and extent of the mass transfer process of the

most important nutritional compounds (monosaccharides, amino acids, and lipids) was evaluated

in chapter 7 as another possible mechanism controlling the GIT processes.

8.1.4. Effect of pectin on the mass transfer of nutrients

We found that the mass transfer process of electrically neutral nutrients (monosaccharides,

glycine, and tyrosine, all neutral at pH 7.0) was governed by the viscosity of the pectin solutions,

whereas the mass transfer process of electrically charged nutrients (aspartic-negatively charged at

pH 7.0 acid and lysine-positively charged at pH 7.0) was governed by both viscosity and the

electrical interactions of these nutrients with pectin. In addition, flocculation was proposed to be

the most important mechanism controlling the mass transfer process of emulsified lipids. Pectins,

especially HMP, were able to inhibit both the rate and extent of the mass transfer process of

electrically neutral nutrients (monosaccharides, glycine, and tyrosine). The apparent viscosity

profiles of LMP and HMP (Figure 7.5) revealed that the viscosity of HMP was significantly

higher than the one of LMP. Therefore, the retardation of the rate and extent of the mass transfer

process of electrically neutral nutrients was higher with HMP than LMP probably due to the

higher viscosity of HMP as compared to that of LMP (Figures 7.2, 7.3a, and 7.3d). Interestingly,

the mass transfer process of electrically charged nutrients (aspartic acid and lysine) was governed

by both the viscosity of the pectin solutions and the electrical interactions of these nutrients with

pectins. The repulsive interactions between aspartic acid and LMP prevented aspartic acid

molecules to be trapped in the structural network of LMP, and therefore, the extent of the mass

transfer process of aspartic acid was not significantly hindered by LMP. Conversely, the

attractive interactions between lysine and LMP promoted significantly its entrapment in the

structural network of LMP, and therefore, both the rate and extent of the mass transfer process of

lysine was significantly hindered upon addition of LMP.

Page 229: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

220

Finally, we suggested that the mechanism by which pectin is able to retard the rate and extent of

the mass transfer process of the emulsified lipids might be related to the high capacity of HMP to

flocculate the lipid droplets of the emulsion as compared to LMP (chapter 5). Figure 8.1 shows

schematically that several mechanisms were proposed in this thesis for controlling the

physiological properties of pectin, including the inhibition of the digestive enzymatic activities

and the modulation of both the mass transfer of nutrients and the digestion of emulsified lipids.

Structural characteristics of pectin (methoxylation degree and molecular weight) significantly

influenced the ability of pectin to interact with nutrients as well as the GIT components.

However, further studies are required to discriminate the relative contribution of each mechanism

to the overall effect, as well as the relative contribution of the individual structural characteristics

of pectin, including the methoxylation degree and molecular weight.

Figure 8.1. Schematic representation of the mechanisms proposed in this thesis for controlling the

physiological properties of pectin. i) Non-competitive inhibition of digestive enzymes, ii) flocculation of

lipid droplets, and iii) hindered diffusion of enzymatic products to epithelial cells (enterocytes). Red,

substrate; blue, product; green, enzyme; and grey-red stick, pectin.

i) Non-competitive inhibition

ii) Flocculation

iii) Hindered diffusion

Ephitelial cells (enterocytes)

Page 230: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

221

8.2. Using artificial chromogenic substrates for evaluating the

activity of digestive enzymes

Pancreatic lipase, -amylase, and protease are the main enzymes responsible for the digestion of

lipids, carbohydrates, and proteins, respectively (Campbell, 2015). These enzymes were studied

because of their biological relevance on the digestives processes. Alkaline phosphatase was also

studied because it participates in the dephosphorylation of nutrients, which is a necessary process

for their further digestion (Vaishnava & Hooper, 2007). In chapter 2, the optimization of the

reaction conditions affecting the activity of these digestive enzymes by using artificial

chromogenic substrates was carried out. The reaction conditions obtained in that chapter allowed

obtaining rapid, reproducible, and reliable measurements of the activities of each of the tested

digestive enzymes. The best reactions conditions obtained in chapter 2 for the measurement of

the aforementioned digestive enzyme activities (conditions summarized in Table 2.1) were then

used in chapter 3 to evaluate the inhibitory effect of pectin upon the activity of these enzymes by

using artificial chromogenic substrates.

Enzymes are often highly specific to its native substrate (Qian, 2008). However, some enzymes

are able to exhibit activity toward several substrates, showing a characteristic known as enzyme-

substrate promiscuity (Hult & Berglund, 2007). In particular, pancreatic lipase and proteases have

been reported to exhibit enzyme-substrate promiscuity by accepting a wide variety of substrates

(decrease in specificity) with which they have higher affinities (increase in selectivity) as

compared to native substrates (Khersonsky, Roodveldt, & Tawfik, 2006). These substrates are

e.g. chromogenic artificial substrates and the high selectivity of digestive enzymes by these

substrates allows obtaining high enzymatic activities in short period of time. In the one hand, the

high selectivity of pancreatic lipase towards p-nitrophenyl ester substrates (e.g., p-nitrophenyl

palmitate) has been reported to be related with the lack of chirality of the alcohol moiety of the

ester substrate (Kapoor & Gupta, 2012). In the other hand, the high selectivity of protease

towards p-nitrophenyl acetate has been reported to be related to the lack of the peptide bond of

the substrate (which makes p-nitrophenyl acetate a suitable substrate for esterases in general) and

Page 231: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

222

because the structural similarity of the substrate with aromatic amino acids (Gossrau, Lojda,

Smith, & Sinha, 1984). The high selectivity of alkaline phosphatase and -amylase towards p-

nitrophenyl phosphate and 2-chloro-p-nitrophenyl--D-maltotrioside, respectively, has been

related to the structural similarity of those substrates with their analogous native substrates

(Copley, 2003; Hult & Berglund, 2007; Khersonsky, Roodveldt, & Tawfik, 2006). Therefore, the

advantages of using artificial chromogenic substrates are the velocity, reproducibility, and

reliability of the results that can be obtained as compared to those obtained when native

substrates are used. Nevertheless, it is important to stress that the use of native substrates is

preferred because of the biological relevance of the results that can be obtained, but native

substrates are often difficult to obtain and the heterogeneity of their chemical structures (e.g.,

chemical composition, molecular weight, and three-dimensional conformation) often prevent

obtaining reproducible results (Jedrzejas, 2000; MacGregor, Janeček, & Svensson, 2001).

One of the disadvantages of using artificial chromogenic substrates is that it is necessary to adjust

the measurement parameters because the experimental conditions reported in literature for native

substrates often differ to those required for artificial ones. For example, the experimental

conditions adjusted for pancreatic lipase in chapter 2 were used in chapter 3 to evaluate the

effect of pectin on the activity of this enzyme in model solutions. However, these experimental

conditions were not suitable to be used to evaluate the effect of pectin on the gastrointestinal fate

of emulsified lipids in chapters 4, 5, and 6 because of the differences on both the chemical nature

of the substrates used in each system and the environmental conditions of the enzymatic

reactions. It has been established that pancreatic lipase has low specificity towards soluble

substrates (e.g., p-nitrophenyl palmitate) but it exhibits a high specificity when the substrate (e.g.,

triacylglyceride) is able to form an interface between the aqueous and the oil phases of an

emulsion (Reis, Holmberg, Miller, Leser, Raab, & Watzke, 2009; Reis, Holmberg, Watzke,

Leser, & Miller, 2009). Therefore, the differences among the chemical nature of the substrate

(e.g., solubility, polarity, hydrophobicity, and three-dimensional conformation) as well as the

environmental factors (e.g., pH, temperature, ionic strength, and cofactors) affecting the rate of

the enzymatic reaction make necessary to adjust the experimental conditions for each reaction

depending on the nature of the system that will be used.

Page 232: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

223

8.3. Emulsified lipids as an experimental model: advantages and

disadvantages

The selection of emulsions as an experimental model representing the wide variety of lipids that

can be consumed in the diet was a critical factor to be considered for obtaining reproducible and

reliable results of the gastrointestinal fate of emulsified lipids by using a simulated in vitro

digestion model. In the oral phase of the in vitro digestion model, dietary lipids (fats and oils) are

mixed with saliva in the mouth. The shear effect of the mastication process (which was simulated

by using mechanical agitation) and the presence of mucin in saliva induced the formation of lipid

droplets which are dispersed in the aqueous phase, leading to the formation of an oil-in-water

emulsion (Vingerhoeds, Blijdenstein, Zoet, & van Aken, 2005). Depending on the chemical

nature and the concentration of both the oil phase (e.g., corn oil) and the surfactant (e.g., Tween

80) as well as the residence time that the emulsion spends in the oral phase, emulsified lipids are

susceptible of suffering flocculation and coalescence induced by mucin (Vingerhoeds,

Blijdenstein, Zoet, & van Aken, 2005). Therefore, the oral phase modifies the structural

characteristics of emulsified lipids and defines their subsequent behavior in the digestion process.

Regardless of the origin, chemical nature, and physical state of the lipids consumed in the diet,

lipids will be structured as an oil-in-water emulsion in the oral phase (Singh, Ye, & Horne, 2009).

This behavior allowed us to use the emulsified lipids as a suitable model for studying the

gastrointestinal fate of lipids when they are subjected to the digestion process. Therefore,

emulsified lipids provide a fundamental model which allows understanding the relationship

between the initial physicochemical properties of the emulsion (e.g., composition, concentration,

particle size, electrical charge, and interfacial characteristics) and its behavior toward the

digestion process (e.g., stability, appearance, rheology, and spatial distribution). Among the

advantages of using emulsified lipids as an experimental model are that they represent the

structure of the wide variety of lipids that can be consumed in the diet, they are easy to prepare

being possible to replicate their initial physicochemical characteristics, they usually have a high

long-term stability to gravitational separation and aggregation, and they are susceptible to present

Page 233: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

224

structural changes depending on the environmental conditions of the digestive process (e.g., pH,

ionic strength, mechanical forces, and the presence of dietary fiber such as pectin). Nevertheless,

some disadvantages of using emulsified lipids as an experimental model are that there are a

limited number of food-grade emulsifiers that can be used to prepare them (e.g., milk proteins,

phospholipids, monoacylglycerides, and diacylglycerides) and that expensive equipment for their

homogenization and particle size reduction is required (McClements, 2010).

8.4. Simulated in vitro digestion of emulsified lipids: Advantages and

disadvantages

The effect of pectin on the gastrointestinal fate of emulsified lipids (chapters 4 and 5) was

evaluated by using a simulated in vitro digestion model designed to mimic the oral, gastric, and

small intestine phases of the human gastrointestinal tract (GIT). These phases involved mixing of

the sample (a corn oil-in-water emulsion mixed with different pectin samples and other sources

of dietary fibers at different concentrations) with simulated digestive fluids of variable

composition depending on the simulated phase. The temperature of the in vitro digestion model

was maintained at 37 °C to mimic human corporal temperature. An accurate simulation of the

human GIT should involve mimicking the exact composition and dynamics of the GIT fluids.

However, this is usually complicated to be implemented because the model would be difficult to

design and operate. Nevertheless, the utilization of the key components of the GIT fluids which

are able to impact the gastrointestinal fate of emulsified lipids as well as the simulation of the

GIT dynamics (e.g., mastication and both gastric and intestinal peristalsis were simulated by

orbital shaking) were taken into account for understanding the mechanisms involved in the lipid

digestion process (Hur, Lim, Decker, & McClements, 2011). Because emulsified lipids are fully

digested in the small intestinal phase (Li, Hu, & McClements, 2011; Singh, Ye, & Horne, 2009),

the simulation of the colonic phase was not performed. Furthermore, the colonic phase is usually

difficult to mimic because simulation of the anaerobic conditions requires special instrumentation

and the wide variety of colonic bacteria representing the colonic microbiota are often difficult to

preserve (Hur, Lim, Decker, & McClements, 2011).

Page 234: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

225

The advantage of using a simulated in vitro digestion model is that it provides a useful alternative

to in vivo models (e.g., cell culture, animal, and human models) by rapidly screening the effect of

different sources of dietary fibers in the gastrointestinal fate of emulsified lipids. Although in

vivo models often provide accurate results, they are time consuming, poorly reproducible,

expensive, and several ethical considerations should be taken into account when using animal

models (Li, Kim, Park, & McClements, 2012). Results obtained with in vitro models are often

different to those obtained with in vivo models because of the high complexity of the dynamics,

structure, and composition of the in vivo models (Koven, Henderson, & Sargent, 1997).

Consequently, a simulated in vitro digestion model must be designed to find an appropriate

balance between the accuracy and biological relevance of the results with the cost and the

affordable operability of the model (Hur, Lim, Decker, & McClements, 2011; McClements & Li,

2010).

The disadvantage of using a simulated in vitro digestion model is that the results are highly

dependent of the experimental conditions (McClements & Li, 2010). For example, the selection

of the type and concentration of enzymes (e.g., selecting between pepsin, trypsin, or

chymotrypsin), ion metal cofactors (e.g., selecting between CaCl2, MgCl2, ZnCl2, or NaCl), and

surface-active compounds (e.g., selecting between phospholipids or bile salts) is a key factor

defining the efficiency of the digestion process of emulsified lipids as well as the reliability and

the biological relevance of the results (Li, Hu, & McClements, 2011). Minor changes in these

parameters can significantly affect the final results and also limit their comparability with those

obtained with other experimental models (Bonnaire, Sandra, Helgason, Decker, Weiss, &

McClements, 2008).

For this reason, the simulated in vitro digestion model used in this thesis was selected because

this model has been widely studied (Hur, Lim, Decker, & McClements, 2011; McClements & Li,

2010; Minekus, Alminger, Alvito, Ballance, Bohn, Bourlieu, et al., 2014), validated (Bonnaire,

Sandra, Helgason, Decker, Weiss, & McClements, 2008; Li, Hu, & McClements, 2011), and the

results obtained with this model have been shown to correlate with in vivo models (Li, Kim, Park,

& McClements, 2012).

Page 235: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

226

8.5. Franz diffusion cell as an in vitro model for evaluating the mass

transfer of nutrients

In chapter 7, the effect of both low (LMP) and high (HMP) methoxylated pectins on the rate and

extent of the mass transfer process of the primary nutritional compounds consumed in the diet or

produced after digestion such as monosaccharides (e.g., glucose, galactose, fructose, and ribose),

amino acids (e.g., glycine, aspartic acid, lysine, and tyrosine), and lipids (represented by a corn

oil-in-water emulsion) was evaluated. Both viscosity of the pectin solutions and the electrical

interactions between pectin and the nutritional compounds were proposed to be the main factors

determining the rate and extent of the mass transfer process. Franz diffusion cells have been used

successfully for several years in the pharmaceutical industry to evaluate transdermal drug

administration (Ng, Rouse, Sanderson, Meidan, & Eccleston, 2010). No studies had been

reported concerning the use of the Franz diffusion cells for studying the mass transfer kinetics of

nutritional compounds. Results presented in chapter 7 are among those few studies related to the

effect of pectin on the mass transfer kinetics of nutritional compounds (monosaccharides, amino

acids, and emulsified lipids) by using the Franz diffusion cell model. There are several

experimental parameters to be considered when using a Franz diffusion cell including i) the cell

design, ii) the agitation of the receptor compartment, and iii) the physicochemical nature of the

barrier membrane. These parameters are related to the reproducibility, reliability, and biological

relevance of the results.

i) The Franz diffusion cell is an experimental setup widely used for studying drug permeation

through several tissues. This has led to a great demand of these devices as well as the emergence

of a wide variety of designs depending on the requirements of each experiment. However, the

design of the Franz diffusion cell was standardized by the U.S. Food and Drug Administration

(FDA) to reduce the variability between the measurements, to reduce the cost of the experimental

setup, and to make the results comparable to other studies (Skelly, Shah, Maibach, Guy, Wester,

Flynn, et al., 1987). It has been established that the minimum volume of the donor and receptor

compartments of the Franz diffusion cell should be 1 and 10 mL, respectively, to prevent

Page 236: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

227

significant evaporation in long-period experiments (Bonferoni, Rossi, Ferrari, & Caramella,

1999). In our experimental setup, the volume of the donor and receptor compartments were 3 and

17 mL, respectively, allowing us to ensure that our results (e.g., effective diffusion coefficients)

are reliable and they may be comparable to other experiments. For example, we found that the

effective diffusion coefficients of glucose through the Franz diffusion cell were 1.44x10-5

and

0.48x10-5

cm2 s

-1 for control (containing no pectin) and 2% (w/w) HMP, respectively, at a

temperature of 37 °C, whereas the effective diffusion coefficients of glucose through a dialysis

membrane were reported to be 1.21x10-5

and 0.27x10-5

cm2 s

-1 for control (containing no malva

nut gum) and 1% (w/w) malva nut gum, respectively, at the same temperature (Srichamroen &

Chavasit, 2011). Therefore, our results concerning the effect of pectin on the mass transfer

process of nutritional compounds through a Franz diffusion cell can be compared to those

obtained with other experimental models.

ii) Agitation of the receptor compartment is a critical parameter for the maintenance of both

uniform nutrient distribution through the cell as well as temperature equilibrium (Ng, Rouse,

Sanderson, Meidan, & Eccleston, 2010). Nevertheless, agitation introduces a convective

parameter that also has an effect on the overall mass transfer process. It is important to consider

that the mass transfer process in the unstirred water layer of the intestinal lumen (where the

absorption process of nutrients takes place) is strictly controlled by diffusion, while convection

can be considered as negligible (Smithson, Millar, Jacobs, & Gray, 1981; Thomson & Dietschy,

1984). Therefore, agitation may have caused an overestimation on the effective diffusion

coefficients (related to the rate at which the mass transfer process occurs) because convective

effects often accelerates the mass transfer process of nutrients (Cussler, 2009). However,

agitation does not necessarily affect the extent of the mass transfer process of nutrients, it only

affects the time at which the mass transfer process can be completed (Cussler, 2009).

iii) Another key parameter affecting the mass transfer process of nutrients relates to the barrier

membrane used to separate the donor and the receptor compartment of the Franz diffusion cell.

Usually, the membranes used to simulate the absorption of nutrients must be biological

membranes capable to mimic both passive and active diffusion mechanisms occurring in in vivo

Page 237: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

228

systems (Schulthess, Lipka, Compassi, Boffelli, Weber, Paltauf, et al., 1994). The most common

biological membranes used to simulate the intestinal absorption process of nutrients are rat

intestinal microvillus membrane and pig small intestine brush border (Deglaire & Moughan,

2012; Gill, Smith, Wissler, & Kunz, 1989). However, biological membranes are often difficult to

obtain and preserve, in addition to the reproducibility of the results is sometimes unacceptable

(Schulthess, et al., 1994; Smithson, Millar, Jacobs, & Gray, 1981). Synthetic membranes (e.g.,

cellulose membrane) allow the permeation process to occur by simple diffusion (in the absence of

membrane proteins) rather than passive and active diffusion, which involves the presence of

membrane proteins (Lack & Weiner, 1961). Usually, simple diffusion is faster than both passive

and active diffusion because there are not proteins restricting the mass transfer process of

nutritional compounds through the membrane (Ng, Rouse, Sanderson, Meidan, & Eccleston,

2010). Therefore, the presence of a cellulose membrane in our experimental setup may have also

contributed to the overestimation on the effective diffusion coefficients of the nutritional

compounds because of the lack of selectivity of the cellulose membrane.

8.6. Future prospects

The effect of pectin on the activity of digestive enzymes was conducted by using artificial

chromogenic substrates (chapter 3). These substrates are structurally similar to their analogous

native substrates and they allowed obtaining fast, reliable, and reproducible results. However,

artificial substrates should only be used for comparison purposes because the enzymatic activities

obtained with these substrates are not necessarily comparable to those obtained by using native

substrates. Therefore, the evaluation of the inhibitory effect of pectin on the activity of the

digestive enzymes by using native substrates might lead to obtain results with greater biological

relevance as compared to those obtained with artificial substrates.

Comparisons of the functional properties of different pectins must be done carefully because of

the diversity of pectins that can be obtained from different plant sources. In further studies,

detailed information on the chemical structure of pectins must be obtained. For example, it has

been pointed out that structural features of pectins such as the acetylation degree, branching

Page 238: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

229

degree, contents of rhamnose, arabinose and galactose, and the water holding capacity are closely

related to their functional properties, besides methoxylation degree and molecular weight (James,

1986). In addition, the evaluation of the functional properties of several pectin samples with

different structural features is required to identify the relative contribution of the individual

structural characteristics to the overall effect. Furthermore, those physicochemical properties

must be independent to each other to discriminate the relative effect produced by each parameter

individually.

Simulated in vitro gastrointestinal digestion of emulsified lipids is widely employed in many

fields of the food science and nutrition research. Several in vitro digestion models have been

proposed depending on the purpose of the research and the availability of resources. Therefore,

the possibility to compare results with other investigations is often difficult because of the

variability of experimental conditions used in each model. For example, a large variety of

digestive enzymes from different sources have been employed. In addition, differences in pH,

ionic strength, and digestion times may also affect the results of the digestion process. Therefore,

we propose to use the recently published standardized in vitro digestion model (Infogest

protocol) coming from an international consensus that it is suitable for application in several

experimental situations and allowing to obtain more comparable results in future research

(Minekus, et al., 2014). In addition, the sophistication level of the simulated in vitro digestion

model can significantly affect the reliability and biological relevance of the results (Hur, Lim,

Decker, & McClements, 2011; McClements & Li, 2010). Analytical instruments designed to

simulate the full digestion process are currently available, e.g., the TIM (Intestinal Tract Model)

lipid absorption system (TNO Innovation for Life, Zeist, The Netherlands) (Dickinson, Abu

Rmaileh, Ashworth, Barker, Burke, Patterson, et al., 2012). The TIM model has been developed

to simulate the full dynamic and physicochemical complexity of the human GIT. The TIM model

allows performing studies concerning the digestibility and bioaccessibility of several food

components including lipids, proteins, minerals, and vitamins (Dickinson, et al., 2012).

Therefore, analytical instruments for simulating the dynamics of the GIT would lead to obtain

reliable results with significant biological relevance.

Page 239: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

230

The rate and extent of the in vitro digestion process of emulsified lipids, as well as the

bioaccessibility and bioavailability of emulsified lipids must be correlated with in vivo

experiments by using cell or animal models. For example, the human intestinal Caco-2 cell line

has been extensively used as a model of the intestinal barrier to simulate the absorptive properties

of the intestinal mucosa and to evaluate both bioaccessibility and bioavailability of food

components (Sambuy, De Angelis, Ranaldi, Scarino, Stammati, & Zucco, 2005). In addition,

animal models have been used for determining several food components digestibility,

bioaccesibility, and bioavailability (Deglaire & Moughan, 2012). For example, Wistar rat (Rattus

norvegicus) is usually considered as a suitable animal model for digestion experiments because

its upper GIT is anatomically and physiologically similar to that of the humans (Gill, Smith,

Wissler, & Kunz, 1989). In addition, Wistar rats are usually easy to rise and relatively

inexpensive to maintain, and the results that can be obtained are biologically relevant. However,

it is important to consider the large variability of the results that can be obtained between

experimental units depending on gender, age, feeding, and environmental conditions, besides the

ethical considerations that must be taken into account when using animals as experimental

models (Deglaire & Moughan, 2012; Gill, Smith, Wissler, & Kunz, 1989). As aforementioned,

the correlation of the results obtained in this study by using in vitro models must be correlated

with results obtained by in vivo models (e.g., cell lines or Wistar rats) to validate the biological

relevance of the results obtained here and to enhance the possibility to extrapolate these results to

more complex models (e.g., clinical trials with humans).

Finally, the results obtained in this thesis would lead to the rational design of pectin-based

functional and nutraceutical foods targeted for populations with different susceptibilities of

suffering several cardiovascular diseases (e.g., adult and elderly populations). It is well known

that pectins are able to significantly impact the sensory and physicochemical properties of pectin-

enriched foodstuffs; therefore, the rational design of pectin-based functional foods would lead to

obtain a functional food which retains its desirable sensory and textural properties.

Page 240: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

231

References

Bonferoni, M. C., Rossi, S., Ferrari, F., & Caramella, C. (1999). A Modified Franz Diffusion Cell for

Simultaneous Assessment of Drug Release and Washability of Mucoadhesive Gels.

Pharmaceutical Development and Technology, 4(1), 45-53.

Bonnaire, L., Sandra, S., Helgason, T., Decker, E. A., Weiss, J., & McClements, D. J. (2008). Influence of

Lipid Physical State on the in Vitro Digestibility of Emulsified Lipids. Journal of Agricultural

and Food Chemistry, 56(10), 3791-3797.

Campbell, I. (2015). Digestion and absorption. Anaesthesia & Intensive Care Medicine, 16(1), 35-36.

Cerda, J. J. (1988). The role of grapefruit pectin in health and disease. Transactions of the American

Clinical and Climatological Association, 99, 203-213.

Copley, S. D. (2003). Enzymes with extra talents: moonlighting functions and catalytic promiscuity.

Current Opinion in Chemical Biology, 7(2), 265-272.

Cussler, E. L. (2009). Diffusion: mass transfer in fluid systems: Cambridge university press.

Deglaire, A., & Moughan, P. J. (2012). Animal models for determining amino acid digestibility in humans

– a review. British Journal of Nutrition, 108(SupplementS2), S273-S281.

Dickinson, P. A., Abu Rmaileh, R., Ashworth, L., Barker, R. A., Burke, W. M., Patterson, C. M.,

Stainforth, N., & Yasin, M. (2012). An Investigation into the Utility of a Multi-compartmental,

Dynamic, System of the Upper Gastrointestinal Tract to Support Formulation Development and

Establish Bioequivalence of Poorly Soluble Drugs. The AAPS Journal, 14(2), 196-205.

Gill, T., Smith, G., Wissler, R., & Kunz, H. (1989). The rat as an experimental animal. Science,

245(4915), 269-276.

Gossrau, R., Lojda, Z., Smith, R., & Sinha, P. (1984). Recent Advances in Protease Research using

Synthetic Substrates. In W. Hörl & A. Heidland (Eds.), Proteases, vol. 167 (pp. 191-207):

Springer US.

Hult, K., & Berglund, P. (2007). Enzyme promiscuity: mechanism and applications. Trends in

Biotechnology, 25(5), 231-238.

Hur, S. J., Lim, B. O., Decker, E. A., & McClements, D. J. (2011). In vitro human digestion models for

food applications. Food Chemistry, 125(1), 1-12.

James, N. B. (1986). An Introduction to Pectins: Structure and Properties. In Chemistry and Function of

Pectins, vol. 310 (pp. 2-12): American Chemical Society.

Jedrzejas, M. J. (2000). Structure, function, and evolution of phosphoglycerate mutases: comparison with

fructose-2,6-bisphosphatase, acid phosphatase, and alkaline phosphatase. Progress in Biophysics

and Molecular Biology, 73(2–4), 263-287.

Kapoor, M., & Gupta, M. N. (2012). Lipase promiscuity and its biochemical applications. Process

Biochemistry, 47(4), 555-569.

Khersonsky, O., Roodveldt, C., & Tawfik, D. S. (2006). Enzyme promiscuity: evolutionary and

mechanistic aspects. Current Opinion in Chemical Biology, 10(5), 498-508.

Koven, W. M., Henderson, R. J., & Sargent, J. R. (1997). Lipid digestion in turbot (Scophthalmus

maximus): in-vivo and in-vitro studies of the lipolytic activity in various segments of the digestive

tract. Aquaculture, 151(1–4), 155-171.

Lack, L., & Weiner, I. M. (1961). In vitro absorption of bile salts by small intestine of rats and guinea

pigs. American Journal of Physiology -- Legacy Content, 200(2), 313-317.

Li, Y., Hu, M., & McClements, D. J. (2011). Factors affecting lipase digestibility of emulsified lipids

using an in vitro digestion model: Proposal for a standardised pH-stat method. Food Chemistry,

126(2), 498-505.

Page 241: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Chapter 8

232

Li, Y., Kim, J., Park, Y., & McClements, D. J. (2012). Modulation of lipid digestibility using structured

emulsion-based delivery systems: Comparison of in vivo and in vitro measurements. Food &

Function, 3(5), 528-536.

MacGregor, E. A., Janeček, Š., & Svensson, B. (2001). Relationship of sequence and structure to

specificity in the α-amylase family of enzymes. Biochimica et Biophysica Acta (BBA) - Protein

Structure and Molecular Enzymology, 1546(1), 1-20.

McClements, D. J. (2010). Emulsion Design to Improve the Delivery of Functional Lipophilic

Components. Annual Review of Food Science and Technology, 1(1), 241-269.

McClements, D. J., & Li, Y. (2010). Review of in vitro digestion models for rapid screening of emulsion-

based systems. Food & Function, 1(1), 32-59.

Minekus, M., Alminger, M., Alvito, P., Ballance, S., Bohn, T., Bourlieu, C., Carriere, F., Boutrou, R.,

Corredig, M., Dupont, D., Dufour, C., Egger, L., Golding, M., Karakaya, S., Kirkhus, B., Le

Feunteun, S., Lesmes, U., Macierzanka, A., Mackie, A., Marze, S., McClements, D. J., Menard,

O., Recio, I., Santos, C. N., Singh, R. P., Vegarud, G. E., Wickham, M. S. J., Weitschies, W., &

Brodkorb, A. (2014). A standardised static in vitro digestion method suitable for food - an

international consensus. Food & Function, 5(6), 1113-1124.

Ng, S.-F., Rouse, J. J., Sanderson, F. D., Meidan, V., & Eccleston, G. M. (2010). Validation of a Static

Franz Diffusion Cell System for In Vitro Permeation Studies. AAPS PharmSciTech, 11(3), 1432-

1441.

Qian, H. (2008). Cooperativity and Specificity in Enzyme Kinetics: A Single-Molecule Time-Based

Perspective. Biophysical Journal, 95(1), 10-17.

Reis, P., Holmberg, K., Miller, R., Leser, M. E., Raab, T., & Watzke, H. J. (2009). Lipase reaction at

interfaces as self-limiting processes. Comptes Rendus Chimie, 12(1–2), 163-170.

Reis, P., Holmberg, K., Watzke, H., Leser, M. E., & Miller, R. (2009). Lipases at interfaces: A review.

Advances in Colloid and Interface Science, 147–148, 237-250.

Sambuy, Y., De Angelis, I., Ranaldi, G., Scarino, M. L., Stammati, A., & Zucco, F. (2005). The Caco-2

cell line as a model of the intestinal barrier: influence of cell and culture-related factors on Caco-2

cell functional characteristics. Cell Biology and Toxicology, 21(1), 1-26.

Schulthess, G., Lipka, G., Compassi, S., Boffelli, D., Weber, F. E., Paltauf, F., & Hauser, H. (1994).

Absorption of Monoacylglycerols by Small Intestinal Brush Border Membrane. Biochemistry,

33(15), 4500-4508.

Singh, H., Ye, A., & Horne, D. (2009). Structuring food emulsions in the gastrointestinal tract to modify

lipid digestion. Progress in Lipid Research, 48(2), 92-100.

Skelly, J., Shah, V., Maibach, H., Guy, R., Wester, R., Flynn, G., & Yacobi, A. (1987). FDA and AAPS

Report of the Workshop on Principles and Practices of In Vitro Percutaneous Penetration Studies:

Relevance to Bioavailability and Bioequivalence. Pharmaceutical Research, 4(3), 265-267.

Smithson, K., Millar, D., Jacobs, L., & Gray, G. (1981). Intestinal diffusion barrier: unstirred water layer

or membrane surface mucous coat? Science, 214(4526), 1241-1244.

Srichamroen, A., & Chavasit, V. (2011). In vitro retardation of glucose diffusion with gum extracted from

malva nut seeds produced in Thailand. Food Chemistry, 127(2), 455-460.

Thomson, A. B. R., & Dietschy, J. M. (1984). The Role of the Unstirred Water Layer in Intestinal

Permeation. In T. Csáky (Ed.), Pharmacology of Intestinal Permeation II, vol. 70 / 2 (pp. 165-

269): Springer Berlin Heidelberg.

Vaishnava, S., & Hooper, L. V. (2007). Alkaline Phosphatase: Keeping the Peace at the Gut Epithelial

Surface. Cell Host & Microbe, 2(6), 365-367.

Vingerhoeds, M. H., Blijdenstein, T. B. J., Zoet, F. D., & van Aken, G. A. (2005). Emulsion flocculation

induced by saliva and mucin. Food Hydrocolloids, 19(5), 915-922.

Page 242: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Concluding remarks

233

Concluding remarks

The results obtained in this thesis showed that pectin is able to exert its physiological functions

by several mechanisms including the inhibition of the activity of the major digestive enzymes

(pancreatic lipase, -amylase, alkaline phosphatase, and protease) and the modulation of the rate

and extent of both the digestion of emulsified lipids (predominantly the extent of the digestion

process of emulsified lipids rather than the digestion rate) and the mass transfer of the most

important nutritional compounds (monosaccharides, amino acids, and emulsified lipids). These

effects may have been due to i) the impact of pectin on the activity of the digestive enzymes

(pectins behave as non-competitive inhibitors of the digestive enzymes), ii) the rheological

properties of the gastrointestinal fluids (pectins inhibited both the rate and extent of the mass

transfer process of nutrients by increasing the viscosity), iii) the interaction of pectin with key

gastrointestinal components [e.g., pancreatic lipase, electrolytes (Na⊕ and Ca2⊕), and bile salts],

and iv) the alteration in the lipid droplet aggregation state (pectin promoted depletion flocculation

of emulsified lipids). In addition, the structural characteristics of pectin (e.g., methoxylation

degree and molecular weight) have a significant effect on the magnitude by which such effects

may occur. However, further studies are required to discriminate the relative contribution of each

mechanism to the overall inhibitory effect as well as the relative contribution of the individual

structural characteristics of pectin, including the methoxylation degree and molecular weight. The

results obtained in this thesis have important implications for the rational design of pectin-based

functional and nutraceutical foods that may modulate lipid digestion within the gastrointestinal

tract. The design of pectin-based functional and nutraceutical foods is focused to give healthier

lipid profiles and thereby promoting health and wellness of the consumers. The inclusion of

pectin in food formulations might be also an effective strategy for controlling the calorie uptake

by limiting the digestion and absorption of nutritional compounds. However, the formulation of

foodstuffs with high pectin contents must be done carefully in countries whose inhabitants are

deficient in vitamins (e.g., B-complex vitamins and vitamin E) and minerals (e.g., iron, calcium,

and zinc) since the consumption of pectin has been associated with a decreased bioaccesibility of

these nutritional compounds.

Page 243: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Academic production

234

Academic production

List of publications

1. Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L. P., & Narváez-Cuenca, C. E.

(2014). Inhibition of digestive enzyme activities by pectic polysaccharides in model

solutions. Bioactive Carbohydrates and Dietary Fibre. 4: 27-38.

2. Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L. P., Narváez-Cuenca, C. E., &

McClements, D. J. (2014). Impact of dietary fibers [methyl cellulose, chitosan, and pectin] on

digestion of lipids under simulated gastrointestinal conditions.

Food & Function. 5: 3083-3095.

3. Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L. P., Narváez-Cuenca, C. E., &

McClements, D. J. (2014). Interaction of a dietary fiber (pectin) with gastrointestinal

components (bile salts, calcium, and lipase): A calorimetry, electrophoresis, and turbidity

study. Journal of Agricultural and Food Chemistry. 62: 12620-12630.

4. Espinal-Ruiz, M., Restrepo-Sánchez, L. P., Narváez-Cuenca, C. E., & McClements, D. J.

(2016). Impact of pectin properties on lipid digestion under simulated gastrointestinal

conditions: Comparison of citrus and banana passion fruit (Passiflora tripartita var.

mollisima) pectins. Food Hydrocolloids. 52: 329-342.

5. Espinal-Ruiz, M., Restrepo-Sánchez, L. P., & Narváez-Cuenca, C. E. (2016). Effect of

pectins on the mass transfer kinetics of monosaccharides, amino acids, and a corn oil-in-

water emulsion in a Franz diffusion cell. Food Chemistry. 209: 144-153.

Page 244: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Academic production

235

Conferences and meetings

1. Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L. P., & Narváez-Cuenca, C. E.

(2014). Inhibition of digestive enzyme activities by pectic polysaccharides. Presented as Oral

communication. Food Structure and Functionality Forum Symposium: From Molecules to

Functionality. March 30 to April 02. Amsterdam, The Netherlands.

2. Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L. P., & Narváez-Cuenca, C. E.

(2014). Inhibition of digestive enzyme activities by pectic polysaccharides in model

solutions. Presented as Poster. International Food Physics Symposium. May 08. Amherst

(MA), United States.

3. Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L. P., Narváez-Cuenca, C. E., &

McClements, D. J. (2014). Impact of polysaccharides on digestion of lipids under simulated

gastrointestinal conditions. Presented as Oral communication. 2014 Annual Conference &

Exhibition: Functional Foods, Nutraceuticals, Natural Health Products, and Dietary

Supplements. October 14 to 17. Istanbul, Turkey.

4. Espinal-Ruiz, M., Parada-Alfonso, F., Restrepo-Sánchez, L. P., Narváez-Cuenca, C. E., &

McClements, D. J. (2015). Interaction of pectin with gastrointestinal components: A

calorimetry, microelectrophoresis, and turbidity study. Presented as Oral communication.

Delivery of Functionality in Complex Food Systems: Physically-Inspired Approaches from

the Nanoscale to the Microscale. July 14 to 17. Paris, France.

5. Espinal-Ruiz, M., Restrepo-Sánchez, L. P., Narváez-Cuenca, C. E., & McClements, D. J.

(2016). Impact of pectin properties on lipid digestion under simulated gastrointestinal

conditions: Comparison of citrus and banana passion fruit (Passiflora tripartita var.

mollisima) pectins. Presented as Oral communication. 2nd

Food Structure and Functionality

Forum Symposium: From Molecules to Functionality. February 28 to March 02. Singex,

Singapore.

Page 245: Efecto de la pectina sobre la actividad de algunas enzimas ... · Carlos Eduardo Narváez Cuenca, PhD ... Palabras clave: Pectina, grado de metoxilación, enzimas digestivas, emulsión,

Acknowledgements

236

Acknowledgements

I would like to express my gratitude to those who have contributed to this thesis, either directly or

indirectly. First of all, I would like to thank my director, Carlos-Eduardo Narváez-Cuenca, for his

continuous guidance during the PhD program and for his contribution to the development of the

work presented here. Thank you for all your feedback and support, not only within the thesis

project, but also for my personal development.

I also want to thank to professors Luz-Patricia Restrepo-Sánchez and Fabián Parada-Alfonso for

giving me the chance to carry out my doctoral thesis. I am proud having both of you as “co-

directors”. I would like to thank to professors Cecilia Anzola and Laura Ortiz for their

contributions to the thesis. For all involved in the research group: Thanks for the collaboration we

had, it was a great experience working in such a special team: Eliana, Katherine, Elizabeth,

Mayra, Jonathan, Nelson, Mónica, Diego, Paola, Catalina, and Patricia.

A special thanks to Dr. David Julian McClements (Department of Food Science, University of

Massachusetts Amherst) and his team for having received me in the Food Biopolymers and

Colloids Research Laboratory and for helping me to achieve the objectives proposed in this

thesis: Jean, Iris, Amir, Ben, Cheryl, Zachary, Vanessa, Gabriel, Penny, Bengü, Becca, Dr. Hu,

Laura, Tommy, Patrick, Izlia, Eric, Cynthia, Jennifer, Saehun, Jenny, Xuan, and Bicheng.

I would like to thank to my granters for funding me along the thesis: COLCIENCIAS-

COLFUTURO (Convocatoria No. 528, Beca Francisco José de Caldas para estudios de doctorado

en Colombia) and Vicerrectoría Académica from Universidad Nacional de Colombia (Beca

Estudiante Sobresaliente de Posgrado).

Finally, I would like to say a very special thank you for my family (my parents and brothers), and

for my wife. Jeimmy, Meeting you is the best, unexpected thing that happened to me. Thanks for

supporting me and understanding me all these years. I would like to dedicate this thesis to you.

Of course, Gretta could not miss. Thanks for brightening our lives.