Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Carbohydrate Polymers 161 (2017) 118–139

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Review

Pectin as a rheology modifier: Origin, structure, commercial


production and rheology
Siew Yin Chan a,b , Wee Sim Choo a,∗ , David James Young a,b,c,∗ , Xian Jun Loh b,d,e,∗
a
School of Science, Monash University Malaysia, 47500 Bandar Sunway, Selangor Darul Ehsan, Malaysia
b
Institute of Materials Research and Engineering (IMRE), Agency for Science, Technology and Research (A*STAR), 2 Fusionopolis Way, Innovis, #08-03,
Singapore 138634, Singapore
c
Faculty of Science, Health, Education and Engineering, University of the Sunshine Coast, Maroochydore DC, Queensland 4558, Australia
d
Department of Materials Science and Engineering, National University of Singapore, 9 Engineering Drive 1, Singapore 117576, Singapore
e
Singapore Eye Research Institute (SERI), 11 Third Hospital Avenue, Singapore 168751, Singapore

a r t i c l e i n f o a b s t r a c t

Article history: Pectins are a diverse family of biopolymers with an anionic polysaccharide backbone of ␣-1,4-linked d-
Received 21 September 2016 galacturonic acids in common. They have been widely used as emulsifiers, gelling agents, glazing agents,
Received in revised form 2 December 2016 stabilizers, and/or thickeners in food, pharmaceutical, personal care and polymer products. Commercial
Accepted 16 December 2016
pectin is classified as high methoxy pectin (HMP) with a degree of methylation (DM) >50% and low
Available online 24 December 2016
methoxy pectin (LMP) with a DM <50%. Amidated low methoxy pectins (ALMP) can be obtained through
aminolysis of HMP. Gelation of HMP occurs by cross-linking through hydrogen bonds and hydrophobic
Keywords:
forces between the methyl groups, assisted by a high co-solute concentration and low pH. In contrast,
Pectin
Polysaccharide
gelation of LMP occurs by the formation of ionic linkages via calcium bridges between two carboxyl
Rheology modifier groups from two different chains in close proximity, known as the ‘egg-box’ model. Pectin gels exhibit
Extraction Newtonian behaviour at low shear rates and shear-thinning behaviour when the shear rate is increased.
Hydrogels An overview of pectin from its origin to its physicochemical properties is presented in this review.
Gelling mechanisms © 2016 Elsevier Ltd. All rights reserved.
Hydrogen bonding
Hydrophobic interactions
Ionic interactions
Rheology

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
2. Occurrence of pectin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
3. Chemical structure of pectin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
3.1. Homogalacturonan (HGA) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
3.2. Rhamnogalacturonan I (RG-I) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
3.3. Rhamnogalacturonan II (RG-II) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
3.4. Analytical tools for pectin macromolecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
4. Sources of pectin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
5. Safety status and classification of pectins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
6. Extraction and purification of pectin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
6.1. De-esterification of pectin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
6.2. Other feasible extraction additives and methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

∗ Corresponding authors at: Wee Sim Choo is from School of Science, Monash University Malaysia, 47500 Bandar Sunway, Selangor Darul Ehsan, Malaysia. David James
Young is from Faculty of Science, Health, Education and Engineering, University of the Sunshine Coast, Maroochydore DC, Queensland 4558, Australia. Xian Jun Loh is from
Institute of Materials Research and Engineering (IMRE), Agency for Science, Technology and Research (A*STAR), 2 Fusionopolis Way, Innovis, #08-03, Singapore 138634,
Singapore.
E-mail addresses: [email protected] (W.S. Choo), [email protected] (D.J. Young), [email protected] (X.J. Loh).

http://dx.doi.org/10.1016/j.carbpol.2016.12.033
0144-8617/© 2016 Elsevier Ltd. All rights reserved.
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 119

7. Rheological properties of pectin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125


7.1. Pectin solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
7.2. Pectin gelation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
7.2.1. High methoxy pectin (HMP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
7.2.2. Low methoxy pectin (LMP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
7.2.3. Amidated low methoxy pectin (ALMP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
7.3. Intrinsic and extrinsic factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
7.4. Comparison in rheology of HMP, LMP and ALMP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
7.5. Rheology of mixed pectin systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
8. Modification of pectin structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
9. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

1. Introduction decrease the viscosity of a system (Rojas, Rosell, & Benedito de


Barber, 1999; Shi & BeMiller, 2002). Herein we provide an overview
Production of pectin originated in the 1900s in Germany when of pectin in terms of origin, structures, commercial production and
an apple juice manufacturer tried to cook dried apple pomace, most importantly, rheological properties.
the by-product from apple juice processing (Ciriminna, Chavarría-
Hernández, Inés Rodríguez Hernández, & Pagliaro, 2015). Pectin
was thereafter marketed as a gelling agent. It is a hydrocolloid
capable of forming networks to trap water and forming gels at low 2. Occurrence of pectin
concentrations (<1%) (Abid et al., 2017; Chan, Choo, Young, & Loh,
2016; Liang et al., 2012; Vriesmann & Petkowicz, 2013; Ptitchkina, Pectin was first discovered in apple juice by Vauquelin (1790)
Danilova, Doxastakis, Kasapis, & Morris, 1994). The prolonged com- and named by Braconnot (1825a,b), borrowing from the Greek
mercial success of pectin has shown the importance of using fruit word pektikos which means to congeal or solidify. Since then, pectin
by-products as raw materials to produce value-added products. has been studied extensively by scientists from different branches
New application opportunities continue to emerge and pectin is of science. Chemically, pectin is defined as an anionic polysaccha-
no longer just a gelling agent but also used as a stabiliser and thick- ride made up largely by covalently ␣-1,4-linked d-galacturonic acid
ener (Brejnholt, 2009). Pectin is seen as an attractive investment (also known as galactosyluronic acid) (GalA) units (Fig. 1A). It is
and has been industrialized by companies such as CP Kelco, Cargill, derived naturally from plants, in which pectin makes up approx-
Calleva, DSM Yantai Andre Pectin, Dupont, FMC Biopolymers, Herb- imately one third of the cell walls in dicotyledonous and some
streith & Fox, etc and contributes significantly to the global market monocotyledonous plants (Braconnot, 1825a,b). Pectin can also be
of hydrocolloids with an estimated $850 million worth of pectin found in small proportions of cell walls in grass (2–10%) and wood
sold in 2013 (Bomgardner, 2013). This market is expected to con- tissue (5%) (Voragen, Coenen, Verhoef, & Schols, 2009).
tinue growing at a rate of 5–6% per year (Bomgardner, 2013). Today, Pectin in plant tissues is associated with other cell wall com-
the food applications of pectin are diverse, ranging from bever- ponents, such as celluloses or hemicelluloses; playing important
age (Nakamura, Yoshida, Maeda, & Corredig, 2006; Zulueta, Esteve, roles in plant growth and development (Ferrari et al., 2013; Rao
Frasquet, & Frígola, 2007), confectionery (Basu & Shivhare, 2010), & Silva, 2006) (Fig. 1B). Pectin is a common component of young
dairy (Joudaki et al., 2013) to meat processing (Pereira, Marques, tissues, fruits and vegetables (de Assis, Lima, & de Faria Oliveira,
Hatano, & Castro, 2010). Pectin has also attracted substantial atten- 2001). In general, it is most abundant in the middle lamella layers
tion from the pharmaceutical (Günter & Popeyko, 2016), cosmetic between adjoining plant cells (Mellerowicz & Sundberg, 2008). This
(Lupi et al., 2014), and polymer (da Costa, de Mello Ferreira, & de is followed by the primary cell walls of plants which also contain
Macedo Cruz, 2016) industries. fairly high amount of pectin (Cosgrove & Jarvis, 2012). The amount
Rheology is the study of the deformation and flow of matter. It of pectin is greatly decreased or even absent beginning from the
defines the relationship between strain, stress and time. Strain is secondary cell walls of plants towards the plasma membranes of
the change in shape or length of the sample while stress is the plant cells (Willats, McCartney, Mackie, & Knox, 2001) (Fig. 1C).
force per unit area producing the change in shape. When sub- Biologically, pectin has specific functions depending on its loca-
jected to external forces, solids (or truly elastic materials) will tion and molecular structure. Pectin can function as a gel and
deform, whereas liquids (or truly viscous materials) will flow. Con- thereby assisting with cell adhesion and softening cell walls for cell
temporary rheology also involves investigating the behaviour of elongation (Parre & Geitmann, 2005; Suárez et al., 2013; Willats
real materials with properties intermediate between those of ideal et al., 2001). This polysaccharide lends strength and support to
solids and ideal liquids. A rheology modifier is a material that alters plants by maintaining cell consistency and mechanical resistance
the rheology of a fluid composition to which it is added; and thus, (Voragen et al., 2009). Additionally, pectin influences various cell
it plays an essential role in achieving desirable flow characteristics wall properties such as porosity, surface charge, pH, and ion bal-
(Braun & Rosen, 2000). Most products in food, pharmaceutical, per- ance by forming networks and trapping solute molecules that
sonal care or household applications contain rheology modifiers to enable ion transportation (Harholt, Suttangkakul, & Vibe Scheller,
achieve appropriate application characteristics. In this review, we 2010; Voragen et al., 2009). Pectin also activates plant defenses
highlight the use of pectin as a rheology modifier. by stimulating the accumulation of phytoalexins which have a
Pectin is characterized as an emulsifier, gelling agent, glaz- wide spectrum of anti-microbial activity (Benedetti et al., 2015;
ing agent, stabilizer, and/or thickener in commercial applications Hahn, Darvill, & Albersheim, 1981). In addition, pectin oligosaccha-
(Codex Alimentarius, 2015). All of these functional terms are in rides induce accumulation of protease inhibitors in plant tissues as
fact subsets of the term “rheology modifier”. Pectin is a rheology wound response (Bishop, Makus, Pearce, & Ryan, 1981; Ferrari et al.,
modifier that thickens many food matrices; conversely, it can also 2013) and also lignification (Matsunaga et al., 2004; Robertsen,
1986; Xiao & Anderson, 2013).
120 S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139

Fig. 1. [A] Pectin with covalently ␣-1,4-linked d-galacturonic acid (GalA) units, [B] The structure of cell wall. Pectins are often associated with other cell wall components such
as cellulose or hemicellulose in plant tissues; and [C] Pectins are most abundant in middle lamella and primary cell walls, and greatly decreased or are absent in secondary
cell walls towards the plasma membrane (Sticklen, 2008).
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 121

3. Chemical structure of pectin the pectin backbone comprises RGI domain with HG being linked as
its side chain (Vincken et al., 2003); and (c) the pectin backbone is
Although it has been more than 210 years since the discov- composed of alternating, perpendicularly linked-HGA strands and
ery of pectins, their chemical and structural properties are still the a RGI domain (Yapo, 2011a).
subject of investigation due to the inhomogeneity of this polymer
family. To begin with basic structure, pectin is assembled by at
least 17 different monosaccharides, of which GalA (Fig. 2Ai) is the
most plentiful, followed by l-arabinose, d-galactose, l-rhamnose
and others (Kaya, Sousa, Crépeau, Sørensen, & Ralet, 2014). These
3.1. Homogalacturonan (HGA)
monosaccharides can be interconnected through 20 different link-
ages (Kaya et al., 2014). Within GalA monomers, the carboxylic
HGA is a linear homopolymer of ␣-1,4-linked GalA and is known
groups or hydroxyls may be methyl-esterified (Fig. 2Aii) and/or
as the “smooth region” (Fig. 2Bi). It is an abundant and widespread
O-acetyl-esterified (Fig. 2Aiii), respectively. O-acetyl-esterification
domain of pectin, accounting for approximately 60–65% of total
occurs predominantly at the O-3 position and occasionally at the
pectin amount (Voragen et al., 2009). HGA appears to be synthe-
O-2 position.
sized in the Golgi apparatus and then transferred to the middle
Native pectins in plants are composed of polysaccharide
lamella and primary cell walls (Willats et al., 2001). It is then
domains: homogalacturonan (HGA), rhamnogalacturonan I (RG-I),
deposited in a form that has 70–80% of GalA units methyl-esterified
rhamnogalacturonan II (RG-II), xylogalacturonan, and apiogalac-
and could be O-acetyl-esterified at O-2 or O-3 (Mohnen, 2008;
turonan. The latter three have been classified as substituted HGA
O’Neill, Albersheim, & Darvill, 1990). The amount of GalA units
(Caffall & Mohnen, 2009; Harholt et al., 2010; Yapo, 2011a). The
present in a HGA chain is estimated to be around 100–200 units,
deposition and the way in which these domains are joined to one
however, this range could be an underestimate (Bonnin, Dolo, Le
another remains a matter of debate. Currently, it is thought that
Goff, & Thibault, 2002). The smooth region can sometimes be joined
the polysaccharide domains are covalently linked, and to a greater
by one or two ␣-1,2-linked l-rhamnopyranose units and most of the
or lesser extent, ionically cross-linked with other pectin strands to
pectins have this structure. In addition, GalA units may be substi-
form pectic networks that branch throughout the primary cell walls
tuted at the C-2 or C-3 positions with residues of xylose or apiose,
(Fig. 2B) (Harholt et al., 2010; Willats, McCartney, & Knox, 2003).
producing domains known as xylogalacturonan or apiogalacturo-
There are three models proposed: (a) the pectin backbone consists
nan, respectively (Ovodov, 2009). These biosynthetic modifications
of alternating HGA and RGI domains (Schols & Voragen, 1996); (b)
alter the functional properties of the HGA domain.

Fig. 2. [A] Forms of galacturonic acids (GalA) found in pectin: (i) GalA, (ii) Methylated GalA and (iii) O-Acetylated GalA; and [B] Pectin chain comprising of covalently linked
(i) homogalacturonan (HGA), (ii) rhamnogalacturonan I (RG-I) and (iii) rhamnogalacturonan II (RG-II). The diagram shown here is intended only to illustrate some of the
major domains found in most pectins rather than indicate definitive structures.
122 S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139

3.2. Rhamnogalacturonan I (RG-I) & Schols, 2000; Limberg et al., 2000). Pectin is an extremely complex
and structurally diverse group of polymers. The chemical structure
RG-I consists of up to 100 or more repeating units of the dis- of pectin varies between plants, tissues and even within a cell wall
accharide ␣-1,2-linked-l-rhamnose-␣-1,4-d-GalA (Fig. 2Bii). The (Willats, Knox, & Mikkelsen, 2006). Although considerable research
rhamnose residues account for >30% of a pectin (Renard, Crépeau, has been devoted to its chemical structure, there is much to be
& Thibault, 1995). Within these rhamnose residues, 50–78% is esti- done to correlate the functionalities and interactions of each divi-
mated to be RGI (Renard et al., 1995). The GalA residues of RG-I are sion of pectin to its unique biomechanical properties. This requires
not methyl-esterified; however, the GalA residues of RG-I could be thoughtful planning and the application of new analytical tools.
O-acetyl-esterified (Caffall & Mohnen, 2009). In most cases, 20–80%
of rhamnose residues in this domain are substituted at the C-4 posi-
4. Sources of pectin
tion with neutral sugar side chains (Mohnen, 2008). Attachment of
neutral sugar side chains to the C-2, C-3 and/or C-4 positions of
In spite of its availability in a large number of plant species, com-
rhamnose residues are also possible. The neutral sugars side chains
mercial sources of pectin are limited. Historically, apple pomace
are predominantly galactose and arabinose, forming galactan, ara-
has been the major source, but there has been an increasing use of
binan and arabinogalactans. Other sugars such as glucose, mannose,
citrus peel in recent years (Brejnholt, 2009). Today, commercially
fucose, xylose, and glucuronic acid are found covalently linked to
available pectin are mostly extracted from citrus peel (85.5%), fol-
the backbones as side chains. The composition and size of neutral
lowed by apple pomace (14.0%) and to a smaller extent, sugar beet
sugar side chains can be a single glycosyl residue up to 50 or more,
pulp (0.5%) (Ciriminna et al., 2015; Staunstrup, 2009). Citrus peel
resulting in a large and highly variable family of polysaccharides
and apple pomace are available in copious amounts as remain-
with a range of glycosidic linkages (Voragen et al., 2009). The highly
ders from juice and essential oil production (Masmoudi et al.,
branched nature of RG-I has led to the name “hairy region”.
2008; Nussinovitch & Hirashima, 2013; Shalini & Gupta, 2010),
while sugar beet pulp is obtained from the sugar industry (Yapo,
3.3. Rhamnogalacturonan II (RG-II)
Robert, Etienne, Wathelet, & Paquot, 2007). Other feasible sources
are shown in Table 1. Different plant species have different pectin
RG-II is not structurally related to RG-I, despite their names
content and the pectins extracted from each plant species have
running in sequence and the fact that it is also included in the
different physicochemical properties.
hairy region (Fig. 2Biii). It is a branched pectic domain contain-
ing a HGA backbone. RG-II is a highly compact homopolymer of
around nine ␣-1,4-linked GalA units (of which some are methyl- 5. Safety status and classification of pectins
esterified) with four structurally different polymeric side chains
attached. These side chains contain eleven rare sugars including The Joint FAO/WHO Expert Committee on Food Additives
apiose, 2-O-methyl-l-fucose, 2-O-methyl-d-xylose, 3-C-carboxy- (JECFA) has recommended pectin (Codex Alimentarius No. 440) as
5-deoxy-l-xylose (aceric acid), 3-deoxy-d-manno-octulosonic acid a safe additive with no limit on acceptable daily intake. Pectin is
and 3-deoxy-d-lyxoheptulosaric acid (Yapo, 2011b). A fascinat- not digested by humans as there are no enzymes in vivo that are
ing feature of RG-II is that it appears to be the only major pectic able to degrade the molecule (HansUlrich & Frank, 2012). However,
domain that does not have significant structural diversity or mod- certain bacteria of the gut flora such as Bacteroides, Bifidobacterium,
ulation of its fine structure. It is thought that the ends of RG-II Clostridium, Erwinia, Escherichia, and Eubacterium strains are able
are linked glycosidically to the HGA domains (Yapo, 2011a). It is a to use pectin as substrates for fermentation (Dongowski, Lorenz, &
highly conserved and widespread domain isolated from cell walls Anger, 2000; HansUlrich & Frank, 2012). Pectin is quite stable under
by endopolygalacturonase cleavage, indicating its covalent attach- the acidic condition of the stomach, although a de-esterification
ment to HGA. Interestingly, RG-II crosslinks two pectin molecules of a small extent may occur. Without the fermentation process,
(apiosyl residues) within the cell wall by borate ester links. pectin would pass almost unchanged through the digestive sys-
tem. Increasing consumer awareness of a healthy lifestyle and
3.4. Analytical tools for pectin macromolecules the emerging trend to produce functional food have made pectin
popular. It has been reported that pectin has numerous positive
There are multiple analytical tools capable of exploring the influences on health (Endress, 1991) including improving colonic
fine structures and functionalities of the pectin macromolecule health (Min et al., 2015), lowering of cholesterol and serum glucose
such as Fourier transform infrared spectroscopy (FT-IR) (Coimbra, levels (Jones et al., 2015; Zhu et al., 2015), reducing cancer propen-
Barros, Barros, Rutledge, & Delgadillo, 1998), FT-Raman spec- sity (Concha, Weinstein, & Zúñiga, 2013; Wang, Chen, & Lü, 2014),
troscopy (Bichara et al., 2016), gas chromatography (GC) (Huisman, and stimulating the immune response (Bernard et al., 2015).
Oosterveld, & Schols, 2004; Walter & Sherman, 1983), high perfor- Commercial pectin is very heterogeneous. The degree of methyl-
mance liquid chromatography (HPLC) (Levigne, Ralet, & Thibault, esterification (also known as degree of methylation) (DM) of GalA
2002; Voragen, Schols, & Pilnik, 1986), gas chromatography mass units is used to classify pectin (Deuel, 1943). DM is a percentage
spectroscopy (GC–MS) (Savary & Nuñez, 2003), electrospray- which expresses the molar ratio of methyl-esters present to GalA
ionization mass spectroscopy (ESIMS) (Ishii, Ichita, Matsue, Ono, & units (includes both free GalA and substituted GalA). It is the major
Maeda, 2002), ion-exchange chromatography (IEC) (Bonnin et al., parameter affecting gelling, influencing surface tension and emul-
2002; Kravtchenko, Voragen, & Pilnik, 1992b; Ralet, Bonnin, & sion formation (Müller-Maatsch, Caligiani, Tedeschi, Elst, & Sforza,
Thibault, 2001), polysaccharide analysis using carbohydrate gel 2014). There are two types of pectin classified according to DM:
electrophoresis (PACE) (Goubet, Ström, Dupree, & Williams, 2005), low methoxy pectin (LMP) and high methoxy pectin (HMP). HMP
capillary electrophoresis (CE) (Ström, Ralet, Thibault, & Williams, forms gels in high co-solute concentration (55–75%) and acidic (pH
2005; Zhong, Williams, Goodall, & Hansen, 1998) and nuclear mag- 2.50–3.50) systems (Oakenfull & Scott, 1984) while LMP forms gels
netic resonance (NMR) spectroscopy (Cardoso, Silva, & Coimbra, in the presence of cations over a broad pH range (D. Gidley, Morris,
2002). Combinations of these methods with enzymatic finger- Murray, Powell, & Rees, 1980; Hoefler, 1991). Aside from the typical
printing is a powerful suite to elucidate the complex structure LMP and HMP, there is another type of commercial pectin, viz ami-
of pectin (Daas, Arisz, Schols, De Ruiter, & Voragen, 1998; Daas, dated low methoxy pectin (ALMP). ALMP is obtained from ammonia
Meyer-Hansen, Schols, De Ruiter, & Voragen, 1999; Daas, Voragen, de-esterification of HMP (Fig. 3A). The difference between LMP and
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 123

Table 1
Pectin content (%) of commercial and other feasible raw materials used for pectin production.

Raw Materials Producing Pectins Pectin Content (%) References

Commercial Sources
Apple pomace 4.60–20.92 Canteri-Schemin, Fertonani, Waszczynskyj, and Wosiacki (2005), Min et al.
(2011)
Citrus peel Orange peel 10.90–24.80 Kaya et al. (2014), Koubala et al. (2008),
Grapefruit peel 21.60–28.00 Venzon et al. (2015)
Lemon peel 20.90–30.60
Lime peel 9.00–33.60
Sugar beet pulp 4.10–24.96 Lv, Wang, Wang, Li, and Adhikari (2013), Yapo et al. (2007)

Other Feasible Sources


Banana peels 2.40–21.70 Happi Emaga, Ronkart, Robert, Wathelet, and Paquot (2008), Oliveira et al.
(2015)
Carrots (rejected) 8.70–9.10 Christiaens et al. (2015)
Carrot peels (steamed) 8.90–9.10
Celeriac peels (steamed) 15.40–16.40
Cocoa husks 3.38–12.60 Chan and Choo (2013), Vriesmann, de Mello Castanho Amboni, and de
Oliveira Petkowicz (2011)
Creeping fig seeds 5.25–6.07 Liang et al. (2012)
Dragon fruit peel 5.60–26.38 Muhammad, Mohd. Zahari, Gannasin, Mohd. Adzahan, and Bakar (2014)
Durian rind 2.10–10.30 Wai, Alkarkhi, and Easa (2010)
Faba bean hulls 9.57–15.75 Korish (2015)
Green beans cutting waste 8.10–8.30 Christiaens et al. (2015)
Jackfruit peel 8.94–15.19 Begum, Aziz, Uddin, and Yusof (2014)
Leek cutting waste 10.60–11.00 Christiaens et al. (2015)
Mangosteen rind 12.00–12.40 Gan and Latiff (2011)
Mango peel 9.20–31.80 Koubala et al. (2008)
Passion fruit rind 2.25–30.30 Liew, Chin, and Yusof (2014), Seixas et al. (2014)
Papaya peel 11.11–49.83 Koubala et al. (2014)
Pomelo peel 8.32–27.63 Methacanon et al. (2014)
Plum pomace 3.80–21.30 Kosmala et al. (2013)
Rapeseed cake 1.80–6.20 Jeong et al. (2013)
Sisal waste 4.61–19.20 Santos, Espeleta, Branco, and de Assis (2013)
Sunflower head 7.40–11.60 Iglesias and Lozano (2004)
Tomato peel 14.90–83.50 Grassino, Halambek et al. (2016), Namir, Siliha, and Ramadan (2015)
Watermelon rind 13.01–25.79 Prakash Maran, Sivakumar, Thirugnanasambandham, and Sridhar (2014)

ALMP is that there are primary amide groups in the chemical struc- agents may selectively extract pectin with different solubility from
ture, replacing methyl-ester groups on carboxyl groups in HMP. the same starting material (Yapo, 2009a,b). This selectivity, how-
Degree of amidation (DAm) is a percentage which expresses the ever, is less evident when the extraction does not occur in the above
molar ratio of primary amide group present to GalA units (includ- sequence, especially when acid is used first or when these agents
ing both free GalA and substituted GalA). Most commercial ALMP are used individually to extract pectin from the same starting
contains a DAm of 15–25. materials (Yapo, 2009a,b). Acids are generally the highest yield-
Deuel (1943) also proposed calculating the degree of acetyl- ing extracting agents (Yapo, 2009a,b). Another advantage of using
esterification (DA). DA is a percentage which expresses the molar acid extraction is that the pectin obtained are usually enriched with
ratio of acetyl-esters present to GalA units (includes both free GalA GalA units (Rao & Silva, 2006).
and substituted GalA). DA is also an important determinant of On a commercial scale, mineral acids are used to extract pectin.
gelling capacity of pectin; however, it is not as significant as DM Mineral acids are cheaper and more efficient in terms of produc-
in commercial pectin classification. Recently, another popular way tion yield. Commonly used minerals acids include hydrochloric
to classify pectin has been developed, the degree of blockiness (DB). acid, nitric acid and sulphuric acid (Fishman, Chau, Hoagland, &
Daas et al. (1999) introduced DB as a percentage expressing the total Hotchkiss, 2006; Iglesias & Lozano, 2004; Mesbahi, Jamalian, &
amount of non-esterified mono-, di- and tri- GalA liberated from Farahnaky, 2005; Pagán, Ibarz, Llorca, Pagán, & Barbosa-Cánovas,
degradation of pectin by endopolygalacturonase relative to the 2001; Rao & Silva, 2006). Pretreatment such as washing, blanching,
total amount of non-esterified GalA present in the pectin. Through or drying increase the stability of raw materials during trans-
DB, the pattern of methyl-esterification can be known. Moreover, portation and storage. These processes inactivate enzymes such
the pattern of methyl-esterification can radically alter gel forma- as pectin methyl-esterase (PME) and bacteria that would other-
tion and the properties of the gels formed. Guillotin et al. (2005) wise rapidly degrade pectin. To prevent fermentation, peels are
have further introduced the expression “absolute degree of block- dried to 10–12% moisture level. The fresh or dried raw materials are
iness” (DBabs ), which takes the original total GalA units (includes then extracted in demineralized water that has been acidified to pH
both free GalA and substituted GalA) into account. DBabs is defined 1.5–3.0 at 50–100 ◦ C for approximately three hours (Nussinovitch
as the amount of non-esterified mono-, di- and tri- GalA liberated & Hirashima, 2013). The extract containing pectin is separated from
from degradation of pectin by endopolygalacturonase relative to the raw materials by filtration or centrifugation. Polyvalent cations
the total amount of GalA present in the pectin (Fig. 3B). such as aluminum or copper salts are also used to precipitate pectin.
These metal ions are then removed by acidified alcohol washes, fol-
lowed by a wash in alkaline alcohol to neutralize the product. The
6. Extraction and purification of pectin precipitates obtained are washed with alcohol again and pressed
dry. Pectin can be further de-esterified in an alcohol suspension. The
Pectin can be extracted via water, buffers, chelating agents, final steps involve drying and milling to yield powdered pectin, fol-
acids, and/or bases. When used successively in the above- lowed by blending with different production batches to standardize
mentioned order (from water to bases), these different chemical
124 S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139

6.1. De-esterification of pectin

Pectins obtained from industrial production are mostly HMP.


LMP does occur naturally in plants but is usually obtained from
HMP by acid, alkaline and/or enzymatic de-esterification. One of
the most common industrial methods to de-esterify HMP to pro-
duce LMP is alkaline de-esterification by ammonia in alcohol at
relatively low temperatures (Fig. 3B). Others reagents used for
aminolysis include hydroxylamine and other ammonia derivatives
(Racape, Thibault, Reitsma, & Pilnik, 1989). The benefits of using
ALMP include the production of relatively high pectin molecular
weights. In this heterogeneous system, the methyl-esterified GalA
residues undergo saponification at a slower rate than with sodium
hydroxide in water at the same pH. As the time of reaction increases,
the DM decreases, DAm increases and the GalA content remains
constant (Renard & Thibault, 1996). The neutral sugar composition
and number of side chains remain constant (Reitsma, Thibault, &
Pilnik, 1986). The molecular weight of ALMPs do not change dras-
tically relative to the initial HMP used, suggesting that the reaction
does not lead to drastic de-polymerization under these conditions
(Anger & Dongowski, 1988).
Acid de-esterification is achieved by using strong inorganic
acids such as hydrochloric acid. Alkaline de-esterification is usually
performed using sodium hydroxide. However, there are two disad-
vantages of these methods that could lead to pectin degradation,
viz hydrolysis and ˇ-elimination. Hydrolysis is a pH-dependent
reaction and it occurs mainly under acidic conditions; (Krall &
McFeeters, 1998; Smidsrød, Haug, & Larsen, 1966). Acid hydrol-
ysis involves the removal of “ballast” which includes neutral
polysaccharides and to a lesser extent, proteins and phenolic com-
pounds (Constenla & Lozano, 2003; Kravtchenko, Berth, Voragen, &
Pilnik, 1992; Kravtchenko, Voragen, & Pilnik, 1992a; Kravtchenko,
Voragen et al., 1992b). By comparison, ˇ-elimination involves
cleavage of glycosidic bonds, leading to formation of a double bond
between C4 and C5 in the GalA units (BeMiller, 1986; Keijbets &
Pilnik, 1974) (Fig. 3C).
The rate of reaction of acid hydrolysis increases as pH decreases.
The pH of plant cell walls generally ranges between 4–6 (Brett &
Waldron, 1990). Fraeye et al. (2007) found that acid hydrolysis at
this pH range is negligible. The lower the DM, the faster pectin is
hydrolyzed, probably due to a lower amount of methyl-esterified
target groups. This in turn influences gel strength, leading to the
formation of a weaker gel. However, at controlled or reduced tem-
peratures, there is a higher possibility that LMP will be obtained
without extensive main-chain breakdown (Rao & Silva, 2006).
Unlike acid hydrolysis, ˇ-elimination is more temperature-
dependent than it is pH-dependent. A kinetic study by Kravtchenko,
Arnould, Voragen, and Pilnik (1992) revealed that any increase in
temperature increases the rate of ˇ-elimination reaction. It was
reported that the activation energies of ␤-elimination are similar
Fig. 3. [A] Ammonia de-esterification of pectin, yielding amidated low methoxy
at pH values ranging from 4.5 to 11.0. However, this mechanism
pectin (ALMP); [B] Schematic diagram of enzymatic digestion on a 50% degree of
methylation (DM) pectin with endopolygalacturonase from Kluyveromyces fragilis
occurs extensively and rapidly under alkaline conditions, or even
(endo-PGkf) and exopolygalacturonase from Aspergillus tubingensis (exo-PGAt). The at near neutral pH (Rao & Silva, 2006; Taylor, 1982). It can occur at
black and white arrows symbolize the action of endo-PGkf and exo-PGAt, respec- weak acidic pH but the reaction rate is very low. There is a prereq-
tively. Oligomers released by endo-PGkf are indicated as (a). The degree of blockiness uisite for ˇ-elimination to occur, which is the presence of a methyl
(DB) and absolute degree of blockiness (DBabs ) are calculated as illustrated. Figure
group at C6 (Keijbets & Pilnik, 1974). As DM decreases, the reaction
adapted from Guillotin et al. (2005); [C] ˇ-elimination cleavage of glycosidic bond
near the presence of a methyl group at C6, leading to formation of a double bond rate of ˇ-elimination decreases.
between C4 and C5 in the galacturonic acid (GalA) unit; and [D] Random and block Enzyme de-esterification has becoming increasingly popular
distribution of GalA units (either unesterified or methyl-esterified) in pectin with for obtaining LMP in an efficient and environmentally sustainable
50% DM.
manner. Most importantly, enzyme de-esterification is capable of
producing random or block distribution of unesterified GalA units
(Denès, Baron, Renard, Péan, & Drilleau, 2000; Savary, Hotchkiss,
Fishman, Cameron, & Shatters, 2003) (Fig. 3D). One of the most
commonly used enzymes for enzyme de-esterification is pectin
structure and functionality. Dilution with sucrose or dextrose may methyl-esterase (PME). There are two mechanisms proposed for
be used to standardize performance (May, 1990). enzyme de-esterification of pectin using PME: single-chain reac-
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 125

tion of the enzyme acts to produce a block pattern of carboxylates


by moving along linearly on a single pectin chain or multiple-attack
mechanism involving multiple attachments of the enzyme, produc-
ing shorter blocks of de-esterified pectin (Hotchkiss et al., 2002;
Limberg et al., 2000).
Rapid-setting HMP is usually obtained after a short extrac-
tion time at temperatures close to boiling. This is due to short
extraction times with high temperature reduces de-esterification.
Conversely, long extraction times with low temperature favors
de-esterification to produce slow-setting HMP or even LMP
(Nussinovitch and Hirashima, 2013). Therefore, it is important
to select suitable extraction conditions to obtain pectin with the
desired properties (Chan & Choo, 2013).

6.2. Other feasible extraction additives and methods

Recent reports suggest that there is a potential for organic acids


such as acetic, citric, lactic, malic and tartaric acid to be employed in
pectin extraction (Chan & Choo, 2013; Jamsazzadeh Kermani et al.,
2014; Yapo, 2009a,b). Chan and Choo (2013), Kumar and Chauhan
(2010) found that organic acids can achieve higher yields of pectin
compared to conventional inorganic acids. Enzymes can also be
used in isolation or employed as an additive to give higher produc-
tion yield or better quality. Wikiera, Mika, Starzyńska-Janiszewska,
and Stodolak (2015) have demonstrated that multicatalytic enzy-
matic preparation Celluclast 1.5L is efficient for pectin extraction
from apple pomace, comparable to acidic treatment and three-fold
higher than water extraction. Recent reports also describe new
extraction techniques using ultrasound (Grassino, Brnčić, et al.,
2016), ultra-high pressure (Guo et al., 2012), microwave (Hosseini,
Khodaiyan, & Yarmand, 2016b), electric field and electromagnetic
waves (Zouambia, Youcef Ettoumi, Krea, & Moulai-Mostefa, 2016).
Polysaccharides extracted by ultrasound possessed lower viscos-
ity, molecular weight and degree of esterification (DE), but with
a greater degree of branching and purity than conventional heat-
extracted pectin (Wang et al., 2015). Ultra-high pressure may be an
eco-friendlier alternative to produce pectin with higher viscosity
and stability (Guo et al., 2012). Microwave can assist acid extrac-
tions of pectin at relatively low temperatures, in contrast to the
rather high temperatures of conventional hot acid extraction and
that the extraction process can be shortened from hours to minutes
(Fishman, Chau, Cooke, & Hotchkiss, 2008). Seixas et al. (2014) have
suggested that milder, weak organic acids could potentially replace
strong acids if microwave is incorporated to assist the pectin extrac-
tion. Electric fields have been used to shorten the extraction time Fig. 4. [A] The effect of shear rate on the [i] apparent viscosity and [ii] shear stress
and reduce degradation brought about by the extended exposure to of aqueous dispersion of high methyl-esterified pectin from tamarillo pulp in water
heat during the conventional method (de Oliveira, Giordani, Gurak, at 3%, 5% and 8%. Figure adapted from Nascimento et al. (2016); and [B] The effect of
Cladera-Olivera, & Marczak, 2015). Combination of the above meth- shear rate on the viscosity of citrus pectin solution at different pectin concentrations.
Figure reprinted from Sousa et al. (2015).
ods can produce pectin with better quality in more efficient ways.
In summary, by understanding the effect by each extraction
methods, food scientists will be able to design a more energy
proportion of hydrated molecules (Guimarães, Coelho Júnior, &
efficient and greener way to deliver a better end-product with
Garcia Rojas, 2009). The viscosities of pectin solutions increase
enhanced properties. In that way, we can eventually reduce waste
with increasing pectin concentration (Fig. 4Ai). This behaviour has
and pollution from conventional extraction processes and also save
been observed for pectin from apple pomace (Min et al., 2011),
costs on both processing and post-treatment of the waste gener-
cacao pod husks (Vriesmann & Petkowicz, 2013), citrus peel (Sousa,
ated.
Nielsen, Armagan, Larsen, & Sørensen, 2015), dragon fruit peel
(Muhammad et al., 2014), mango pulp (Iagher, Reicher, & Ganter,
7. Rheological properties of pectin 2002), pomelo peel (Methacanon, Krongsin, & Gamonpilas, 2014),
sour orange peel (Hosseini, Khodaiyan, & Yarmand, 2016a) and
7.1. Pectin solution tamarillo fruit pulp (Nascimento, Simas-Tosin, Iacomini, Gorin, &
Cordeiro, 2016). When the concentration is increased continu-
Dilute solutions of pectin contain homogeneously dispersed ously, the intermolecular distances between the pectin molecules
pectin molecules that are too far apart to interact with one decrease, facilitating intermolecular interactions such as hydrogen
another, and thus there is only a slight increase in viscosity due bonding (Guimarães et al., 2009). A linear relationship is observed
to distortion in the velocity pattern of the liquid by an increased between shear stress and shear rate of pectin solutions, indicating
126 S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139

Newtonian behaviour (Fig. 4Aii), but only up to a certain con- and deduced that this was due to an increase in these hydrophobic
centration. Pectin dispersions at concentrations up to 3% behave interactions.
like Newtonian liquids and subsequent increases in concentration Recent studies have demonstrated the importance of the pattern
results in shear thinning behaviour (i.e. a decrease in viscosity with of methylation on gelation of HMP (Löfgren et al., 2005; Slavov et al.,
increase in shear rate) (Fig. 4B). It should be noted that the actual 2009). Pectins gel at a faster rate when they possess a block distribu-
pectin concentration at which the solution transforms from New- tion of methyl esterified-GalA units, relative to pectin with the same
tonian to shear thinning behaviour is dependent on the molecular DM, but with a random distribution of methyl esters (Löfgren et al.,
weight of the pectin used. There are many studies reporting the 2005) (Fig. 5B). Pectins with randomly distributed methyl esters
observation of an increase in pseudoplasticity with an increase in have considerably lower storage moduli (G’) relative to pectins with
pectin concentration (Muhammad et al., 2014; Nascimento et al., a higher DB and therefore the latter form stronger gels.
2016; Sato, Oliveira, & Cunha, 2008; Sousa et al., 2015). This pseu- Sucrose is often used as an additive for HMP gel preparation
doplastic behaviour is magnified with increasing concentration but it can be replaced by other co-solutes such as glucose, fruc-
(characterized by increasing slope at higher shear rate) (Fig. 4B). tose or polyols such as glycerol (Evageliou, Richardson, & Morris,
While at near-zero shear rate, the viscosity of pectin increases with 2000; Tsoga, Richardson, & Morris, 2004a). Hydrophobic interac-
increasing concentration (Fig. 4B). This phenomenon is common for tions depend largely on the molecular geometry of the co-solutes
polysaccharides in which the zero shear viscosity value becomes and the interaction with neighboring water molecules (Matubayasi,
higher with an increase in polymer concentration (Muhammad 1994; Tsoga, Richardson, & Morris, 2004b). Oakenfull and Scott
et al., 2014). It is also noteworthy that the Newtonian plateau limit (1984) deduced that the function of sugar in the formation of gels
is shifted to a low shear rate region with increasing pectin concen- by HMP is to stabilize junction zones by promoting hydrophobic
tration. interaction between methyl-ester groups. The effect of sucrose can
be explained by the concept of “preferential hydration” introduced
7.2. Pectin gelation by Lee and Timasheff (1981). “Preferential hydration” describes the
heterogeneous distribution of solvent around the macromolecules.
Gelling is the formation of a three-dimensional (3D) network Sucrose at a concentration of 50% by weight increases hydrophobic
of polymer chains with solvent and solutes trapped within. (Loh interaction between two methyl esters by 67% compared to water
et al., 2014; Loh, Nam Nguyen, Kuo, & Li, 2011; Nam Nguyen, Kuo, alone (Oakenfull & Scott, 1984).
& Loh, 2011; Wu, Loh, Wu, Lay, & Liu, 2010) The conditions required pH is also an important gelation factor. Pectin is an anion
for gelation and the properties of the gel ultimately depend on the polysaccharide and lowering the pH protonates carboxylic groups,
molecular structure, the intermolecular forces which hold the net- reducing electrostatic repulsions along and between pectin chains
work together and the nature of the junction zones at which the (Thakur et al., 1997). Moreover, at low pH, non-dissociated car-
polymer molecules are cross-linked (Axelos & Thibault, 1991a). The boxylic groups form inter- and intramolecular hydrogen bonds
junction zones in polysaccharide gels are complex and the molec- with secondary alcohol groups.
ular structures are held together by a large number of individually The formation of HMP gels is time-dependent (Seshadri, Weiss,
weak interactions such as electrostatic interactions and hydrogen Hulbert, & Mount, 2003) (Fig. 5C). In this study, a native pectin mix-
bonds. HMP and LMP have different gelation mechanisms, although ture (1.15% apple pectin, 41.40% sucrose, pH 1.50) at time zero was
the gel characteristics are governed by the same macromolecular highly viscous in nature. This solution was then subjected to stress
properties such as the composition, size and conformation of the at different gelation times and the magnitude of the stress response
polymers (Axelos & Thibault, 1991a). to the oscillatory sweeps increased as gelation time increased,
indicating that the solution became more elastic in nature. The
7.2.1. High methoxy pectin (HMP) frequency dependence of the storage modulus (G’) became simul-
HMP gels are obtained at low pH in the presence of a high taneously less pronounced, corresponding to a change from viscous
sucrose concentration or similar co-solute. A high sucrose con- to a rubbery behavior. At t = 225 min, G’ was higher than G”, indi-
centration reduces water activity which is necessary to promote cating that an elastic gel was formed. This rheological data can be
chain–chain interactions rather than chain-solvent interactions; explained in terms of network formation. Initially, there is no net-
while low pH protonates carboxylate residues, minimizing elec- work formed and each pectin molecule behaves very much like
trostatic repulsion (Thibault & Ralet, 2003). Walkinshaw and Arnott a single, dispersed molecule. As the pectin dispersion is sheared,
(1981) inferred from X-ray diffraction (XRD) studies that HMP gels the molecules relax quickly due to their high molecular mobility.
are stabilized by intermolecular hydrogen bonds and hydropho- As a consequence, there is no extensive stress buildup, particu-
bic bonding between methyl esters. This finding was subsequently larly at low oscillation frequencies. As time progresses, the pectin
confirmed by Oakenfull and Scott (1984). Fig. 5A schematically molecules begin to associate by the formation of hydrogen bond
illustrates the gelation of HMP. and with the help of hydrophobic interactions. The resulting 3D
Hydrogen bonds between pectin molecules are favored by the network restricts the freedom of molecules to react to the super-
conformation of adjacent GalA units. Individual hydrogen bonds imposed flow. As a result, the relaxation time increases, which is
are relatively weak, but a large number of them confer significant indicated by a shift of the storage modulus from the viscous region
thermodynamic stability to the gel. Hydrogen bonding is the main to the rubber plateau.
interaction that sustains the HMP gel structure, but it is insuffi- The gelation of HMP is also highly dependent on temperature,
cient to overcome the entropic barrier to gelation. Hydrophobic as demonstrated in a study by da Silva, Gonçalves, and Rao (1995)
interactions are essential for HMP gelation. The contribution of (Fig. 5D). The authors performed gelation of HMP at a tempera-
hydrophobic interactions to the free energy of junction zones in ture of 5 ◦ C and found that gelation was slow and resulted in a
HMP (with DM = 70%) is half that arising from hydrogen bonding weak gel. They attributed this to the lack of hydrophobic interac-
(Oakenfull & Scott, 1984). The unfavourable interaction between tion at low temperature. The pseudo-equilibrium G’ increased for
the non-polar methyl esters of pectin and water molecules result temperatures up to 30 ◦ C and then decreased. It was inferred that
in this hydrophobic effect (Thakur, Singh, Handa, & Rao, 1997). hydrogen bonds and hydrophobic interactions are best facilitated
These methyl esters coalescence, thereby reducing the surface at an intermediate temperature range and contribute to a network
area exposed to water and therefore overall entropy is decreased. with higher elasticity. The structure of HMP-sugar gels is reported
Guimarães et al. (2009) have reported that HMP is more viscous to be irreversible on heating (Rao & Silva, 2006). Evageliou et al.
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 127

(2000) reported an unusual “thermal annealing” behaviour of HMP instead of hydrogen bonds and hydrophobic interactions (Axelos
when a gel was heated and cooled. This behaviour, however, was & Thibault, 1991b) (Fig. 6 A). This is known as the “egg-box” model
pH-dependent. At pH 4.7, a HMP (of DM = 70%, 0.5 wt%) gel in the (Gidley, Morris, Murray, Powell, & Rees, 1979; Grant, Morris, Rees,
presence of 65 wt% sucrose was fully reversible on heating with no Smith, & Thom, 1973; Morris, Powell, Gidley, & Rees, 1982) and
detectable thermal hysteresis. At pH 3.0 and 3.5, there was a slight is characterized as junction zones formed by the ordered, side-
reduction in G’ and G” on heating. However, stability at 30 ◦ C was by-side associations of GalA, whereby specific sequences of GalA
observed to be virtually constant despite the pH difference. monomers in adjacent chains are linked intermolecularly through
electrostatic and ionic bonding of carboxyl groups, forming struc-
7.2.2. Low methoxy pectin (LMP) tures like an egg-box (Braccini & Pérez, 2001). The “egg-box” model
was first introduced to explain the gelation mechanism of alginate.
LMP gels are stabilized by ionic cross-linkages via calcium Although the “egg-box” model can explain the gelation of alginate
bridges between two carboxylates from two different chains, (Low, Chee, Kai, & Loh, 2015), it is only an approximation for the

Fig. 5. [A] HMP gelation is governed by intermolecular hydrogen bonds and hydrophobic interactions between methyl esters; [B] Gelation time for 0.75% of citrus pectin (i)
A, (ii) B and (iii) C with DM 74%, 72% and 82%, respectively, in the presence of 60% sucrose at pH 3. The gelation time was plotted as a function of (iv) degree of blockiness (DB).
Figure adapted from Löfgren, Guillotin, Evenbratt, Schols, and Hermansson (2005); [C] Frequency sweeps on apple pectin dispersion at pH 1.5, 22 ◦ C: (i) storage moduli and
(ii) loss moduli. Figures adapted from Seshadri et al. (2003); and [D] Storage moduli (G’) as a function of ageing time for 1% HMP from citrus in the presence of 60% sucrose at
pH 3 at several ageing temperatures: 5 ◦ C (shaded-triangle), 10 ◦ C (shaded-square), 15 ◦ C (open square), 20 ◦ C (shaded-circle), 30 ◦ C (open circle), 50 (open triangle). Figure
adapted from da Silva et al. (1995).
128 S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139

Fig. 5. (Continued)

mechanism of gelation of LMP. The latter is better described as a gel. Fraeye et al. (2009) further observed that the gel strength of
shifted “egg-box”, since one of the chains is slightly shifted with LMP-calcium gels was better explained by the pattern of methyla-
respect to the other. The “egg-box” models for the gelation of algi- tion than by the degree of methylation. G’ increased as DB increased
nate and pectin have been revisited by Braccini and Pérez (2001); from 0 to 40%, indicating that stronger gels were formed when
Fang et al. (2008); and Ventura, Jammal, and Bianco-Peled (2013). the NMG residues were distributed blockwise (Fig. 6Bii). The same
A two-stage process has been suggested with an initial dimeriza- correlation is true for the DBabs (Fig. 6Biii).
tion of two macromolecules and subsequent aggregation of these Fraeye et al. (2009) reported that the gel strength of
dimers (Axelos & Thibault, 1991b; Rao & Silva, 2006). The junc- LMP-calcium gels increases with increasing GalA concentration.
tions are formed between unbranched, non-esterified galacturonan However, it was suggested that this is also affected by the distri-
blocks bound together by coordinating calcium ions. The intra- bution of non-methylated GalA available for egg-box formation.
and intermolecular interactions of the polymer are modified by Dobies, Kozak, and Jurga (2004) reported that the strength of LMP-
the charge density and distribution of the non-methyl-esterified calcium gel is stronger with a higher concentration of calcium
carboxyl groups. The new conformation changes the placement of ions, as evidenced by the increase in G’ values with calcium con-
carboxyl groups to be outward facing from the polymer backbone, centration. An increase in concentration of calcium ions allows
facilitating inter-chain hydrogen bonding (Gilsenan, Richardson, & additional cross-linking of non-methylated GalA, generating a
Morris, 2000; Mansel et al., 2015). In general, egg-boxes formed denser and more elastic gel network (Basak & Bandyopadhyay,
between two neighbouring chains are stabilized predominantly by 2014; Ngouémazong et al., 2012). The storage moduli of pectin
electrostatic interactions, followed by hydrogen bonds and then by solutions also increase with increasing concentration of divalent
van der Waals interactions. iron ions (Mierczyńska et al., 2015) (Fig. 6C). Addition of calcium
Formation of cooperative egg-boxes is only possible when or iron ions increases pseudoplasticity, providing an alternative
stretches of non-methylated GalA units of a minimum length are to higher concentrations of thickening agents in fluids (Fig. 6Ci
present. The “egg-boxes” are stable only when there are at least six and ii). Conversely, addition of magnesium ions causes a decrease
consecutive carboxyl groups on the interior of each participating in G’, suggesting a weakening of the pectin gel by magnesium
chain (Braccini & Pérez, 2001; Powell, Morris, Gidley, & Rees, 1982). (Mierczyńska et al., 2015) (Fig. 6Ciii). It could be assumed that
The occurrence of methyl ester groups in the primary backbone lim- divalent iron ions could probably bind to pectin in an egg-box like
its the extent of such junction zones leading to formation of the gel. model; however there is sparse information on the interaction of
As DM decreases, LMP-calcium mixtures transform from liquid-like pectin with metal ions other than calcium.
to gel-like with G’ increasing, hence increasing gel strength with Commercially, LMP is usually dispersed in aqueous calcium
decreasing DM (Fraeye et al., 2009) (Fig. 6Bi). With decreasing DM, solutions at about 70 ◦ C and then cooled slowly. At high temper-
there is a higher probability of the number of sequences of non- ature, gelation of LMP is dominated by the formation of egg-box
methylated GalA residues being long enough to form “egg-boxes”. junction zones via calcium bridges, supported by hydrophobic
Formation of a higher amount of “egg-boxes” results in a stronger interactions. On cooling, hydrogen bonding increases, supported
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 129

Fig. 6. (Continued)

by inter-dimer associations facilitating the gelation process. Ran-


dom electrostatic interactions of calcium with single dissociated
carboxyl groups of pectin chains (calcium cross-linking) also pro-
mote the structuring process. When the temperature is increased,
gelation is accelerated. Durand, Bertrand, Clark, and Lips (1990)
reported that the gelation time of LMP in the presence of calcium is
increased considerably by a small increase in the gel formation tem-
perature. The sol-gel transition temperature is influenced mostly by
the amount of pectin and calcium. The binding of calcium ions to
pectin is less stable at high temperature and thus more calcium ions
are required to form an elastically active junction zone (Garnier,
Axelos, & Thibault, 1993). An increase of metal ion concentration
increases the temperature gel point to a concentration limit and
then decreases.
The number of methoxy groups is relatively low in LMP (<50%)
but it should not be ignored and also contributes to gelling through
hydrophobic interactions (Kastner, Einhorn-Stoll, & Senge, 2012).
The free carboxyl groups are assumed to be randomly distributed.
The presence of small amounts of sugar (10–30%) in LMP gel tends
to impart desirable textural properties (Fu & Rao, 2001). The pres-

solution depended on the DM of each pectin and was adjusted so that in the final
gel the stoichiometric ratio R = 2[Ca2+ ]/[COO− ] was equal to 0.2. Figure adapted
from Fraeye et al. (2009); [C] Storage moduli of apple pectin with the addition of
(i) calcium hydroxide [Ca(OH)2 ], (ii) iron lactate [C6 H10 FeO6 ] and (iii) magnesium
Fig. 6. [A] LMP gelation mechanism is governed by ionic cross-linking via calcium hydroxide [Mg(OH)2 ]. Figure adapted from Mierczyńska, Cybulska, Sołowiej, and
(divalent) ions between two carboxyl groups from two different chains in close Zdunek (2015); [D] (i) Gelation of LMP via hydrogen bonds in the presence of acid,
proximity; [B] Effect of (i) degree of methylation (DM), (ii) degree of blockiness (DB) (ii) Evolution of storage moduli for 1 and 2% acid LMP gels and time resolved small-
and (iii) absolute degree of blockiness on storage moduli of pectin de-esterified angle X-ray scattering (SAXS) for the corresponding gels recorded at four different
by sodium hydroxide (symbol: rhombus), pectin de-esterified by Aspergillus times after mixing (tw ). Figures adapted from Mansel et al. (2015); and [E] Storage
aculeatus pectin methyl-esterase (PME) (symbol: square) and pectin de-esterified moduli (G’) as a function of time for 1% pectin samples at pH 3 with the absence of
by tomato PME (symbol: triangle). Data points placed between brackets represent salt (open symbol), and the addition of 0.05 M LiCl (light gray), 0.05 M NaCl (dark
pectin-calcium mixtures that did not exhibit gel character. The final GalA concen- gray) and 0.05 M KCl (black). The temperature from the left to the vertical dotted line
tration was fixed at 43 mM. The calcium concentration of the added calcium chloride was decreased from 40 ◦ C to 5 ◦ C over time, and kept constant at T = 5 ◦ C throughout
the time frame investigated. Figure adapted from Ström, Schuster, and Goh (2014).
130 S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139

Fig. 6. (Continued)

ence of sugar can also promote inter-chain interactions, as in HMP 21 helices are highly extended and reduction in temperature and/or
gelation. However, high concentrations of sugar will affect the pH can promote a conformational transition to the more compact
gelling of LMP adversely as the dehydrating effect of sugar favours 31 helices. This transition is reversible. Gelation occurs by dimer-
hydrogen bonding and decreases cross-linking by calcium ions (Fu ization of (antiparallel) three-fold helices, with more extensive
& Rao, 2001). aggregation at very low pH where the chains become essentially
Recent studies have reported new ways to make LMP gels via uncharged (Gilsenan et al., 2000). Protonation of carboxyl groups
acid-induced (Gilsenan et al., 2000; Ström et al., 2007) and mono- appears to promote conformational ordering and association by
valent ion-induced gelation (Ström et al., 2014; Wehr, Menzies, & two different mechanisms: (a) suppression of electrostatic repul-
Blamey, 2004). Pectin has a conformational transition below a cer- sion, and (b) allowing the carboxyl groups to act as hydrogen-bond
tain pH that depends on both the temperature and pH and induces donors (Gilsenan et al., 2000).
aggregation and eventually gelation. At low pH and/or low tem- Calcium-pectin interactions are essentially cooperative at low
perature, the conformation of pectin are predominantly 31 helices ionic strength; however, if the DM is too low, pectin can be precipi-
(Gilsenan et al., 2000). At high pH and/or high temperature, the tated out of solution instead of forming a stable gel. The number
solution conformation are predominantly 21 helices, with only lim- of dissociated free carboxyl groups is relatively low at a pH of
ited conversion to the 31 form on cooling (Gilsenan et al., 2000). The about 3 (i.e. below the pKa value of 3.5). Therefore, the formation
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 131

of typical egg-box-junction zones is limited with more interac- calcium. Monovalent ions such as sodium, potassium and lithium
tions between undissociated carboxyl groups via hydrogen bonds induce a stronger acid gel for LMP by permitting both hydrogen and
instead. The hydrogen bonds formed between protonated and ionic bonding (Ström et al., 2014) (Fig. 6E).
unprotonated carboxyl groups are weakened by protonation and
so acid-induced pectin gels are not stable at very low pH (Gilsenan
et al., 2000; Mansel et al., 2015) (Fig. 6Di). As gelation proceeds,
small-angle X-ray scattering (SAXS) data indicates that the entan-
7.2.3. Amidated low methoxy pectin (ALMP)
gled pectin solution starts to aggregate and form cross-links with
Löfgren, Guillotin, and Hermansson (2006) have demonstrated
time (Schuster, Cucheval, Lundin, & Williams, 2011), but not at low
that ALMP and LMP appear very similar in gel microstructure and
pH (Mansel et al., 2015) (Fig. 6Dii). In the early stages of network
kinetic behaviour at pH 3 and pH 7. ALMP shows rapid gel formation
build-up, pectin molecules in solution are far apart. The polymers
with G’ > G” throughout, with and without 30% sucrose (Fig. 7A). The
then cluster, thereby lowering the effective concentration of the
presence of sucrose influences neither the kinetic behaviour nor
solution. As more clusters nucleate, grow and connect, a stronger
the microstructure of the gels, but strongly increases the storage
network is slowly formed, as reflected by the increase in G’. Acid-
modulus. ALMP generally has reduced sensitivity towards calcium
induced gels have less junction zones than pectin-calcium gels,
(Löfgren et al., 2006). However, some studies have suggested that
highlighting the importance of hydrogen bond formation (Mansel
amidation might have little influence on sensitivity towards cal-
et al., 2015).
cium (Capel, Nicolai, Durand, Boulenguer, & Langendorff, 2006). By
Ström et al. (2014) have reported that gelling of LMP at low pH
comparison, ALMP strongly favours acid-induced gelation (Capel
may be assisted by monovalent ions up to a critical limit (Wang, Lee,
et al., 2006; Lootens et al., 2003). The hydrogen bonds between
Wang, & Huang, 2007). Interestingly, ionic cross-linking by diva-
amide groups help to promote gel formation at low pH. Gels from
lent ions is inhibited by monovalent ions such as sodium, which
ALMP are pH sensitive and thermally reversible. They can be heated
tie up the free carboxyl groups. However, monovalent ions can also
and solidified again on cooling, while conventional pectin-gels
be beneficial, improving the solubility of LMP in the presence of
remain liquid under the same conditions (Chen et al., 2014).

Fig. 7. [A] G’ and G” during gel formation of 8% pectin G (non-amidated pectin) at (a) pH 3; (b) pH 7. Normalized G’ values during gel formation of 0.8% pectin G and pectin
C30 (amidated pectin) at (c) pH 3; (d) pH 7. The initial 9 min involves a temperature decrease from 80 to 25 ◦ C. Figure reprinted from Löfgren et al. (2006); and [B] Effect of
(i) sonication intensity and (ii) sonication time on the development of storage moduli of apple pectin gels. Figures adapted from Seshadri et al. (2003).
132 S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139

Table 2
Molecular weight and rhamnose content (%) of commercial pectin.

Pectins Molecular Rhamnose Neutral sugar References


weight content (%) (%)
(Da)

Apple 63,000–81,000 2.30 12.60 Schmidt, Schmidt, Kurz, Endreß, and Schuchmann (2015),
Citrus 38,000–162,000 1.40 5.10 Yapo et al. (2007), Kravtchenko, Voragen et al. (1992a),
Sugar beet 20,200–90,100 5.40 6.80–32.90 Leroux, Langendorff, Schick, Vaishnav, and Mazoyer (2003)

Table 3
Young’s modulus of high methoxyl pectin (HMP), low methoxyl pectin (LMP) and amidated low methoxyl pectin (ALMP) gels (Vithanage et al., 2010).

Young’s modulus (Nm−2 )


Pectins
At 5 ◦ C At 20 ◦ C

HMP (6.40 ± 0.20) × 103 (4.91 ± 0.20) × 103


LMP (1.80 ± 0.05) × 103 (9.50 ± 0.05) × 102
ALMP (7.90 ± 0.08) × 102 (5.60 ± 0.01) × 102

7.3. Intrinsic and extrinsic factors ing extraction or purification and are therefore not usually present
in significant amounts in commercial pectins.
The gelation process is affected by both intrinsic and extrinsic
parameters. Intrinsic factors include the number and distribution 7.4. Comparison in rheology of HMP, LMP and ALMP
pattern of free carboxyl groups, the molecular weight and types of
pectin while extrinsic factors include pectin concentration, calcium Comparing the three commercially available pectins, HMP
concentration, pH, ionic strength and temperature (Gigli, Garnier, shows the highest apparent viscosity, followed by LMP and ALMP
& Piazza, 2009). (G. Vithanage, Grimson, Wills, Harrison, & Smith, 2010). Young’s
Comparing commercial pectin from different sources; apple modulus is used to define the force displacement of matter. It is
pectin produces a more elastic-viscous gel and citrus pectin pro- determined by calculating the slope from the plot of stress versus
duces a more elastic-brittle gel. By contrast, sugar beet pectin is strain. Vithanage et al. (2010) found that Young’s modulus for HMP
less effective as a gelling agent. However, it possesses emulsify- gel was the highest, followed by LMP and ALMP (Table 3). In other
ing properties (Chen, Fu, & Luo, 2016) and is capable of forming words, HMP gel requires a higher force to break its structure com-
covalently cross-linked hydrogels and so is of commercial value pared to LMP and ALMP gels, indicating that hydrogen bonding in
(May, 1990). Gelation of pectin is influenced by molecular weight HMP pectin is stronger than the ionic bonding (with calcium ions)
because the rigidity of the gel is determined by the number of effec- in LMP and ALMP. Young’s modulus is temperature-dependent
tive junctions formed per chain (Axelos & Thibault, 1991a). The (Table 3).
lower the molecular weight, the less junction zones per molecule HMP gels demonstrate the broadest range of elastic behaviour
can be formed, decreasing the extent of cross-linking and thereby in puncture and strain sweep tests, followed by LMP and ALMP
weakening the gel. Although molecular weight is mainly affected gels (Vithanage et al., 2010). Comparing LMP and ALMP gels, the
by the origin of the pectin, handling methods such as extraction strength of the former is reported to be greater than that of the
also affect the molecular weight of the polymer significantly. For latter; however, ALMP gels exhibit a higher storage modulus than
example, ultrasound may be used to increase the yield of pectin LMP gels (Vithanage et al., 2010). The G’ and G” values for ALMP gel
extraction. However, Seshadri et al. (2003) found that the rheo- increase with increasing temperature, forming a more rubbery elas-
logical properties of pectin that had been treated with ultrasound tic medium; whereas HMP and LMP gels show a decrease of G’ and
were inferior. As sonication intensity and time was increased, gel G” with increasing temperature, forming a more liquidly medium
strength was reduced. As shown in Fig. 7Bi and ii, G’ decreased as (Vithanage et al., 2010). This is postulated to be due to the forma-
sonication intensity and time increased. This result was attributed tion of stable junctions by hydrogen bonding through amide groups
to the reduction in the molecular weight of pectin by cavitational in ALMP. At lower temperature, hydrogen bonds are favoured, rein-
promoted chain degradation. forcing the junction zones; while at higher temperatures, the loss of
The presence of sugar monomers such as rhamnose, whose hydrogen bonding is compensated for with reinforced hydrophobic
dimensions are not compatible with the geometry of the junction interactions (Alonso-Mougán, Meijide, Jover, Rodrıı́guez-Núñez, &
zones formed by GalA monomers, obstructs the molecular orien- Vázquez-Tato, 2002). The sol/gel transition of LMP is sensitive to
tation necessary for junction-zone formation (Axelos & Thibault, the ionic strength of the medium, while the viscoelastic properties
1991a). Rhamnose influences the conformation of the polymer in of the gel structure is time-dependant and resistant to moderate
solution and ultimately its gelling properties (Oakenfull, 1991). This temperature (retained up to 60 ◦ C) (Fu & Rao, 2001; Gigli et al.,
sugar is most abundant in rhamnogalacturonan, disturbing the reg- 2009).
ularity of the galacturonan backbone and forming the “hairy region”
which limits inter-chain association. This could further explain 7.5. Rheology of mixed pectin systems
the reason sugar beet pectin exhibited the poorest gelling prop-
erties of the three commercial pectins, as it possesses a relatively Synergistic mixed gels are also of interest to industry. Syner-
high rhamnose content (Table 2). However, it should be noted that gism can lead to different microstructures to that of pure gels with
the regularity and frequency of the rhamnose interruptions to the improved texture and enhanced rheological properties that result
pectin chain are different in pectins from different plant species in cost savings during manufacturing. Synergistic effects can be
or even within the same species. The presence of neutral sug- achieved by mixing HMP, LMP and/or ALMP to produce a strength-
ars also limit inter-chain association (Kravtchenko, Voragen et al., ened mixed gel. The G’ of mixed HMP/LMP gel is higher than that of
1992a) (Table 2). However, these branches are often removed dur- HMP or LMP (Löfgren & Hermansson, 2007). Fig. 8A demonstrates
that G’ values for mixed HMP/LMP are considerably higher than
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 133

for HMP gels. The mixed gel exhibits similar rheological behaviour Polysaccharides and proteins are two major components of
to HMP gels of higher sugar concentration. The strong aggregation food products. They are often used to control structure, texture,
of LM pectin in the presence of sucrose and calcium contributes and stability. Textural and structural properties depend largely
to a large influence on the storage modulus of the mixed gels, on these biopolymers forming organized structures. The mixture
increasing the storage modulus by about thirty times (based on of different components or even the same class of biopolymer
60% sucrose) (Löfgren & Hermansson, 2007). By comparison, the (carbohydrate–carbohydrate or protein-protein) can lead to phase
addition of HMP in a mixed gel based on 30% sucrose increased the separation. There are two types of phase separation: segregative
storage modulus five times (Löfgren & Hermansson, 2007). When which results in an unmixingof the two phases; and associative
ALMP is mixed with HMP, it is possible to obtain high viscosity and which leads to complex formation of the two biopolymers. Associa-
reduced pseudoplasticity (Sato et al., 2008). tive phase separation has been widely applied to stabilize products
in the food, cosmetic, pharmaceutical, medical and biotechnolog-

Fig. 8. [A] Storage moduli during gel formation in mixed 0.8% HMP/0.4% LMP compared to 1.2% HMP gels. A, B, and HMC are HMP while C30 and LMC are LMP. (a), (b) and
(c) were gels formed in the presence of 60% sucrose while (d), (e) and (f) were gels formed in the presence of 45% sucrose. Figure reprinted from Löfgren and Hermansson
(2007); [B] (i) Temperature dependences of a fixed dynamic moduli (frequency of 1.2 rad/s) for acid aqueous mixtures of pectin-chitosan at the compositions indicated. (ii)
Frequency dependences of storage modulus and loss modulus at different temperatures (stages) of the gel-forming process for pectin-chitosan mixtures at compositions
indicated. Figures adapted from Nordby, Kjøniksen, Nyström, and Roots (2003); [C] (i) The complex viscosity, ␩*, storage modulus G’, and loss modulus G” versus angular
frequency for ␤-lacglobulin-pectin coaservates at composition 5:1 prepared at a concentration of sodium chloride = 0.02 M. (ii) The complex viscosity, ␩*, storage modulus G’,
and loss modulus G” versus angular frequency for 2% pectin solution prepared at a concentration of sodium chloride = 0.02 M. (iii) The storage modulus G’ versus frequency
curves for ␤-lacglobulin-pectin coaservates at different compositions prepared at a concentration of sodium chloride = 0.11 M. Figures adapted from Xiaoyong Wang et al.
(2007); and [D] Temperature sweep of low-fat spread with varying ratios of fish gelatin and pectin. Figure reprinted from Cheng, Lim, Chow, Chong, and Chang (2008).
134 S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139

Fig. 8. (Continued)

ical industries. In the pectin-casein system, for example, pectin interconnected gel-like network structure. The typical viscoelas-
can prevent aggregation of casein micelles or be the cause of it tic behaviour of pectin solution (Fig. 8C) has been observed in
(Maroziene & de Kruif, 2000), changing the microstructure stabil- coacervates with similar pectin content, indicating that the elas-
ity and rheology of dairy products. Pectin and milk protein are tic behaviour of ␤-lacglobulin-pectin coaservates is primarily due
generally not considered compatible at neutral pH; however, the to the interactions between ␤-lacglobulin molecules and pectin
interaction can be mediated by calcium bridging between pectin chains. Sadahira, Rodrigues, Akhtar, Murray, and Netto (2016)
and casein micelles. deduced that egg white and pectin complexes increased continu-
Agoda-Tandjawa, Durand, Gaillard, Garnier, and Doublier (2012) ous phase viscosity and enhanced foam stability by slowing liquid
demonstrated that the presence of LMP does not modify the drainage. When incorporated with fish gelatin in a low-fat spread,
viscoelasticity and microstructural properties of microfibrillated an increase in pectin content could also increase the consistency
cellulose suspensions but it does increase the shear-thinning and melt ability of the low-fat spread (Cheng et al., 2008) (Fig. 8D). It
behaviour and thixotropic characteristic of the suspensions. They was also predicted that the low-fat spread sample with more pectin
suggested that this outcome was due to the viscosifier effect of content would perform better with ‘melt in the mouth’ character-
Nordby et al. (2003) demonstrated that a pectin-chitosan mix- istics and a better instant in-mouth flavour release effect.
ture at different ratios at very low pH (≈1.7) behaves more elastic
at lower temperatures (Fig. 8Bi). At temperatures below the gel 8. Modification of pectin structure
temperature, G’ is always larger than G”, indicating that the pectin-
chitosan mixture is in a solid-like state. At temperatures above Pectin formulations can be manipulated to achieve gels, 3-
the gel temperature (42–50 ◦ C), G” is always larger than G’, indi- D matrices, films and micro-/nano-particles. However, there can
cating that the mixture is in a more viscous-like state (Fig. 8Bii). be a lack of reproducible performance due to the large diversity
Khondkar, Tester, Hudson, Karkalas, and Morrow (2007) used phos- of molecular structures, leading to problems in quality control
phate cross-linking between starch and pectin to prepare gels with and quality assurance (Günter et al., 2014). Solutions to this
greater elasticity and better microstructure. complication fall into two categories: the development of new
Wang et al. (2007) have demonstrated that increasing the technologies for pectin isolation, purification, and characteriza-
amount of pectin in a protein-polysaccharide mixture favours tion and the modification of pectin macromolecules (Liu, Fishman,
the formation of stronger gels. The significantly higher G’ than & Hicks, 2007). Chen et al. (2014) recently published a thorough
G” indicates that ␤-lacglobulin-pectin coaservates form a highly review on chemical modifications of pectins detailing substitutions,
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 135

chain elongation and depolymerization. Alkylation and thiolation References


are among these many modifications related to modifying gelling
behavior. Through modification, the limitations of pectin such as Abid, M., Cheikhrouhou, S., Renard, C. M. G. C., Bureau, S., Cuvelier, G., Attia, H., &
Ayadi, M. A. (2017). Characterization of pectins extracted from pomegranate
poor solubility in organic solvents could possibly be resolved and a peel and their gelling properties. Food Chemistry, 215, 318–325.
greater range of functionalities and applications may be achieved. Agoda-Tandjawa, G., Durand, S., Gaillard, C., Garnier, C., & Doublier, J. L. (2012).
Rheological behaviour and microstructure of microfibrillated cellulose
suspensions/low-methoxyl pectin mixed systems. Effect of calcium ions.
Carbohydrate Polymers, 87(2), 1045–1057.
Alonso-Mougán, M., Meijide, F., Jover, A., Rodrıı́guez-Núñez, E., & Vázquez-Tato, J.
(2002). Rheological behaviour of an amide pectin. Journal of Food Engineering,
9. Conclusion 55(2), 123–129.
Anger, H., & Dongowski, G. (1988). Amidated pectins—Characterization and
Pectins are a family of versatile biopolymers abundant in enzymatic degradation. Food Hydrocolloids, 2(5), 371–379.
Axelos, M. A. V., & Thibault, J. F. (1991a). The chemistry of low-Methoxyl pectin
plants and used commercially as emulsifiers, gelling agents, glaz- gelation. In R. H. Walter (Ed.), The chemistry and technology of pectin (pp.
ing agents, stabilizers, and/or thickeners. The carboxylic groups 109–118). San Diego: Academic Press.
of pectins are usually methyl-esterified to some degree and this Axelos, M. A. V., & Thibault, J. F. (1991b). Influence of the substituents of the
carboxyl groups and of the rhamnose content on the solution properties and
biosynthetic modification alters their functional properties. HMP flexibility of pectins. International Journal of Biological Macromolecules, 13(2),
requires high co-solute content and acid to form gels whereas LMP 77–82.
can form gels in the absence of a co-solute and at a wider range Basak, R., & Bandyopadhyay, R. (2014). Formation and rupture of Ca2+ induced
pectin biopolymer gels. Soft Matter, 10(37), 7225–7233.
of pH values. Pectin produced industrially is typically HMP. LMP Basu, S., & Shivhare, U. S. (2010). Rheological, textural, micro-structural and
does occur naturally in plants but is usually obtained commercially sensory properties of mango jam. Journal of Food Engineering, 100(2), 357–365.
from HMP by acid, alkaline and/or enzymatic de-esterification. One BeMiller, J. N. (1986). An introduction to pectins: Structure and properties. In
Chemistry and function of pectins. pp. 2–12. Washington, DC: American
of the most common de-esterification methods is aminolysis, yield- Chemical Society.
ing ALMP. The presence of a primary amide group in the chemical Begum, R., Aziz, M. G., Uddin, M. B., & Yusof, Y. A. (2014). Characterization of
structure alters the gelation properties of pectin. jackfruit (Artocarpus heterophyllus) waste pectin as influenced by various
extraction conditions. Agriculture and Agricultural Science Procedia, 2, 244–251.
Pectin exhibits Newtonian behaviour at low shear rates and
Benedetti, M., Pontiggia, D., Raggi, S., Cheng, Z., Scaloni, F., Ferrari, S., . . . & De
pseudoelastic behaviour when shear rates are increased. The pre- Lorenzo, G. (2015). Plant immunity triggered by engineered in vivo release of
cise role of each structural feature on rheology and the gelling oligogalacturonides, damage-associated molecular patterns. Proceedings of the
mechanism is of considerable commercial interest. The gelation National Academy of Sciences, 112(17), 5533–5538.
Bernard, H., Desseyn, J.-L., Gottrand, F., Stahl, B., Bartke, N., & Husson, M.-O. (2015).
process is affected by both intrinsic and extrinsic factors includ- Pectin- derived acidic oligosaccharides improve the outcome of pseudomonas
ing the number and distribution pattern of free carboxyl groups, aeruginosa lung infection in C57BL/6 mice. Public Library of Science, 10(11),
the molecular weight and types of pectin, pectin concentration, e0139686.
Bichara, L. C., Alvarez, P. E., Fiori Bimbi, M. V., Vaca, H., Gervasi, C., & Brandán, S. A.
calcium ion concentration, pH, ionic strength and temperature. (2016). Structural and spectroscopic study of a pectin isolated from citrus peel
Understanding these influences enables us to predict the rheolog- by using FTIR and FT-Raman spectra and DFT calculations. Infrared Physics &
ical behaviour of pectin during industrial processing. Technology, 76, 315–327.
Bishop, P. D., Makus, D. J., Pearce, G., & Ryan, C. A. (1981). Proteinase
It is also important to understand the rheology and net- inhibitor-inducing factor activity in tomato leaves resides in oligosaccharides
work structures of mixed pectin-protein and pectin-polysaccharide enzymically released from cell walls. Proceedings of the National Academy of
systems that are used commercially for cosmetic, pharmaceu- Sciences, 78(6), 3536–3540.
Bomgardner, M. M. (2013). Pushing pectin chemical & engineering news 91.
tical, medical and biotechnological applications. Pectin can be Washington, DC: American Chemical Society.
incorporated into mixed structures and stabilize other system Bonnin, E., Dolo, E., Le Goff, A., & Thibault, J.-F. (2002). Characterisation of pectin
(polysaccharides and/or proteins), or vice versa. Synergism can lead subunits released by an optimised combination of enzymes. Carbohydrate
Research, 337(18), 1687–1696.
to a different microstructure from that of pure gels and can improve
Braccini, I., & Pérez, S. (2001). Molecular Basis of Ca2+ -Induced Gelation in Alginates
gel quality. Although extensive studies have been conducted on the and Pectins:The Egg-Box Model Revisited. Biomacromolecules, 2(4), 1089–1096.
rheology of mixed systems involving pectin, there are many other Braconnot, H. (1825a). Nouvelles observations sur l’acide pectique. Annales De
potential combinations that have yet to be investigated. Chimie Et De Physique, 30, 96–102.
Braconnot, H. (1825b). Recherches sur un nouvel acide universellement répandu
Pectin is a natural, biocompatible, biodegradable and renewable dans tous les végétaux. Annales De Chimie Et De Physique, 28(2), 173–178.
polysaccharide but suffers from the irreproducibility and inhomo- Braun, D. B., & Rosen, M. R. (2000). Preface. In rheology modifiers handbook −
geneity of its structure. This complicates quality assurance and Practical use and application. New York, United States of America: William
Andrew Publishing.
control. Structural modification can greatly alter the solubility Brejnholt, S. M. (2009). Pectin. In Food stabilisers, thickeners and gelling agents. pp.
and gelling properties of pectins. Collaboration between different 237–265. Wiley-Blackwell.
disciplines of science can provide a better understanding of this Brett, C., & Waldron, K. (1990). Cell wall structure and the skeletal functions of the
wall. In Physiology and biochemistry of plant cell walls. pp. 4–57. Dordrecht,
biopolymer’s properties so that its true commercial potential may Netherlands: Springer.
be realized. While it is common to focus on applications of pectin in Caffall, K. H., & Mohnen, D. (2009). The structure, function, and biosynthesis of
the food and pharmaceutical industries, pectin may potentially be plant cell wall pectic polysaccharides. Carbohydrate Research, 344(14),
1879–1900.
used in a number of other industries as a renewable and biocom- Canteri-Schemin, M. H., Fertonani, H. C. R., Waszczynskyj, N., & Wosiacki, G.
patible material. (2005). Extraction of pectin from apple pomace. Brazilian Archives of Biology
and Technology, 48, 259–266.
Capel, F., Nicolai, T., Durand, D., Boulenguer, P., & Langendorff, V. (2006). Calcium
and acid induced gelation of (amidated) low methoxyl pectin. Food
Hydrocolloids, 20(6), 901–907.
Cardoso, S. M., Silva, A. M. S., & Coimbra, M. A. (2002). Structural characterisation of
Acknowledgements the olive pomace pectic polysaccharide arabinan side chains. Carbohydrate
Research, 337(10), 917–924.
The authors thank B. Q. Y. Chan and C. E. Wong for proofread- Chan, S.-Y., & Choo, W.-S. (2013). Effect of extraction conditions on the yield and
chemical properties of pectin from cocoa husks. Food Chemistry, 141(4),
ing the manuscript. The authors acknowledge B. Q. Y. Chan and C. 3752–3758.
Owh for their assistance with illustrating the schematic diagrams in Chan, S., Choo, W., Young, D., & Loh, X. (2016). Thixotropic supramolecular
this manuscript. S. Y. Chan acknowledges the funding and support pectin-poly(ethylene glycol) methacrylate (PEGMA) hydrogels. Polymers,
8(11), 404.
from Monash University Malaysia and the Institute for Materials
Research and Engineering, A*STAR.
136 S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139

Chen, J., Liu, W., Liu, C.-M., Li, T., Liang, R.-H., & Luo, S.-J. (2014). Pectin Fraeye, I., De Roeck, A., Duvetter, T., Verlent, I., Hendrickx, M., & Van Loey, A.
modifications: A review. Critical Reviews in Food Science and Nutrition, 55(12), (2007). Influence of pectin properties and processing conditions on thermal
1684–1698. pectin degradation. Food Chemistry, 105(2), 555–563.
Chen, H.-m., Fu, X., & Luo, Z.-g. (2016). Effect of molecular structure on emulsifying Fraeye, I., Doungla, E., Duvetter, T., Moldenaers, P., Van Loey, A., & Hendrickx, M.
properties of sugar beet pulp pectin. Food Hydrocolloids, 54(Part A), 99–106. (2009). Influence of intrinsic and extrinsic factors on rheology of
Cheng, L. H., Lim, B. L., Chow, K. H., Chong, S. M., & Chang, Y. C. (2008). Using fish pectin–calcium gels. Food Hydrocolloids, 23(8), 2069–2077.
gelatin and pectin to make a low-fat spread. Food Hydrocolloids, 22(8), Fu, J. T., & Rao, M. A. (2001). Rheology and structure development during gelation
1637–1640. of low-methoxyl pectin gels: The effect of sucrose. Food Hydrocolloids, 15(1),
Christiaens, S., Uwibambe, D., Uyttebroek, M., Van Droogenbroeck, B., Van Loey, A. 93–100.
M., & Hendrickx, M. E. (2015). Pectin characterisation in vegetable waste Günter, E. A., & Popeyko, O. V. (2016). Calcium pectinate gel beads obtained from
streams: A starting point for waste valorisation in the food industry. LWT − callus cultures pectins as promising systems for colon-targeted drug delivery.
Food Science and Technology, 61(2), 275–282. Carbohydrate Polymers, 147, 490–499.
Ciriminna, R., Chavarría-Hernández, N., Inés Rodríguez Hernández, A., & Pagliaro, Günter, E. A., Popeyko, O. V., Markov, P. A., Martinson, E. A., Litvinets, S. G., Durnev,
M. (2015). Pectin: A new perspective from the biorefinery standpoint. Biofuels E. A., . . . & Ovodov, Y. S. (2014). Swelling and morphology of calcium pectinate
Bioproducts and Biorefining, 9(4), 368–377. gel beads obtained from Silene vulgaris callus modified pectins. Carbohydrate
Codex Alimentarius. (2005). General standards for food additives, codex stan Polymers, 103(0), 550–557.
192–1995. Pectin. Rome, Italy: Food and Agriculture Organization of the United Gan, C.-Y., & Latiff, A. A. (2011). Extraction of antioxidant pectic-polysaccharide
Nations., p. 169. from mangosteen (Garcinia mangostana) rind: Optimization using response
Coimbra, M. A., Barros, A., Barros, M., Rutledge, D. N., & Delgadillo, I. (1998). surface methodology. Carbohydrate Polymers, 83(2), 600–607.
Multivariate analysis of uronic acid and neutral sugars in whole pectic samples Garnier, C., Axelos, M. A. V., & Thibault, J.-F. (1993). Phase diagrams of
by FT-IR spectroscopy. Carbohydrate Polymers, 37(3), 241–248. pectin-calcium systems: Influence of pH, ionic strength, and temperature on
Concha, J., Weinstein, C., & Zúñiga, M. (2013). Production of pectic extracts from the gelation of pectins with different degrees of methylation. Carbohydrate
sugar beet pulp with antiproliferative activity on a breast cancer cell line. Research, 240(0), 219–232.
Frontiers of Chemical Science and Engineering, 7(4), 482–489. Gidley, M. J., Morris, E. R., Murray, E. J., Powell, D. A., & Rees, D. A. (1979).
Constenla, D., & Lozano, J. E. (2003). Kinetic model of pectin demethylation. Latin Spectroscopic and stoicheiometric characterisation of the calcium-mediated
American Applied Research, 33, 91–95. association of pectate chains in gels and in the solid state. Journal of the
Cosgrove, D., & Jarvis, M. (2012). Comparative structure and biomechanics of plant Chemical Society, Chemical Communications, 22, 990–992.
primary and secondary cell walls. Frontiers in Plant Science, 3(204). Gidley, M. J., Morris, E. R., Murray, E. J., Powell, D. A., & Rees, D. A. (1980). Evidence
Daas, P. J. H., Arisz, P. W., Schols, H. A., De Ruiter, G. A., & Voragen, A. G. J. (1998). for two mechanisms of interchain association in calcium pectate gels.
Analysis of partially methyl-esterified galacturonic acid oligomers by International Journal of Biological Macromolecules, 2(5), 332–334.
high-performance anion-exchange chromatography and matrix-assisted laser Gigli, J., Garnier, C., & Piazza, L. (2009). Rheological behaviour of low-methoxyl
desorption/ionization time-of-flight mass spectrometry. Analytical pectin gels over an extended frequency window. Food Hydrocolloids, 23(5),
Biochemistry, 257(2), 195–202. 1406–1412.
Daas, P. J. H., Meyer-Hansen, K., Schols, H. A., De Ruiter, G. A., & Voragen, A. G. J. Gilsenan, P. M., Richardson, R. K., & Morris, E. R. (2000). Thermally reversible
(1999). Investigation of the non-esterified galacturonic acid distribution in acid-induced gelation of low-methoxy pectin. Carbohydrate Polymers, 41(4),
pectin with endopolygalacturonase. Carbohydrate Research, 318(1–4), 135–145. 339–349.
Daas, P. J. H., Voragen, A. G. J., & Schols, H. A. (2000). Characterization of Goubet, F., Ström, A., Dupree, P., & Williams, M. A. K. (2005). An investigation of
non-esterified galacturonic acid sequences in pectin with pectin methylesterification patterns by two independent methods: capillary
endopolygalacturonase. Carbohydrate Research, 326(2), 120–129. electrophoresis and polysaccharide analysis using carbohydrate gel
da Costa, M. P. M., de Mello Ferreira, I. L., & de Macedo Cruz, M. T. (2016). New electrophoresis. Carbohydrate Research, 340(6), 1193–1199.
polyelectrolyte complex from pectin/chitosan and montmorillonite clay. Grant, G. T., Morris, E. R., Rees, D. A., Smith, P. J. C., & Thom, D. (1973). Biological
Carbohydrate Polymers, 146, 123–130. interactions between polysaccharides and divalent cations: the egg-box
da Silva, J. A. L., Gonçalves, M. P., & Rao, M. A. (1995). Kinetics and thermal model. FEBS Letters, 32(1), 195–198.
behaviour of the structure formation process in HMP/sucrose gelation. Grassino, A. N., Brnčić, M., Vikić-Topić, D., Roca, S., Dent, M., & Brnčić, S. R. (2016).
International Journal of Biological Macromolecules, 17(1), 25–32. Ultrasound assisted extraction and characterization of pectin from tomato
de Assis, S. A., Lima, D. C., & de Faria Oliveira, O. M. M. (2001). Activity of waste. Food Chemistry, 198, 93–100.
pectinmethylesterase, pectin content and vitamin C in acerola fruit at various Grassino, A. N., Halambek, J., Djaković, S., Rimac Brnčić, S., Dent, M., & Grabarić, Z.
stages of fruit development. Food Chemistry, 74(2), 133–137. (2016). Utilization of tomato peel waste from canning factory as a potential
de Oliveira, C. F., Giordani, D., Gurak, P. D., Cladera-Olivera, F., & Marczak, L. D. F. source for pectin production and application as tin corrosion inhibitor. Food
(2015). Extraction of pectin from passion fruit peel using moderate electric Hydrocolloids, 52, 265–274.
field and conventional heating extraction methods. Innovative Food Science & Guillotin, S. E., Bakx, E. J., Boulenguer, P., Mazoyer, J., Schols, H. A., & Voragen, A. G.
Emerging Technologies, 29, 201–208. J. (2005). Populations having different GalA blocks characteristics are present
Denès, J.-M., Baron, A., Renard, C. M. G. C., Péan, C., & Drilleau, J.-F. (2000). Different in commercial pectins which are chemically similar but have different
action patterns for apple pectin methylesterase at pH 7.0 and 4.5. Carbohydrate functionalities. Carbohydrate Polymers, 60(3), 391–398.
Research, 327(4), 385–393. Guimarães, G. C., Coelho Júnior, M. C., & Garcia Rojas, E. E. (2009). Density and
Deuel, H. (1943). Pektin als hochmolekularer elektrolyt. Mitteilungen aus dem kinematic viscosity of pectin aqueous solution†. Journal of Chemical &
Gebiete der Lebensmitteluntersuchung und Hygiene, 34, 41–51. Engineering Data, 54(2), 662–667.
Dobies, M., Kozak, M., & Jurga, S. (2004). Studies of gelation process investigated by Guo, X., Han, D., Xi, H., Rao, L., Liao, X., Hu, X., & Wu, J. (2012). Extraction of pectin
fast field cycling relaxometry and dynamical rheology: The case of aqueous from navel orange peel assisted by ultra-high pressure, microwave or
low methoxyl pectin solution. Solid State Nuclear Magnetic Resonance, 25(1–3), traditional heating: A comparison. Carbohydrate Polymers, 88(2), 441–448.
188–193. Hahn, M. G., Darvill, A. G., & Albersheim, P. (1981). Host-pathogen Interactions:
Dongowski, G., Lorenz, A., & Anger, H. (2000). Degradation of pectins with different XIX. The endogenous elicitor, a fragment of a plant cell wall polysaccharide
degrees of esterification by Bacteroides thetaiotaomicron isolated from human that elicits phytoalexin accumulation in soybeans. Plant Physiology, 68(5),
gut flora. Applied and Environmental Microbiology, 66(4), 1321–1327. 1161–1169.
Durand, D., Bertrand, C., Clark, A. H., & Lips, A. (1990). Calcium-induced gelation of HansUlrich, E., & Frank, M. (2012). Pectin. In Dietary fiber and health. pp. 385–408.
low methoxy pectin solutions — Thermodynamic and rheological CRC Press.
considerations. International Journal of Biological Macromolecules, 12(1), 14–18. Happi Emaga, T., Ronkart, S. N., Robert, C., Wathelet, B., & Paquot, M. (2008).
Endress, H. U. (1991). Nonfood uses of pectin. In R. H. Walter, & S. Taylor (Eds.), The Characterisation of pectins extracted from banana peels (Musa AAA) under
chemistry and technology of pectin (pp. 251–268). San Diego: Academic Press. different conditions using an experimental design. Food Chemistry, 108(2),
Evageliou, V., Richardson, R. K., & Morris, E. R. (2000). Effect of pH, sugar type and 463–471.
thermal annealing on high-methoxy pectin gels. Carbohydrate Polymers, 42(3), Harholt, J., Suttangkakul, A., & Vibe Scheller, H. (2010). Biosynthesis of pectin. Plant
245–259. Physiology, 153(2), 384–395.
Fang, Y., Al-Assaf, S., Phillips, G. O., Nishinari, K., Funami, T., & Williams, P. A. Hoefler, A. C. (1991). Other pectin food products. In R. H. Walter, & S. Taylor (Eds.),
(2008). Binding behavior of calcium to polyuronates: Comparison of pectin The chemistry and technology of pectin (pp. 51–66). San Diego: Academic Press.
with alginate. Carbohydrate Polymers, 72(2), 334–341. Hosseini, S. S., Khodaiyan, F., & Yarmand, M. S. (2016a). Aqueous extraction of
Ferrari, S., Savatin, D., Sicilia, F., Gramegna, G., Cervone, F., & De Lorenzo, G. (2013). pectin from sour orange peel and its preliminary physicochemical properties.
Oligogalacturonides: Plant damage-associated molecular patterns and International Journal of Biological Macromolecules, 82, 920–926.
regulators of growth and development. Frontiers in Plant Science, 4(49). Hosseini, S. S., Khodaiyan, F., & Yarmand, M. S. (2016b). Optimization of microwave
Fishman, M. L., Chau, H. K., Hoagland, P. D., & Hotchkiss, A. T. (2006). assisted extraction of pectin from sour orange peel and its physicochemical
Microwave-assisted extraction of lime pectin. Food Hydrocolloids, 20(8), properties. Carbohydrate Polymers, 140, 59–65.
1170–1177. Hotchkiss, A. T., Savary, B. J., Cameron, R. G., Chau, H. K., Brouillette, J., Luzio, G. A.,
Fishman, M. L., Chau, H. K., Cooke, P. H., & Hotchkiss, A. T., Jr. (2008). Global & Fishman, M. L. (2002). Enzymatic modification of pectin to increase its
structure of microwave-assisted flash-extracted sugar beet pectin. Journal of calcium sensitivity while preserving its molecular weight. Journal of
Agricultural and Food Chemistry, 56(4), 1471–1478. Agricultural and Food Chemistry, 50(10), 2931–2937.
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 137

Huisman, M. M. H., Oosterveld, A., & Schols, H. A. (2004). Fast determination of the from creeping fig (Ficus pumila Linn.) seeds. Carbohydrate Polymers, 87(1),
degree of methyl esterification of pectins by head-space GC. Food 76–83.
Hydrocolloids, 18(4), 665–668. Liew, S. Q., Chin, N. L., & Yusof, Y. A. (2014). Extraction and characterization of
Iagher, F., Reicher, F., & Ganter, J. L. M. S. (2002). Structural and rheological pectin from passion fruit peels. Agriculture and Agricultural Science Procedia, 2,
properties of polysaccharides from mango (Mangifera indica L.) pulp. 231–236.
International Journal of Biological Macromolecules, 31(1–3), 9–17. Limberg, G., Körner, R., Buchholt, H. C., Christensen, T. M. I. E., Roepstorff, P., &
Iglesias, M. T., & Lozano, J. E. (2004). Extraction and characterization of sunflower Mikkelsen, J. D. (2000). Analysis of different de-esterification mechanisms for
pectin. Journal of Food Engineering, 62(3), 215–223. pectin by enzymatic fingerprinting using endopectin lyase and
Ishii, T., Ichita, J., Matsue, H., Ono, H., & Maeda, I. (2002). Fluorescent labeling of endopolygalacturonase II from A. Niger. Carbohydrate Research, 327(3),
pectic oligosaccharides with 2-aminobenzamide and enzyme assay for pectin. 293–307.
Carbohydrate Research, 337(11), 1023–1032. Liu, L., Fishman, M., & Hicks, K. (2007). Pectin in controlled drug delivery – A
Jamsazzadeh Kermani, Z., Shpigelman, A., Kyomugasho, C., Van Buggenhout, S., review. Cellulose, 14(1), 15–24.
Ramezani, M., Van Loey, A. M., & Hendrickx, M. E. (2014). The impact of Loh, X. J., Nam Nguyen, V. P., Kuo, N., & Li, J. (2011). Encapsulation of basic
extraction with a chelating agent under acidic conditions on the cell wall fibroblast growth factor in thermogelling copolymers preserves its bioactivity.
polymers of mango peel. Food Chemistry, 161, 199–207. Journal of Materials Chemistry, 21(7), 2246–2254.
Jeong, H.-S., Kim, H.-Y., Ahn, S. H., Oh, S. C., Yang, I., & Choi, I.-G. (2013). Effects of Loh, X. J., Gan, H. X., Wang, H., Tan, S. J. E., Neoh, K. Y., Jean Tan, S. S., . . . & Chan, K.
combination processes on the extraction of pectins from rapeseed cake H. (2014). New thermogelling poly(ether carbonate urethane)s based on
(Brassica napus L.). Food Chemistry, 139(1–4), 9–15. pluronics F127 and poly(polytetrahydrofuran carbonate). Journal of Applied
Jones, M., Gu, X., Stebbins, N., Crandall, P., Ricke, S., & Lee, S.-O. (2015). Effects of Polymer Science, 131(5) [n/a–n/a.
soybean pectin on blood glucose and insulin responses in healthy men. The Lootens, D., Capel, F., Durand, D., Nicolai, T., Boulenguer, P., & Langendorff, V.
FASEB Journal, 29(1 Supplement). (2003). Influence of pH, Ca concentration, temperature and amidation on the
Joudaki, H., Mousavi, M., Safari, M., Razavi, S. H., Emam-Djomeh, Z., & gelation of low methoxyl pectin. Food Hydrocolloids, 17(3), 237–244.
Gharibzahedi, S. M. T. (2013). A practical optimization on salt/high-methoxyl Low, Z. W., Chee, P. L., Kai, D., & Loh, X. J. (2015). The role of hydrogen bonding in
pectin interaction to design a stable formulation for Doogh. Carbohydrate alginate/poly(acrylamide-co-dimethylacrylamide) and alginate/poly(ethylene
Polymers, 97(2), 376–383. glycol) methyl ether methacrylate-based tough hybrid hydrogels. RSC
Kastner, H., Einhorn-Stoll, U., & Senge, B. (2012). Structure formation in sugar Advances, 5(71), 57678–57685.
containing pectin gels −Influence of Ca2+ on the gelation of low-methoxylated Lupi, F. R., Gabriele, D., Seta, L., Baldino, N., de Cindio, B., & Marino, R. (2014).
pectin at acidic pH. Food Hydrocolloids, 27(1), 42–49. Rheological investigation of pectin-based emulsion gels for pharmaceutical
Kaya, M., Sousa, A. G., Crépeau, M.-J., Sørensen, S. O., & Ralet, M.-C. (2014). and cosmetic uses. Rheologica Acta, 54(1), 41–52.
Characterization of citrus pectin samples extracted under different conditions: Lv, C., Wang, Y., Wang, D., Li, L.-j., & Adhikari, B. (2013). Optimization of production
Influence of acid type and pH of extraction. Annals of Botany, 114(6), yield and functional properties of pectin extracted from sugar beet pulp.
1319–1326. Carbohydrate Polymers, 95(1), 233–240.
Keijbets, M. J. H., & Pilnik, W. (1974). ␤-Elimination of pectin in the presence of Müller-Maatsch, J., Caligiani, A., Tedeschi, T., Elst, K., & Sforza, S. (2014). Simple and
anions and cations. Carbohydrate Research, 33(2), 359–362. validated quantitative 1 H NMR method for the determination of methylation,
Khondkar, D., Tester, R. F., Hudson, N., Karkalas, J., & Morrow, J. (2007). Rheological acetylation, and feruloylation degree of pectin. Journal of Agricultural and Food
behaviour of uncross-linked and cross-linked gelatinised waxy maize starch Chemistry, 62(37), 9081–9087.
with pectin gels. Food Hydrocolloids, 21(8), 1296–1301. Mansel, B. W., Chu, C.-Y., Leis, A., Hemar, Y., Chen, H.-L., Lundin, L., & Williams, M.
Korish, M. (2015). Faba bean hulls as a potential source of pectin. Journal of Food A. K. (2015). Zooming in: Structural investigations of rheologically
Science and Technology-Mysore, 52(9), 6061–6066. characterized hydrogen-bonded low-methoxyl pectin networks.
Kosmala, M., Milala, J., Kołodziejczyk, K., Markowski, J., Zbrzeźniak, M., & Renard, C. Biomacromolecules, 16(10), 3209–3216.
M. G. C. (2013). Dietary fiber and cell wall polysaccharides from plum (Prunus Maroziene, A., & de Kruif, C. G. (2000). Interaction of pectin and casein micelles.
domestica L.) fruit, juice and pomace: Comparison of composition and Food Hydrocolloids, 14(4), 391–394.
functional properties for three plum varieties. Food Research International, Masmoudi, M., Besbes, S., Chaabouni, M., Robert, C., Paquot, M., Blecker, C., & Attia,
54(2), 1787–1794. H. (2008). Optimization of pectin extraction from lemon by-product with
Koubala, B. B., Kansci, G., Mbome, L. I., Crépeau, M. J., Thibault, J. F., & Ralet, M. C. acidified date juice using response surface methodology. Carbohydrate
(2008). Effect of extraction conditions on some physicochemical Polymers, 74(2), 185–192.
characteristics of pectins from Améliorée and Mango mango peels. Food Matsunaga, T., Ishii, T., Matsumoto, S., Higuchi, M., Darvill, A., Albersheim, P., &
Hydrocolloids, 22(7), 1345–1351. O’Neill, M. A. (2004). Occurrence of the primary cell wall polysaccharide
Koubala, B. B., Christiaens, S., Kansci, G., Van Loey, A. M., & Hendrickx, M. E. (2014). rhamnogalacturonan II in pteridophytes, lycophytes, and bryophytes.
Isolation and structural characterisation of papaya peel pectin. Food Research Implications for the evolution of vascular plants. Plant Physiology, 134(1),
International, 55, 215–221. 339–351.
Krall, S. M., & McFeeters, R. F. (1998). Pectin hydrolysis:Effect of temperature, Matubayasi, N. (1994). Matching-mismatching of water geometry and hydrophobic
degree of methylation, pH, and calcium on hydrolysis Rates. Journal of hydration. Journal of the American Chemical Society, 116(4), 1450–1456.
Agricultural and Food Chemistry, 46(4), 1311–1315. May, C. D. (1990). Industrial pectins: Sources, production and applications.
Kravtchenko, T. P., Arnould, I., Voragen, A. G. J., & Pilnik, W. (1992). Improvement of Carbohydrate Polymers, 12(1), 79–99.
the selective depolymerization of pectic substances by chemical ␤-elimination Mellerowicz, E. J., & Sundberg, B. (2008). Wood cell walls: Biosynthesis,
in aqueous solution. Carbohydrate Polymers, 19(4), 237–242. developmental dynamics and their implications for wood properties. Current
Kravtchenko, T. P., Berth, G., Voragen, A. G. J., & Pilnik, W. (1992). Studies on the Opinion in Plant Biology, 11(3), 293–300.
intermolecular distribution of industrial pectins by means of preparative size Mesbahi, G., Jamalian, J., & Farahnaky, A. (2005). A comparative study on functional
exclusion chromatography. Carbohydrate Polymers, 18(4), 253–263. properties of beet and citrus pectins in food systems. Food Hydrocolloids, 19(4),
Kravtchenko, T. P., Voragen, A. G. J., & Pilnik, W. (1992c). Analytical comparison of 731–738.
three industrial pectin preparations. Carbohydrate Polymers, 18(1), 17–25. Methacanon, P., Krongsin, J., & Gamonpilas, C. (2014). Pomelo (Citrus maxima)
Kravtchenko, T. P., Voragen, A. G. J., & Pilnik, W. (1992d). Studies on the pectin: Effects of extraction parameters and its properties. Food Hydrocolloids,
intermolecular distribution of industrial pectins by means of preparative 35, 383–391.
ion-exchange chromatography. Carbohydrate Polymers, 19(2), 115–124. Mierczyńska, J., Cybulska, J., Sołowiej, B., & Zdunek, A. (2015). Effect of Ca2+ , Fe2+
Kumar, A., & Chauhan, G. S. (2010). Extraction and characterization of pectin from and Mg2+ on rheological properties of new food matrix made of modified cell
apple pomace and its evaluation as lipase (steapsin) inhibitor. Carbohydrate wall polysaccharides from apple. Carbohydrate Polymers, 133, 547–555.
Polymers, 82(2), 454–459. Min, B., Lim, J., Ko, S., Lee, K.-G., Lee, S. H., & Lee, S. (2011). Environmentally friendly
Löfgren, C., & Hermansson, A.-M. (2007). Synergistic rheological behaviour of preparation of pectins from agricultural byproducts and their
mixed HM/LM pectin gels. Food Hydrocolloids, 21(3), 480–486. structural/rheological characterization. Bioresource Technology, 102(4),
Löfgren, C., Guillotin, S., Evenbratt, H., Schols, H., & Hermansson, A.-M. (2005). 3855–3860.
Effects of calcium, pH, and blockiness on kinetic rheological behavior and Min, B., Koo, O. K., Park, S. H., Jarvis, N., Ricke, S. C., Crandall, P. G., & Lee, S.-O.
microstructure of HM pectin gels. Biomacromolecules, 6(2), 646–652. (2015). Fermentation patterns of various pectin sources by human fecal
Löfgren, C., Guillotin, S., & Hermansson, A.-M. (2006). Microstructure and kinetic microbiota. Food and Nutrition Sciences, 6(12), 1103.
rheological behavior of amidated and nonamidated LM pectin gels. Mohnen, D. (2008). Pectin structure and biosynthesis. Current Opinion in Plant
Biomacromolecules, 7(1), 114–121. Biology, 11(3), 266–277.
Lee, J. C., & Timasheff, S. N. (1981). The stabilization of proteins by sucrose. Journal Morris, E. R., Powell, D. A., Gidley, M. J., & Rees, D. A. (1982). Conformations and
of Biological Chemistry, 256(14), 7193–7201. interactions of pectins. Journal of Molecular Biology, 155(4), 507–516.
Leroux, J., Langendorff, V., Schick, G., Vaishnav, V., & Mazoyer, J. (2003). Emulsion Muhammad, K., Mohd Zahari, N. I., Gannasin, S. P., Mohd Adzahan, N., & Bakar, J.
stabilizing properties of pectin. Food Hydrocolloids, 17(4), 455–462. (2014). High methoxyl pectin from dragon fruit (Hylocereus polyrhizus) peel.
Levigne, S., Ralet, M.-C., & Thibault, J.-F. (2002). Characterisation of pectins Food Hydrocolloids, 42(Part 2), 289–297.
extracted from fresh sugar beet under different conditions using an Nakamura, A., Yoshida, R., Maeda, H., & Corredig, M. (2006). The stabilizing
experimental design. Carbohydrate Polymers, 49(2), 145–153. behaviour of soybean soluble polysaccharide and pectin in acidified milk
Liang, R.-h., Chen, J., Liu, W., Liu, C.-m., Yu, W., Yuan, M., & Zhou, X.-q. (2012). beverages. International Dairy Journal, 16(4), 361–369.
Extraction, characterization and spontaneous gel-forming property of pectin
138 S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139

Nam Nguyen, V. P., Kuo, N., & Loh, X. J. (2011). New biocompatible thermogelling pectin products. In F. Voragen, H. Schols, & R. Visser (Eds.), Advances in pectin
copolymers containing ethylene-butylene segments exhibiting very low and pectinase research (pp. 345–361). Dordrecht: Springer Netherlands.
gelation concentrations. Soft Matter, 7(5), 2150–2159. Schmidt, U. S., Schmidt, K., Kurz, T., Endreß, H. U., & Schuchmann, H. P. (2015).
Namir, M., Siliha, H., & Ramadan, M. F. (2015). Fiber pectin from tomato pomace: Pectins of different origin and their performance in forming and stabilizing
characteristics, functional properties and application in low-fat beef burger. oil-in-water-emulsions. Food Hydrocolloids, 46, 59–66.
Journal of Food Measurement and Characterization, 9(3), 305–312. Schols, H. A., & Voragen, A. G. J. (1996). Complex pectins: structure elucidation
Nascimento, G. E. d., Simas-Tosin, F. F., Iacomini, M., Gorin, P. A. J., & Cordeiro, L. M. using enzymes. In J. Visser, & A. G. J. Voragen (Eds.), Progress in biotechnology
C. (2016). Rheological behavior of high methoxyl pectin from the pulp of (pp. 3–19). Elsevier.
tamarillo fruit (Solanum betaceum). Carbohydrate Polymers, 139, 125–130. Schuster, E., Cucheval, A., Lundin, L., & Williams, M. A. K. (2011). Using SAXS to
Ngouémazong, D. E., Jolie, R. P., Cardinaels, R., Fraeye, I., Van Loey, A., Moldenaers, reveal the degree of bundling in the polysaccharide junction zones of
P., & Hendrickx, M. (2012). Stiffness of Ca2+-pectin gels: Combined effects of microrheologically distinct pectin gels. Biomacromolecules, 12(7), 2583–2590.
degree and pattern of methylesterification for various Ca2+ concentrations. Seixas, F. L., Fukuda, D. L., Turbiani, F. R. B., Garcia, P. S., Petkowicz, C. L. d. O.,
Carbohydrate Research, 348, 69–76. Jagadevan, S., & Gimenes, M. L. (2014). Extraction of pectin from passion fruit
Nordby, M. H., Kjøniksen, A.-L., Nyström, B., & Roots, J. (2003). Thermoreversible peel (Passiflora edulis f. flavicarpa) by microwave-induced heating. Food
gelation of aqueous mixtures of pectin and chitosan. Rheology. Hydrocolloids, 38, 186–192.
Biomacromolecules, 4(2), 337–343. Seshadri, R., Weiss, J., Hulbert, G. J., & Mount, J. (2003). Ultrasonic processing
Nussinovitch, A., & Hirashima, M. (2013). Cooking innovations: Using hydrocolloids influences rheological and optical properties of high-methoxyl pectin
for thickening, gelling, and emulsification. London: CRC Press. dispersions. Food Hydrocolloids, 17(2), 191–197.
O’Neill, M., Albersheim, P., & Darvill, A. (1990). 12 − The pectic polysaccharides of Shalini, R., & Gupta, D. K. (2010). Utilization of pomace from apple processing
primary cell walls. In D. E. Y. P.M (Ed.), Methods in plant biochemistry (pp. industries: A review. Journal of Food Science and Technology, 47(4), 365–371.
415–441). Academic Press. Shi, X., & BeMiller, J. N. (2002). Effects of food gums on viscosities of starch
Oakenfull, D., & Scott, A. (1984). Hydrophobic interaction in the gelation of high suspensions during pasting. Carbohydrate Polymers, 50(1), 7–18.
methoxyl pectins. Journal of Food Science, 49(4), 1093–1098. Slavov, A., Garnier, C., Crépeau, M.-J., Durand, S., Thibault, J.-F., & Bonnin, E. (2009).
Oakenfull, D. G. (1991). The chemistry of high-methoxyl pectins. In R. H. Walter, & Gelation of high methoxy pectin in the presence of pectin methylesterases and
S. Taylor (Eds.), The chemistry and technology of pectin (pp. 87–108). San Diego: calcium. Carbohydrate Polymers, 77(4), 876–884.
Academic Press. Smidsrød, O., Haug, A., & Larsen, B. (1966). The influence of pH on the rate of
Oliveira, T. Í. S., Rosa, M. F., Cavalcante, F. L., Pereira, P. H. F., Moates, G. K., Wellner, hydrolysis of acidic polysaccharides. Acta Chemica Scandinavica, 20, 1026–1034.
N., . . . & Azeredo, H. M. C. (2015). Optimization of pectin extraction from Sousa, A. G., Nielsen, H. L., Armagan, I., Larsen, J., & Sørensen, S. O. (2015). The
banana peels with citric acid by using response surface methodology. Food impact of rhamnogalacturonan-I side chain monosaccharides on the
Chemistry, 198, 113–118. rheological properties of citrus pectin. Food Hydrocolloids, 47, 130–139.
Ovodov, Y. S. (2009). Current views on pectin substances. Russian Journal of Staunstrup, J. (2009). Citrus pectin production and world market. The international
Bioorganic Chemistry, 35(3), 269–284. citrus & beverage conference.
Pagán, J., Ibarz, A., Llorca, M., Pagán, A., & Barbosa-Cánovas, G. V. (2001). Extraction Sticklen, M. B. (2008). Plant genetic engineering for biofuel production: towards
and characterization of pectin from stored peach pomace. Food Research affordable cellulosic ethanol. Nature Reviews Genetics, 9(6), 433–443.
International, 34(7), 605–612. Ström, A., Ralet, M.-C., Thibault, J.-F., & Williams, M. A. K. (2005). Capillary
Parre, E., & Geitmann, A. (2005). Pectin and the role of the physical properties of the electrophoresis of homogeneous pectin fractions. Carbohydrate Polymers, 60(4),
cell wall in pollen tube growth of Solanum chacoense. Planta, 220(4), 582–592. 467–473.
Pereira, C. M., Marques, M. F., Hatano, M. K., & Castro, I. A. (2010). Effect of the Ström, A., Ribelles, P., Lundin, L., Norton, I., Morris, E. R., & Williams, M. A. K. (2007).
partial substitution of soy proteins by highly methyl-esterified pectin on Influence of pectin fine structure on the mechanical properties of
chemical and sensory characteristics of sausages. Food Science and Technology calcium-pectin and acid-pectin gels. Biomacromolecules, 8(9), 2668–2674.
International, 16(5), 401–407. Ström, A., Schuster, E., & Goh, S. M. (2014). Rheological characterization of acid
Powell, D. A., Morris, E. R., Gidley, M. J., & Rees, D. A. (1982). Conformations and pectin samples in the absence and presence of monovalent ions. Carbohydrate
interactions of pectins: II. Influence of residue sequence on chain association in Polymers, 113, 336–343.
calcium pectate gels. Journal of Molecular Biology, 155(4), 517–531. Suárez, C., Zienkiewicz, A., Castro, A. J., Zienkiewicz, K., Majewska-Sawka, A., &
Prakash Maran, J., Sivakumar, V., Thirugnanasambandham, K., & Sridhar, R. (2014). Rodríguez-García, M. I. (2013). Cellular localization and levels of pectins and
Microwave assisted extraction of pectin from waste Citrullus lanatus fruit arabinogalactan proteins in olive (Olea europaea L.) pistil tissues during
rinds. Carbohydrate Polymers, 101, 786–791. development: Implications for pollen–pistil interaction. Planta, 237(1),
Ptitchkina, N. M., Danilova, I. A., Doxastakis, G., Kasapis, S., & Morris, E. R. (1994). 305–319.
Pumpkin pectin: Gel formation at unusually low concentration. Carbohydrate Taylor, A. J. (1982). Intramolecular distribution of carboxyl groups in low methoxyl
Polymers, 23(4), 265–273. pectins — A review. Carbohydrate Polymers, 2(1), 9–17.
Racape, E., Thibault, J. F., Reitsma, J. C. E., & Pilnik, W. (1989). Properties of Thakur, B. R., Singh, R. K., Handa, A. K., & Rao, M. A. (1997). Chemistry and uses of
amidated pectins. II. Polyelectrolyte behavior and calcium binding of amidated pectin — A review. Critical Reviews in Food Science and Nutrition, 37(1), 47–73.
pectins and amidated pectic acids. Biopolymers, 28(8), 1435–1448. Thibault, J.-F., & Ralet, M.-C. (2003). Physico-chemical properties of pectins in the
Ralet, M.-C., Bonnin, E., & Thibault, J.-F. (2001). Chromatographic study of highly cell walls and after extraction. In F. Voragen, H. Schols, & R. Visser (Eds.),
methoxylated lime pectins deesterified by different pectin methyl-esterases. Advances in pectin and pectinase research (pp. 91–105). Netherlands, Dordrecht:
Journal of Chromatography B: Biomedical Sciences and Applications, 753(1), Springer.
157–166. Tsoga, A., Richardson, R. K., & Morris, E. R. (2004a). Role of cosolutes in gelation of
Rao, M. A., & Silva, J. A. L. d. (2006). Pectins. In Food polysaccharides and their high-methoxy pectin. Part 1. Comparison of sugars and polyols. Food
applications. pp. 353–411. CRC Press. Hydrocolloids, 18(6), 907–919.
Reitsma, J. C. E., Thibault, J. F., & Pilnik, W. (1986). Properties of amidated pectins. I. Tsoga, A., Richardson, R. K., & Morris, E. R. (2004b). Role of cosolutes in gelation of
Preparation and characterization of amidated pectins and amidated pectic high-methoxy pectin. Part 2. Anomalous behaviour of fructose: Calorimetric
acids. Food Hydrocolloids, 1(2), 121–127. evidence of site-binding. Food Hydrocolloids, 18(6), 921–932.
Renard, C. M. G. C., & Thibault, J.-F. (1996). Degradation of pectins in alkaline Vauquelin, M. (1790). Analyse du tamarin. Annales de Chimie, 5, 92–106.
conditions: Kinetics of demethylation. Carbohydrate Research, 286(0), 139–150. Ventura, I., Jammal, J., & Bianco-Peled, H. (2013). Insights into the nanostructure of
Renard, C. M. G. C., Crépeau, M.-J., & Thibault, J.-F. (1995). Structure of the low-methoxyl pectin–calcium gels. Carbohydrate Polymers, 97(2), 650–658.
repeating units in the rhamnogalacturonic backbone of apple, beet and citrus Venzon, S. S., Canteri, M. H. G., Granato, D., Junior, B. D., Maciel, G. M., Stafussa, A.
pectins. Carbohydrate Research, 275(1), 155–165. P., & Haminiuk, C. W. I. (2015). Physicochemical properties of modified citrus
Robertsen, B. (1986). Elicitors of the production of lignin-like compounds in pectins extracted from orange pomace. Journal of Food Science and Technology,
cucumber hypocotyls. Physiological and Molecular Plant Pathology, 28(1), 52(7), 4102–4112.
137–148. Vincken, J.-P., Schols, H. A., Oomen, R. J. F. J., McCann, M. C., Ulvskov, P., Voragen, A.
Rojas, J. A., Rosell, C. M., & Benedito de Barber, C. (1999). Pasting properties of G. J., & Visser, R. G. F. (2003). If homogalacturonan were a side chain of
different wheat flour-hydrocolloid systems. Food Hydrocolloids, 13(1), 27–33. rhamnogalacturonan I. implications for cell wall architecture. Plant Physiology,
Sadahira, M. S., Rodrigues, M. I., Akhtar, M., Murray, B. S., & Netto, F. M. (2016). 132(4), 1781–1789.
Effect of egg white protein-pectin electrostatic interactions in a high sugar Vithanage, C. R., Grimson, M. J., Wills, P. R., Harrison, P., & Smith, B. G. (2010).
content system on foaming and foam rheological properties. Food Rheological and structural properties of high-methoxyl esterified,
Hydrocolloids, 58, 1–10. low-methoxyl esterified and low-methoxyl amidatated pectin gels. Journal of
Santos, J. D. G., Espeleta, A. F., Branco, A., & de Assis, S. A. (2013). Aqueous Texture Studies, 41(6), 899–927.
extraction of pectin from sisal waste. Carbohydrate Polymers, 92(2), 1997–2001. Voragen, A. G. J., Schols, H. A., & Pilnik, W. (1986). Determination of the degree of
Sato, A. K., Oliveira, P., & Cunha, R. (2008). Rheology of mixed pectin solutions. Food methylation and acetylation of pectins by h.p.l.c. Food Hydrocolloids, 1(1),
Biophysics, 3(1), 100–109. 65–70.
Savary, B. J., & Nuñez, A. (2003). Gas chromatography–mass spectrometry method Voragen, A. J., Coenen, G.-J., Verhoef, R., & Schols, H. (2009). Pectin, a versatile
for determining the methanol and acetic acid contents of pectin using polysaccharide present in plant cell walls. Structural Chemistry, 20(2),
headspace solid-phase microextraction and stable isotope dilution. Journal of 263–275.
Chromatography A, 1017(1–2), 151–159. Vriesmann, L. C., & Petkowicz, C. L. O. (2013). Highly acetylated pectin from cacao
Savary, B. J., Hotchkiss, A. T., Fishman, M. L., Cameron, R. G., & Shatters, R. G. (2003). pod husks (Theobroma cacao L.) forms gel. Food Hydrocolloids, 33(1), 58–65.
Development of a valencia orange pectin methylesterase for generating novel
S.Y. Chan et al. / Carbohydrate Polymers 161 (2017) 118–139 139

Vriesmann, L. C., de Mello Castanho Amboni, R. D., & de Oliveira Petkowicz, C. L. Wu, D.-C., Loh, X. J., Wu, Y.-L., Lay, C. L., & Liu, Y. (2010). ‘Living’ controlled in situ
(2011). Cacao pod husks (Theobroma cacao L.): Composition and gelling systems: Thiol-disulfide exchange method toward tailor-made
hot-water-soluble pectins. Industrial Crops and Products, 34(1), biodegradable hydrogels. Journal of the American Chemical Society, 132(43),
1173–1181. 15140–15143.
Wai, W. W., Alkarkhi, A. F. M., & Easa, A. M. (2010). Effect of extraction conditions Xiao, C., & Anderson, C. (2013). Roles of pectin in biomass yield and processing for
on yield and degree of esterification of durian rind pectin: An experimental biofuels. Frontiers in Plant Science, 4(67).
design. Food and Bioproducts Processing, 88(2–3), 209–214. Yapo, B. M., Robert, C., Etienne, I., Wathelet, B., & Paquot, M. (2007). Effect of
Walkinshaw, M. D., & Arnott, S. (1981). Conformations and interactions of pectins: extraction conditions on the yield, purity and surface properties of sugar beet
II. Models for junction zones in pectinic acid and calcium pectate gels. Journal pulp pectin extracts. Food Chemistry, 100(4), 1356–1364.
of Molecular Biology, 153(4), 1075–1085. Yapo, B. M. (2009a). Biochemical characteristics and gelling capacity of pectin from
Walter, R. H., & Sherman, R. M. (1983). The induced stabilization of aqueous pectin yellow passion fruit rind as affected by acid extractant nature. Journal of
dispersions by ethanol. Journal of Food Science, 48(4), 1235–1237. Agricultural and Food Chemistry, 57(4), 1572–1578.
Wang, X., Lee, J., Wang, Y.-W., & Huang, Q. (2007). Composition and rheological Yapo, B. M. (2009b). Pectin quantity, composition and physicochemical behaviour
properties of ␤-lactoglobulin/pectin coacervates: Effects of salt concentration as influenced by the purification process. Food Research International, 42(8),
and initial protein/polysaccharide ratio. Biomacromolecules, 8(3), 992–997. 1197–1202.
Wang, X., Chen, Q., & Lü, X. (2014). Pectin extracted from apple pomace and citrus Yapo, B. M. (2011a). Pectic substances: From simple pectic polysaccharides to
peel by subcritical water. Food Hydrocolloids, 38, 129–137. complex pectins—A new hypothetical model. Carbohydrate Polymers, 86(2),
Wang, W., Ma, X., Xu, Y., Cao, Y., Jiang, Z., Ding, T., . . . & Liu, D. (2015). 373–385.
Ultrasound-assisted heating extraction of pectin from grapefruit peel: Yapo, B. M. (2011b). Pectin rhamnogalacturonan II: on the small stem with four
Optimization and comparison with the conventional method. Food Chemistry, branches; in the primary cell walls of plants. International Journal of
178, 106–114. Carbohydrate Chemistry, 2011 http://dx.doi.org/10.1155/2011/964521, 11
Wehr, J. B., Menzies, N. W., & Blamey, F. P. C. (2004). Alkali hydroxide-induced pages, article ID: 964521
gelation of pectin. Food Hydrocolloids, 18(3), 375–378. Zhong, H.-J., Williams, M. A. K., Goodall, D. M., & Hansen, M. E. (1998). Capillary
Wikiera, A., Mika, M., Starzyńska-Janiszewska, A., & Stodolak, B. (2015). electrophoresis studies of pectins. Carbohydrate Research, 308(1–2), 1–8.
Application of Celluclast 1.5L in apple pectin extraction. Carbohydrate Polymers, Zhu, R.-G., Sun, Y.-D., Li, T.-P., Chen, G., Peng, X., Duan, W.-B., . . . & Jin, X.-Y. (2015).
134, 251–257. Comparative effects of hawthorn (Crataegus pinnatifida Bunge) pectin and
Willats, W. T., McCartney, L., Mackie, W., & Knox, J. P. (2001). Pectin: cell biology pectin hydrolyzates on the cholesterol homeostasis of hamsters fed
and prospects for functional analysis. In N. C. Carpita, M. Campbell, & M. high-cholesterol diets. Chemico-Biological Interactions, 238, 42–47.
Tierney (Eds.), Plant cell walls (pp. 9–27). Netherlands: Springer. Zouambia, Y., Youcef Ettoumi, K., Krea, M., & Moulai-Mostefa, N. (2016). A new
Willats, W. T., McCartney, L., & Knox, J. P. (2003). Pectin cell biology: complexity in approach for pectin extraction: Electromagnetic induction heating. Arabian
context. In F. Voragen, H. Schols, & R. Visser (Eds.), Advances in Pectin and Journal of Chemistry, http://dx.doi.org/10.1016/j.arabjc.2014.11.011
Pectinase Research (pp. 147–157). Netherlands: Springer. Zulueta, A., Esteve, M. J., Frasquet, I., & Frígola, A. (2007). Vitamin C, vitamin A,
Willats, W. G. T., Knox, J. P., & Mikkelsen, J. D. (2006). Pectin: New insights into an phenolic compounds and total antioxidant capacity of new fruit juice and skim
old polymer are starting to gel. Trends in Food Science & Technology, 17(3), milk mixture beverages marketed in Spain. Food Chemistry, 103(4), 1365–1374.
97–104.

You might also like