You are on page 1of 14

Interpolymeric Complexes Formed Between Whey

Proteins and Biopolymers: Delivery Systems of


Bioactive Ingredients
Monique Barreto Santos, Naiara Rocha da Costa, and Edwin Elard Garcia-Rojas

Abstract: Whey proteins are obtained from dairy industry waste. Studies involving the analysis of the bioactive
compounds in whey show health benefits, as it is an excellent source of indispensable amino acids. Milk whey contains
principally β-lactoglobulin, α-lactoglobulin, bovine serum albumin, and lactoferrin, proteins with innumerable functional
and technological properties. One application of these proteins in food is the formation of interpolymer complexes,
along with other proteins or anionic polysaccharides. The formation of complexes occurs mainly through electrostatic
interactions between a negatively charged biopolymer and a positively charged biopolymer. This formation is influenced
by factors such as pH, ionic strength, and biopolymer ratio. Because they do not use high temperatures and chemical
reagents and have additional nutritional and functional value, these complexes have been used as encapsulating agents for
bioactive ingredients. Recent studies on their training and applications are addressed in this review to boost new research
and applications in the food industry, thus increasing opportunities for utilizing whey proteins.
Keywords: bovine serum albumin, coacervate complexes, electrostatic interactions, lactoglobulin, microencapsulation

Introduction The main advantage of the use of coacervate complex from


Whey protein (WP) is the result of whey utilization and is whey proteins compared to complexes from other proteins is their
obtained as a byproduct of the dairy industry, specifically cheese- high added value due to the diverse nutritional and functional
making, from the coagulation of casein. Until recently, whey was properties of whey proteins added to high productivity and low
massively discarded as effluent in rivers, causing significant en- cost because it results from the reuse of a byproduct (Almeida,
vironmental damage due to its high biological oxygen demand Conte-Júnior, Silva, & Alvares, 2013).
(Walstra, Wouters, & Geurts, 2006). More recently, whey has be- When the biopolymers have the same charge, the separation
come of great interest to the food industry after the discovery of occurs in two phases each one rich in a biopolymer, known as in-
the high biological value of the soluble protein fractions, which compatibility thermodynamic (segregation). However, when the
have begun to be used as a raw material in the elaboration of new biopolymers have opposite charges, the separation occurs in one
products with special characteristics and profit opportunity (Urista, phase rich, mainly of the biopolymers, and the other essentially
Álvarez, Riera, Cuenca, & Téllez, 2011). Among the most studied in the solvent called the thermodynamic compatibility (coacerva-
proteins found at the highest concentrations are β-lactoglobulin tion). In the last case, the intensity of formation can be influenced
(β-Lg), α-lactoglobulin (α-Lg), bovine serum albumin (BSA), and by intrinsic characteristics, such as molecular mass, conforma-
lactoferrin (LF; Sgarbieri, 2004). tion, and density of charges, and extrinsic characteristics, such
One alternative for the utilization of the whey proteins is in the as pH, ionic strength, the ratio of biopolymers, total concentra-
formation of interpolymeric complexes or coacervates through tion, temperature, and pressure (de Kruif, Weinbreck, & de Vries,
interactions with other natural polymers (biopolymers), such as 2004; Tolstoguzov, 1991; Turgeon et al., 2007; Turgeon, Beaulieu,
polysaccharides and proteins, which are then used as food ingre- Schmitt, & Sanchez, 2003).
dients, encapsulating agents for bioactive compounds, texturing, Several studies have proposed elucidating the formation charac-
and stabilizing properties (Turgeon, Schmitt, & Sanchez, 2007). teristics of complexes between proteins and polysaccharides from
the standpoints of thermodynamics (de Kruif at al., 2004; Tur-
geon et al., 2007), kinetics of formation (Schmitt & Turgeon,
CRF3-2017-0241 Submitted 12/7/2017, Accepted 3/9/2018. Authors Santos, da 2011), and the functional properties of complexes (Zhang, Zhong,
Costa, and Garcia-Rojas are with Programa de Pós-graduação em Ciência e Tecnologia & Vardhanabhuti, 2014; Norton, Espinosa, Watson, Spyropoulos,
de Alimentos (PPGCTA), Univ. Federal Rural de Rio de Janeiro (UFRRJ), Rodovia & Norton, 2015), as well as in applications as an encapsulating
BR 465, Km 7, Seropédica/RJ, 23890-000, Brazil. Author Garcia-Rojas is also agent (Weinbreck, Kruif, & Schrooyen, 2003a; Mellema, 2004;
with Laboratório de Engenharia e Tecnologia Agroindustrial (LETA), Univ. Federal
Fluminense (UFF), Av. dos Trabalhadores, 420, Volta Redonda/RJ, 27255-125,
Zhang & Mutilangi, 2013).
Brazil. Direct inquiries to author Garcia-Rojas (E-mail: edwin@vm.uff.br). Due to the growing importance given to whey proteins added
to the great interest expressed by the numerous studies found in


C 2018 Institute of Food Technologists®

doi: 10.1111/1541-4337.12350 Vol. 0, 2018 r Comprehensive Reviews in Food Science and Food Safety 1
Complexes between WPI and biopolymers . . .

the literature about delivery systems of bioactive ingredients in Proteins of animal origin are important in the human diet, as
pharmaceutical, food, and cosmetic applications, this review pro- they are the main source of nitrogen and essential amino acids
poses to provide an overview of the literature published over the (Lowery, Edel, & Mcbride, 2012). The nutritional quality of a
last 15 years of the use of whey proteins as an encapsulating agent protein will depend on numerous factors, such as composition,
of different ingredients. It will first address the characteristics of digestibility, absorption, bioavailability of essential amino acids,
the main whey proteins currently used and the main mechanisms, and total nitrogen. Digestibility is the first factor that reflects the
parameters, and interactions involved during the formation of the efficiency of protein utilization and is considered a condition of
complexes. Then we present the applicability of whey proteins in quality (Lemon, Berardi, & Noreen, 2002; Tang, Moore, Kujbida,
the formation of complexes between biopolymers and on bioac- Tarnopolsky, & Phillips, 2009).
tive ingredient delivery systems such as oils, vitamins, pigments, Quantitatively, the principal whey proteins account for approx-
proteins, and probiotics. imately 20% of total bovine milk proteins. They consist primarily
of β-Lg, α-La, BSA, immunoglobulins (Ig), LF, lactoperoxidase,
Whey glycomacropeptides, and proteose-peptone (Almeida et al., 2013;
Whey, also called dairy serum or serum of cheese, is a yellow- Etzel, 2004). They have excellent amino acid composition and
green colored liquid obtained from the coagulation of milk in- high digestibility and bioavailability of essential amino acids, espe-
tended for the manufacture of cheeses, casein, or similar products cially branched-chain amino acids (BCAA, leucine, isoleucine, and
(Codex Alimentarius, 2011). It is a byproduct of significant rele- valine), in addition to having good technological properties such
vance in the dairy industry due to the high volume produced and as solubility (Pelegrine & Gasparetto, 2003), emulsification (Gi-
its nutritional composition (Walstra et al., 2006). rard, Turgeon, & Paquin, 2002a), gelatinization (Ngarize, Adams,
The composition of fresh whey is 94.2% water, 0.8% whey pro- & Howell, 2005), and water retention capacity (Vidigal et al.,
teins, 4.3% lactose, 0.5% minerals, and 0.1% fat (Almeida et al., 2012). In addition, they have antimicrobial and antioxidant func-
2013). As with milk, whey composition may vary with diet, race, tions (Balcão et al., 2013). Due to these factors, whey protein
individual, stage of lactation, management, climate, season, envi- has been used in bakery products, salad dressings, infant formulas,
ronmental condition, and animal health (Park, Juárez, Ramos, & and medical nutritional formulas (Jovanovi, Bara, & Ma, 2005).
Haenlein, 2007). The composition and type of industrially pro- Among the most studied proteins are β-Lg, α-Lg, BSA, and LF
duced whey also depends on the type of coagulation to which the (Sgarbieri, 2004).
casein is subjected. Sweet whey is obtained via enzymatic coag- β-Lg is higher in whey (60%) and has a higher BCAA content,
ulation of casein, and acid whey is obtained through coagulation with a total of 162 amino acid residues. It has a molecular mass of
with lactic acid bacteria or the addition of acid to milk (Etzel, 18.4 kDa and an isoelectric point (pI) of 5.3 (Farrell et al., 2004;
2004; Walstra et al., 2006). Guimont, Marchall, Girardet, & Linden, 1997). β-LG belongs
Traditionally, whey was considered by cheese producers as a to the lipocalin family, which is characterized by a high affinity
byproduct of manufacturing with little or no commercial value. for hydrophobic ligands that leaves them with highly allergenic
For a long time, it was used in animal feed or discarded directly potential (Stojadinovic et al., 2012). Eleven β-LG variants have
in the effluent (Walzen, Dillard, & German, 2002). However, been identified, which include A, B, C, D, E, F, G, H, I, J, and W; A
it has become an important raw material for the production and B occur most frequently in most bovine breeds (Caroli, Chessa,
of value-added products due to the discovery of functional and & Erhardt, 2009). In addition, the protein appears as a dimer but
bioactive properties of its protein components (Marshall, 2004; dissociates into monomers when denatured (Kontopidis, Holt,
Urista et al., 2011). Advances in processing technology have re- & Sawyer, 2004). It is recognized for its high technological and
sulted in the development of various whey protein presentation functional value; however, because it is the most abundant protein
forms (Voswinkel & Kulozik, 2014). This variation is related to and the most allergenic and antigenic, it may present health risks
the protein content present in the product. Thus, whey protein to patients allergic to cow milk proteins, especially children. This
concentrate (WPC) varies between 25% and 80% in protein con- is because its globular structure is stable against acid hydrolysis and
tent. However, whey protein isolate (WPI) is approximately 90% proteolytic enzymes present in the stomach, and thus it remains
protein. Whey protein hydrolyzate (WPH) is the whey isolated intact after digestion (Sgarbieri, 2004; Stojadinovic et al., 2012).
after enzymatic hydrolysis and consists of broken down proteins to The α-Lg protein represents 20% of total whey protein, con-
produce even smaller molecules that are more easily digested and taining 123 amino acid residues and having the highest tryptophan
absorbed into the bloodstream (Carunchia, Croissant, & Drake, content (6%) among all dietary protein sources. It is also rich in
2005). lysine, leucine, threonine, and cysteine. Its molar mass is 14.2 kDa
and it has a pI of 4.8 (Kinsella & Whitehead 1989; Farrell et al.,
2004). To date, three variants (A, B, and C) have been identified
Whey Proteins (Caroli et al., 2009). Its biological properties include anticancer ac-
In general, proteins are natural polymers (biopolymers) that, tivity and antimicrobial activity against pathogenic bacteria, such
due to the chemical properties of their basic building units, as Escherichia coli, Staphylococcus aureus, and Klebsiella pneumonia,
share properties common to amphiphilic and polyampholyte in addition to being rich in tryptophan, a precursor of niacin
molecules. They are amphiphilic because they have both a hy- and serotonin (Haraguchi, Abreu, & Paula, 2006; Santos, Teix-
drophobic and a hydrophilic region and polyampholytes because eira, & Rodrigues, 2011). Moreover, α-Lg is a metallo-protein
they contain aliphatic and carboxylic acids that may behave as a with a calcium atom that enables it to bind to other proteins.
polyacid or polyanion. These characteristics allow proteins to in- Among the whey proteins, α-Lg is the most thermally stable,
teract with molecules with different properties (hydrophobic, hy- because calcium promotes the formation of intermolecular ionic
drophilic, and/or positively and negatively charged), depending on bonds, which makes the molecule resistant to thermal unfolding
the physicochemical conditions (Croguennec, Tavares, & Bouhal- (Yada, 2004). Commercially, α-Lg has been widely used in in-
lab, 2017; Mezzenga & Fischer, 2013). fant formulas due to its structural and conformational similarity to

2 Comprehensive Reviews in Food Science and Food Safety r Vol. 0, 2018 


C 2018 Institute of Food Technologists®
Complexes between WPI and biopolymers . . .

breast milk proteins, and it is also used in sports nutrition foods due promoted the interaction between the charged molecules and
to its branched amino-acid content (Walzem, Dillard, & German, the entropic effects that tended to disperse the interaction. This
2002). theory was based on the following assumptions: (a) molecules
BSA corresponds to approximately 3% of whey proteins, has 583 have a random chain configuration, (b) solvent–solute in-
amino acid residues, and is rich in cysteine (6%) which has affinity teractions are less significant, (c) interacting forces are dis-
for free fatty acids and other lipids, and favors its transport in the tributed according to the nature orientation of the biopoly-
bloodstream. It also performs many other functions, such as elim- mer, and (d) there is no specific local interaction between
ination of oxygen-free radicals and deactivation of various toxic molecules.
lipophilic metabolites such as bilirubin; it has a molecular mass of Later, Tainaka (1980) developed a model to understand the for-
66.4 kDa and pI 4.8 (Emerson, 1989; Kinsella & Whitehead 1989; mation of complex coacervation that is more comprehensive than
Farrell et al., 2004; Neemann, Rosenberger, Jefferson, & Mcadam, all previous theories and is still accepted today. It is applicable
2013). Among the whey proteins, BSA is the most widely studied, to a great number of systems containing a high and low density
particularly because of its structural similarity to human serum al- of charges. According to Tainaka, the driving forces for phase
bumin. The BSA molecule is composed of a chain organized separation are electrostatics and the force of attraction between
into three homologous helical domains, which have two subdo- aggregates, which become stronger when the charge density and
mains that share a common helical form joined by peptide bonds. the molar mass of the polymers increase. However, both the in-
They are characterized by a small number of tryptophan residues crease in molar mass and density of charges have a maximum and a
and a high number of cysteines. BSA has only one free cysteine minimum limit that prevent coacervate formation (Tainaka, 1979,
residue, and the remaining residues form 17 disulfide bonds that 1980).
help maintain the tertiary structure (Dong, Chen, & Hu, 2007; In general, proteins and polysaccharides, when added to aque-
Sinha, Radha, Prakash, & Kaul, 2007). X-ray crystallography data ous solution, may present three thermodynamic phenomena: co-
showed that the structure of albumin is predominantly composed solubility, thermodynamic incompatibility, and complex coacerva-
of α-helix (67%; Bujacz, 2012; Damodaran & Fennema, 2010). tion (de Kruif & Tuinier 2001; Tolstoguzov, 1991). Co-solubility
BSA can also form dimers, especially at high concentrations or in occurs when intermolecular interactions are absent; this occurs
its crystallized form (Carter & Ho, 1994). when entropy is prevailing, and in this case, a stable and homoge-
LF accounts for approximately 1% to 2% of total whey pro- neous solution results. The total concentration of biopolymers in
teins, contains approximately 680 amino acid residues, and has a this situation is below the critical limit (3% to 4%; Boland, Singh,
molar mass of 76.1 kDa, and a pI of 8.9 (Farrell et al., 2004; Mar- & Thompson, 2014; de Kruif & Tuinier, 2001). Thermodynamic
shall, 2004). It transports iron ions (Fe2+ and Fe3+ ), in addition to incompatibility occurs when the ionic strength is high or when
other metals such as Cu2+ , Zn2+ , and Mn2+ (González-Chavez, the pH of the medium is above the protein pI (Benichou, Aserin,
Arévalo-Gallegos, & Rascón-Cruz, 2009). Due to its iron bind- & Garti, 2002; Boland et al., 2014). Thus, solvent–protein interac-
ing role, LF can be found in the form of apo-lactoferrin (apo-LF) tions are stronger than protein-polysaccharides, because proteins
when it is less than 5% iron-saturated, as opposed to the holo- and polysaccharides carry negative charges (Doublier, Garnier,
lactoferrin (holo-LF) form when it has higher saturation (Steijns Renard, & Sanchez, 2000). In the case of interactions between
& Hooijdonk, 2000). Because of its high affinity to the ion, it is the proteins, thermodynamic incompatibility occurs when the pH of
only protein capable of binding to the metal at various pH values the medium is not between the pI of the proteins. At this point,
and in a reversible form (Wally & Buchanan, 2007). Its molecular the two proteins have the same charge and, therefore, interactions
structure consists of a simple polypeptide chain with two highly with the solvent are stronger (Santos, Carvalho, & Garcia-Rojas,
homologous terminal globular (C and N) lobes. These are linked 2018).
together by parts of an α-helix, with each lobe containing a bind- Coacervation is defined by the International Union of Pure
ing site (Öztaş Yeşim, & Ozguneş, 2005). Each lobe consists of two and Applied Chemistry (IUPAC; 1997) as a colloidal separa-
sublobes called N1, N2, C1, and C2 (Steijns & Hooijdonk, 2000). tion of systems in two liquid phases and occurs when biopoly-
It is considered a multifunctional protein because it plays impor- mers exert strong electrostatic attraction to one another. These
tant roles in biological systems, such as absorption of intestinal iron molecules form complexes of three-dimensional structures during
during breastfeeding of newborns; it also has antibacterial, antifun- charge neutralization. Highly hydrated and concentrated droplets
gal, antiviral, anti-inflammatory, and immunoregulatory activities in biopolymers tend to coalesce to minimize energy. The phase
and performs antioxidant actions (Balcão et al., 2013). separation at the macroscopic level occurs in an electrically neutral
phase concentrated in the two biopolymers (coacervate) and an-
Formation of inter-polymer complexes other phase essentially composed of solvent. This phenomenon is
Phase separation was first observed by Tiebackx (1911) in a called complex coacervation (Bungenberg de Jong, 1949; Burgess,
study of gelatin and gum acacia (GA). In 1929, Bungenberg de 1990; de Kruif et al., 2004; Phillips & Williams, 2009). Complex
Jong and Kruyt proposed the first theoretical explanation for the coacervation differs because it generally occurs at low concentra-
phenomenon of simple coacervation. When studying complex tions of biopolymers (<3 to 4 wt% total solids) and low ionic
coacervation formation between gelatin/GA in 1949, Bungenberg strength (<400 mM), and when both molecules carry net oppo-
de Jong realized that the interaction occurred due to changes in the site electric charges (Tolstoguzov, 2007). Table 1 shows the most
electrostatic charges around the biopolymers (GA with negative recent work reported in the literature on the formation of inter-
charge and gelatin with positive charge). It was also determined polymeric complexes formed from whey proteins and biopoly-
that these changes in charge were promoted by changes in pH and mers, the main analyzes performed by the authors (zeta potential,
NaCl (Bungenberg de Jong, 1949; Bungenberg de Jong & Kruyt, turbidimetry, particle size, microscopy, thermal analysis, and rheol-
1929). ogy) and the most relevant parameters observed (Influence of pH,
Overbeek and Voom (1957) reported that coacervation oc- molecular mass, salt, and ratio of biopolymers) on the formation of
curred via competition between the electrostatic forces. This complexes.


C 2018 Institute of Food Technologists® Vol. 0, 2018 r Comprehensive Reviews in Food Science and Food Safety 3
Complexes between WPI and biopolymers . . .

Table 1–Systems composed of whey proteins and biopolymers, the main analyses performed and the most relevant parameters.

WP Biopolymer Parameters Tests Reference


β-Lg Sodium alginate pH ITC, DLS, τ , ζ , Harnsilawat, Pongsawatmanit,
and McClements (2006)
β-Lg Pectin pH, CNaCl τ , AFM Wang, Wang, Ruengruglikit, and
Huang (2007)
BSA Chitosan pH, CNaCl DLS, MO, SANS, USANS, τ , Kayitmazer, Strand, Tribet,
G’, and G’’ Jaeger, and Dubin (2007)
β-Lg Lysozyme T, r MO, ITC, HPLC, τ , DLS Nigem et al. (2007)
BSA Hyaluronic acid pH, CNaCl , r DLS, τ Lenormand, Deschrevel,
Tranchepain, and Vincent
(2008)
β-Lg/α−Lg Chitosan pH, T τ , SDS–PAGE, DSC, SEM Lee and Hong (2009)
WPI Chitosan pH, r η, G’, and G’’, ITC, τ , Bastos et al. (2010)
β-Lg GA T, r DLS, μ, τ , MO, ITC Aberkane et al. (2010)
LF Pectin pH, T τ , DLS, DSC Bengoechea et al. (2011)
β-Lg k-carrageenan pH, r SEM, AFM, μ, SANS Jones et al. (2011)
β-Lg GA r, T ITC, DLS, μ, τ , MO, FTIR Aberkane et al. (2012)
BSA Beet pectin pH, r ζ , DLS, τ , SANS, GPC Li et al. (2012)
BSA Pectin r, CNaCl, pH τ , ζ , G’, and G’’ Ru, Wang, Lee, Ding, and Huang
(2012)
β-Lg LF pH, CNaCl , Ct, r ζ , DLS, SEC, Yan et al. (2013)
BSA Sodium alginate pH, r, Ct τ , ζ , TMP, MO Neemman et al. (2013)
BSA k-carrageenan CNaCl , pH, r τ , G’, and G’’∗∗∗ Chai et al. (2014)
LF GA CNaCl , CT , pH, r, T τ , ζ , XRD, AFM Gulão et al. (2014)
β-Lg LF pH, r, CNaCl τ, ζ Anema and de Kruif (2014)
BSA/ β-Lg Hyaluronic acid pH, CNaCl τ , ITC, SEC, CLSM Du, Dubin, Hoagland, and Sun
(2014)
β-Lg Pectin pH, r, kind of salt, CNaCl τ , ζ , scattering intensity Hirt and Jones (2014)
WPI Arabinogalactan-protein CNaCl , CT , pH, r, τ , ζ , DLS, MO, G’, and G’’ Wee et al. (2014)
Lactalbumin Xanthan gum and pectin pH, CT , CNaCl τ , MEV Zuvanov, Garcia-Rojas, Souza, de
Gulão, and Pereira (2014)
BSA k-carrageenan pH, r, CNaCl τ , MEV, G’, and G’’, XRD Souza et al. (2015)
WPI GA pH, r SDS-PAGE, τ , ELISA Ach et al. (2015)
BSA Linseed flour pH, r, CNaCl τ , ζ , DLS, μ Liu et al. (2015)
WPI Pectin pH, CNaCl MO, DLS, τ , δ Thongkaew et al. (2015)
β-Lg Pectin r ITC, CD, IFS Xu, Melton, Jameson, Williams,
and McGillivray (2015)
LF Sodium alginate pH, r FTIR, CD, μ, ζ , DSC Bokkhim, Bansal, Grøndahl, and
Bhandarim (2015)
β-Lg LF pH, r τ , DLS, ITC Flanagan et al., 2015
β-Lg Lysozyme pH, r ζ , DLS, τ , TEM, SEM, HPLC Diarrassouba et al. (2015)
β-Lg LF pH, r, Ct, CNaCl τ , ITC, MO Tavares et al. (2015)
α-Lg Carboxymethyl-dextran pH μ, ζ , τ Du, Reuhs,and Jones Du, Reuhs,
and Jones, 2016)
β-Lg Gum of Persia pH, r, Ct, tipo de sal, T, ITC Hadian et al. (2016)
força iônica, T
BSA/ β-Lg Gelatine pH, r, T SANS, G’ and G’’ Pathak, Rawat, Aswal, and
Bohidar (2016)
LF Casein pH, r, ionic strength τ , ζ , DLS Anema and de Kruif, (2016)
LF Pea protein pH, r τ , ζ , DLS, SAXS, AFM Adal et al. (2017)
WPI Tamarind seed mucilage pH, r ζ , τ , FTIR, XRD, MO, SEM, González-Martı́nez et al. (2017)
DSC, TGA
LF Sodium alginate pH, r τ , ζ , yield, SEM, digestion Wang, Blanch, Barro, and
in vitro, SDS-PAGE, CD, Adhikari (2017)
DPPH,
WPI Flaxseed pH, r τ , ζ , G’, and G’’, MO Liu, Shim, Shen, Wang, and
Reaney (2017)
WPI Pectin pH, T τ , ζ , DLS Zeeb, Mi-Yeon, Gibis, and Weiss
(2018)
BSA Ovalbumin pH, r, CNaCl τ , ζ , DLS, SEM, MO, DSC, Santos et al. (2018)
FTIR

Interactions between whey proteins and biopolymers protein–polysaccharide complexes are induced electrostatically it
In thermodynamically compatible systems, whey proteins and is assumed that a stabilization should be achieved through sec-
polysaccharides have opposite charges, leading to electrostatic. ondary hydrogen bonds or hydrophobic interactions (Doublier
However, in some systems, depending on the biopolymer used, et al., 2000).
other interactions such as hydrogen bond van der Waals, or hy- The hydrophobic interactions are important for the biopolymers
drophobic interactions may occur (Boland et al., 2014; Doublier that have nonpolar groups in aqueous solutions and are manifested
et al., 2000). Complexation is favored thermodynamically by the by the association of the nonpolar groups (McClements, 2006).
enthalpy gain from the formed electrostatic interactions. These Proteins have hydrophobic groups, which when exposed due to
interactions are stronger when the pH of the system is below structural modifications can interact in a nonpolar manner. The
the pI of the protein, especially when there is low ionic strength increase in temperature or addiction of the small amounts of salt,
(Schmitt, Sanchez, Desobry-Banon, & Hardy, 1998). When the

4 Comprehensive Reviews in Food Science and Food Safety r Vol. 0, 2018 


C 2018 Institute of Food Technologists®
Complexes between WPI and biopolymers . . .

for example, contributes to this type of interaction (Boland et al., Parameters that influence the formation of complexes
2014). between whey proteins and biopolymers
By characterizing the interaction between β-Lg and GA, Numerous studies have highlighted both the conditions of the
Aberkane et al. (2012) verified two steps of interaction using medium, such as pH and ionic strength, and the physicochemical
isothermal titration calorimetry (ITC) at pH 4.5 and 25 °C. The characteristics of the polymers, charge density, and molar mass, as
authors suggest that the first step would be related to electrostatic parameters capable of controlling the coacervation process (Wein-
and hydrogen interactions for having been enthalpically driven, breck, de Vries, Schrooyen, & de Kruif, 2003b).
and that the second step would have been driven by hydropho-
bic contributions related to conformational changes due to the pH
favorable binding entropy. In the studies of Gulão, Souza, da Silva, The interactions between whey proteins and polysaccharides
Coimbra, and Garcia-Rojas (2014) aggregates were formed be- are directly controlled by the degree of ionization of the biopoly-
tween LF and GA at 80 °C, possibly indicating the participation mers, and thus the pH of the medium is important in electrostatic
of hydrophobic interactions by exposing the hydrophobic groups control (Turgeon et al., 2007). For each pair of cationic/anionic
of the protein after a conformational change promoted by the biopolymers, there is a pH at which the charge density will be
increase in temperature. maximal and equivalent and, consequently, the intensity of the
Hydrogen bonds may occur between polysaccharide ester electrostatic interactions and formation of coacervates will also be
groups and the hydroxyl, amide and carboxylic groups of the maximal (Phillips & Williams 2009). These interactions generally
proteins, being negatively influenced by the presence of urea and occur at a pH below the pI of the protein and above the pKa of
high temperatures occurring generally at pHs above the pI of the the polysaccharide (Doublier et al., 2000).
protein where the density of charges is lower (Girard, Turgeon, & The complex formation mechanism can be evidenced by tur-
Gouthier, 2002b). bidimetric techniques, zeta potential, and particle size (Weinbreck
The van der Waals forces, however, are extremely weak elec- et al., 2003b). From the turbidimetric titration of the basic pH
trical attractions, because they promote temporary interactions of to the acidic pH of a system composed of protein and polysac-
dipoles. However, they may influence macromolecule interactions charide, the following stages of formation can be evidenced:
in addition to electrostatic forces (Damodaran & Fennema 2010). (a) pH > pI, protein and polysaccharide have a high negative
Liu, Shim, Wang, and Reaney (2015) studied the influence of charge density that promotes the repulsion between them; (b) the
urea on the formation of complexes between BSA and flax gum, pH is reduced to a critical value (pHc ), and there is a slight increase
using turbidimetric analysis data to verify the participation of hy- in turbidity due to the weak association between the polymers,
drogen bonds in the process. The authors observed a reduction forming soluble complexes; (c) with further acidification of the
in turbidity values and concluded that the electrostatic interaction system, the turbidity increases abruptly due to the association of
between the biopolymers primarily directed the initial stage of the protein and polysaccharide forming the insoluble or coacer-
formation of the complexes. Afterwards, hydrogen bonds served vate complexes, denoted as pHθ 1 ; (d) when the pH is reduced to
to stabilize them. Tavares, Croguennec, Hamon, Carvalho, and values close to the polysaccharide pKa , its anionic groups lose neg-
Bouhallab (2015) characterized the interaction between β-Lg and ative charge density, weakening the interactions and dissociation
LF using ITC, and they verified that the interactions were exother- of the complex, denoted as pHθ 2 .
mic. This indicates the predominance of electrostatic interactions, The turbidity variation through acid titration in a system con-
as they are generally encountered during protein complexation taining BSA and k-carrageenan was studied. As the pH decreases,
and in oppositely charged polyelectrolyte systems. In a study that there is a decrease in electrostatic repulsion and a slight increase in
evaluated the effect of temperature on complexes between β-Lg the turbidity of the system due to the formation of soluble com-
and Persian gum, Hadian et al. (2016) demonstrated that elec- plexes (pHc ). This is followed by a remarkable increase in turbidity
trostatic interactions, hydrogen bonding, and hydrophobic inter- due to the increase in the size and number of complexes formed
actions were involved in the complexation process. The authors between BSA and k-carrageenan, which occurs up to pHmax , with
also concluded that the enthalpy of binding obtained by ITC was the maximum formation of coacervate complexes (Chai, Lee, &
independent of the temperature in the studied range (25, 40, and Huang, 2014). In the pH range above the pI of the protein, the
55 °C). In addition, they found that the temperature increase from biopolymers begin to carry similar charges of sufficient magnitude
25 to 40 °C increases the significance of hydrophobic interactions to favor the separation of phases, and thermodynamic incompat-
in complex formation. The hydrogen bonds are also weakened, ibility occurs. This behavior was reported by Thongkaew, Hin-
giving rise to hydrophobic interactions and exposing nonpolar richs, Gibis, and Weiss (2015) when studying the charge density
groups. of whey protein and pectin as a function of pH. At pH 3.2, the
Fourier transform infrared spectroscopy (FTIR) can be used to pectin presents negative charges, whereas the protein is positively
verify the participation of hydrogen bonds. González-Martı́nez charged, favoring an interaction between them. However, the val-
et al. (2017) characterized coacervate complexes formed between ues obtained at pH 6.5 for the zeta potential for both are negative,
WPI and tamarind gum. They observed a shift in the band rep- confirming strong repulsion between the two polymers.
resenting the stretching of N–H and O–H groups, indicating the The pKa of polysaccharides can determine the limit of com-
presence of hydrogen bonds, and therefore concluded that the hy- plex formation. According to Bengoechea, Jones, Guerrero, and
drogen bonds represent a significant force in the stability of the McClements (2011), sediments formed during titration of LF
protein–polysaccharide complexes. Likewise, Santos, de Carvalho, and pectin are solubilized at pH 2.0. From this pH, the pectin
and Garcia-Rojas (2018) verified the participation of electrostatic molecules lose most of their negative charge, and the electrostatic
interactions in the reduction of bands related to amides and the interactions are weaker. In the acid titration between LF and GA,
participation of hydrogen bonds in the reduction of bands repre- Gulão, de Souza, da Silva, Coimbra, and Garcia-Roja (2014) re-
senting the stretching of N–H and O–H groups in their study of ported a reduction of turbidity from pH 3.0 and total solubilization
heteroprotein complexes between BSA and lysozyme. of the sediments at pH 2.0, called pHθ 2 .


C 2018 Institute of Food Technologists® Vol. 0, 2018 r Comprehensive Reviews in Food Science and Food Safety 5
Complexes between WPI and biopolymers . . .

Ionic strength
The coacervation process is influenced by the addition of salt
and is generally verified by the addition of NaCl. Low concentra-
tions may favor the formation of complexes to a critical level where
attractive forces between polymers are diminished (Li et al., 2012;
Weinbreck et al., 2003b). Under these conditions, salt favors ac-
cess to polymer charges, as it “unwinds” the molecule, facilitating
electrostatic attraction (Weinbreck et al., 2003b). However, at high
salt concentrations the competition created by the Na+ ions by
the negative sites of the polysaccharide molecule, and the Cl− ions
by the positively charged sites of the protein, shield the biopoly-
mer charges, decreasing the interaction between them (Schmitt,
Aberkane, & Sanchez, 2009). The ionic force acts by covering
the coulombic forces between the oppositely charged biopoly-
Figure 1–Graphical representation of the turbidity variation as a function mers which can completely suppress the complexation by having
of pH of a system composed of BSA and lysozyme. The turbidity value has an effect on the Debye length (Dubin, Gao, & Mattison, 1994).
a small increase from pHc followed by a sharp rise in pHθ1 , to maximum As reported by previous work (Gulão et al., 2014; Gurov,
value (pHmax ) with subsequent reduction in pHθ2 (Adapted Santos et al., Gurova, Leontiev, and Tolstoguzov, 1988) also found that low salt
2018).
concentrations (100 to 250 mM NaCl) favored interactions be-
tween lactoferrin and gum arabic by observing increased turbidity
with increased particle size; 250 mM was the largest diameter. In
Unlike the authors cited above, Liu et al. (2015) reported a pHc contrast, a decrease in particle diameter correlated to the turbidity
above the protein pI for the BSA solution and flaxseed gum at data, indicating its negative influence on the interaction at higher
the ratio of 1:1. The formation of soluble complexes occurred salt concentrations (300 and 500 mM NaCl). Similar results were
due to the interaction of a few cationic groups of BSA with the observed by Laos, Brownsey, and Ring (2007) when studying the
anionic groups of the gum. Following acidification, the pHθ 1 complexation between furcellaran gum and BSA and/or β−Lg.
found was 4.8, just below the pI of the BSA (4.9). In the same They reported that the best salt concentration for the process was
study, the authors discussed the results obtained using dynamic 30 mM. Liu et al. (2015) found that the addition of salt (NaCl)
light scattering (DLS), which determines particle size, and they in the system containing BSA and flaxseed gum not only reduces
verified the same sizes found for solutions of biopolymers in the the formation of the complex to half in the concentration of 50
mixture at pH above pHc , suggesting no formation of complexes. mM but also alters the values of pHc and pHθ 1 . In this study,
However, at pH 5.0 > pHθ 1 , the observed particle sizes suggested they observed a decrease in pHc values from 5.4 to 4.8 and from
the formation of soluble complexes, and at pH 4.0 < pHθ 1, particle pHθ 1 from 5.0 to 4.4, leading to a decrease in the number and
sizes matched the formation of insoluble complexes. stoichiometry of coacervate complexes formed.
Similar to the formation of complexes between proteins and In studies of complexation between β-lactoglobulin and pectin
polysaccharides, the formation of heteroprotein complexes is dis- proposed by Sperber, Stuart, Schols, Voragen, and Norde (2010)
tinguished by the formation of complexes at pH values between the results regarding the influence of ionic strength are in agree-
the pI of the proteins. In this case, one protein behaves as a neg- ment with those previously mentioned. The authors observed that
atively charged anionic biopolymer (below its pI), and the other the increase in ionic strength reduced the enthalpy contribution
behaves as a positively charged cationic biopolymer (above its pI; of the interaction, concluding that the addition of NaCl inhibits
Santos et al., 2018). The effect of pH on heteroprotein complex the electrostatic interactions fundamental to complexation.
formation between β-Lg and lysozyme was studied by Diarras- Hadian et al. (2016) studied the influence of concentration (0 to
souba et al. (2015) through turbidimetric analysis and zeta poten- 100 mM) and the type of ions (Na+1 and Ca+2 ) on the formation
tial. In this study, turbidity was observed due to the formation of of complexes between β-Lg and the water-soluble fraction of
complexes over a wide pH range (6.8 to 10.0). At pH 5.0 and Persian gum (WPG). Salt concentrations of 50 mM (NaCl) and 15
11.0, near the pI of β-Lg (5.3) and lysozyme (10.7), respectively, mM (CaCl2 ) were sufficient to promote a significant reduction of
this increase was not verified. These results are justified because coacervation. In addition, the authors concluded that the effects
at pH 6.8, 7.5, and 8.0, the proteins carry negative (β-Lg) and of divalent ions (Ca+2 ) were higher than those of monovalent
positive (lysozyme) charges, which allow their interaction. Ana- ions (Na+1 ) by suppressing complex coacervation at lower ionic
lyzing zeta potential data, the authors observed that the pH with strength values. According to Schmitt et al. (2009), this behavior
highest turbidity coincided with the lowest zeta potential value. is due to the double gain in entropy after the release of two
This is because there is greater interaction and equilibrium of the monovalent ions compared to a single divalent ion. Moreover, the
charges, as the supercritical charge is closer to 0 mV. addition of mono- and divalent ions shift critical values of pH
A similar study was performed by Santos et al. (2018) using BSA to more acidic values to compensate for the partial blockade of
and lysozyme. The complexes formed in the pH range between charges induced by the additional ions; that is, the charge density
8.0 and 11.0, presenting maximum complexion at pH 9.0. Figure 1 of the proteins must increase to form the same amount of soluble
shows the data obtained by the turbidimetric analysis at a given complexes or comparable neutralization of charges (Hadian et al.,
ratio. In this Figure, the three sequential formation processes are 2016; Schmitt et al., 2009).
represented: pHc , representing the beginning of the formation of In addition to the pH shift’s relationship to complex formation,
the soluble complexes; pHƟ1 , which determines the formation the Debye length (Rd), a scale associated with the length of ionized
of the insoluble complexes; and pHƟ2 , which determines the end groups on the surface of the molecule, can be related to changes
of the process, indicating the co-solubility of the biopolymers. in pHc and pHθ (Orme & Giocondi, 2007). Chai et al. (2014)

6 Comprehensive Reviews in Food Science and Food Safety r Vol. 0, 2018 


C 2018 Institute of Food Technologists®
Complexes between WPI and biopolymers . . .

associated this parameter with distance from the protein to the Santos et al. (2018) found higher turbidity in a 1:2 ratio of BSA:
polysaccharide, and the electrostatic attraction between the protein 1ysozyme, and BSA (66 kDa) is a larger protein than lysozyme.
and the polysaccharide is suppressed when the distance between
them is greater than the Debye length. The addition of salt to the Temperature
BSA and k-carrageenan system suppressed electrostatic interactions Temperature may alter the conformation of proteins and
between them, thus displacing pHc , pHθ , and pHmax . polysaccharides, favoring nonelectrostatic interactions. At low
The salt effect was also analyzed for heteroprotein complexes. temperatures, hydrogen bonds are favored, whereas at higher tem-
The influence of NaCl on complex formation between β-Lg and peratures hydrophobic interactions take precedent due to the ex-
LF was reported by some authors. Yan et al. (2013) and Tavares posure of the binding sites after denaturation of the globular pro-
et al. (2015) found similar results, where the maximum ionic teins, as well as conformational rearrangement of the structure of
strength for coacervation was ࣘ20 mM; unlike Anema and de polysaccharides (Schmitt et al., 1998; Mc Clements, Decker, Park,
Kruif (2014), who found coacervate formation at ࣘ100 mM. Ac- & Weiss, 2009; Phillips & Williams, 2009).
cording to Tavares et al. (2015), results may vary depending on Nigem, Croguennec, Renard, and Bouhallab (2007) evaluated
the source variation and quality of the protein. These can include the effect of temperature on the interaction between two proteins.
the degree of iron saturation of the LF, which may affect flexi- The authors concluded that the interaction between α-Lg and
bility and the ability to coacervate, the presence of denatured or lysozyme occurred both at high (45 °C) and at low temperature
aggregative complexes, or even the type of coacervation medium. (5 °C), leading to the formation of precipitates. Reversible aggre-
Coacervation generally occurs in aqueous media with relevant pH gates were formed at 5 °C, whereas coacervates were obtained at
variation during protein mixing, which alters the ionic equilib- 45 °C. According to the authors, the formation of coacervates is
rium kinetics. Santos et al. (2018) also evaluated the influence of influenced by temperature because they are dependent on both
NaCl on the formation of heteroprotein complexes between BSA electrostatic interactions and the conformational change that oc-
and lysozyme, and they observed that, for minimal concentrations curs in the proteins above 27 °C. In addition, the capacity of
of salt (<10 mM), the interaction was reduced. aggregates formed at 5 °C to be converted to coacervation at
45 °C is due to the equilibrium of several forces involved in the
Molecular mass and ratio of biopolymers process (hydrogen, electrostatic, and hydrophobic interactions).
In the coacervation phase, the molecular agglomerates overlap Aberkane, Jasniewski, Gaiani, Scher, and Sanchez (2010) inves-
each other and increase in ionic density, resulting in an electro- tigated the effect of temperature (25 to 50 °C) to verify whether
static energy gain. The forces governing this separation are propor- hydrophobic interactions and hydrogen bonds play a significant
tional to molecular mass and charge density (Burgess, 1990). The role in the complexation of β-Lg and GA. This was also done
molecular mass corresponds to the volume that the biomolecule to calculate the heat capacity variation ( Cp ), which was highly
occupies, and thus a larger volume leads to more interactions that sensitive to interactions between macromolecule residues and sol-
can lead to formation of complexes (Phillips & Williams, 2009). vent molecules. The authors verified that at 25 and 35 °C an
According to Weinbreck, Tromp, and de Kruif (2004a), additional exothermic–endothermic sequence was recorded during the titra-
positively charged proteins are required to compensate for the tion, indicating that the first stage was under the control of en-
negative charges of the polysaccharide at higher pH values. The thalpy and the second of entropy. In contrast, at 45 and 50 °C,
protein becomes more charged and the polysaccharide less charged the isotherm exhibited an endothermic-exothermic sequence. Al-
at lower pH, and thus fewer protein molecules will be required for though other factors are also responsible for the negative value in
charge-balancing. Cp, because it is related to the proportion between polar and
In studies by Ach et al. (2015), the pH of higher yield of com- nonpolar groups, according to the authors, the negative value in
plex formation increased positively as the ratio between α-Lg the second binding stage may indicate a greater contribution of
/β-Lg proteins and GA increased. This yield was measured from hydrophobic interactions in the binding process compared to the
the protein concentration of the supernatant. In this same study, first stage where there was predominance of ionic bonds veri-
the authors compared α-Lg/β-Lg behavior in the coacervation fied by positive Cp (Kayitmazer, 2017). The basis for the author’s
process. First, maximum coacervation occurred at various pH val- argument lies in the fact that removal of the contact between
ues according to their pI, that is, pH 4.5 for β-Lg and pH 3.4 the surface area of the protein and the solvent results in a negative
for α-Lg. Next, the authors concluded that the interaction was Cp . This behavior is particularly related to the interaction of wa-
more intense with β-Lg than with α-Lg, according to the residual ter molecules with hydrophobic sites. Through this correlation of
concentration of protein in the aqueous phase at maximum pH. Cp with the burial of the surface area, the heat capacity provides
Based on these results and comparing the characteristics of each a link between thermodynamic data and structural information
protein, they suggested that the molar mass difference and chemi- (Perozzo, Folkers, & Scapozza, 2004).
cal structure were mutually responsible for the interaction affinity Hadian et al. (2016) studied the interaction between β-Lg and
difference (Ach et al., 2015). WPG, finding a high affinity at 25 °C. However, affinity and sto-
In the case of heteroprotein complexes, the difference in size ichiometry both underwent changes with increasing temperature
between protein pairs (acidic and basic) is a key element that (40 °C). They concluded that hydrophobic interactions became
defines both the optimal pH and the formation ratio (Desfougeres, more significant than hydrogen bonds with increased temperature.
Croguennec, Lechevalier, Bouhallab, & Nau, 2010). Anema and
de Kruif (2014) found complexes with a constant stoichiometry Application of whey proteins as a bioactive encapsulating
for LF (18.3 kDa): β-Lg (80 kDa) of 1:3 at all pH values due to the agent: complex coacervation
regularization of the charge. In contrast to Yan et al. (2013), they Microencapsulation by complex coacervation is classified by
found a better ratio of 1:1 for the same biopolymers. Diarrassouba McClements as delivery systems based on biopolymers. In this
et al. (2015) observed higher turbidity at a β-Lg: lysozyme ratio case, the basis for the construction of this system are biopoly-
of 2:1, in favor of β-Lg which is larger than lysozyme (14 kDa). mers, such as proteins and polysaccharides (McClements, 2014).


C 2018 Institute of Food Technologists® Vol. 0, 2018 r Comprehensive Reviews in Food Science and Food Safety 7
Complexes between WPI and biopolymers . . .

Its use as an encapsulation method has the advantage of simplic- omega-3 rich oil and probiotic separately, the authors were able
ity and does not require the use of high temperatures or organic to conclude that there was a synergistic effect providing greater
solvents. However, its use is limited by pH range, colloid concen- fermentability of bacteria and greater stability of omega-3 oil in
tration, and/or electrolyte concentration based on the polymers spray-dried microcapsules. In a subsequent work, Eratte, Dowling,
used (Comunian et al., 2013). Barrow, and Adhikari (2017) performed in vitro digestion (salivary,
The basic procedure for encapsulation by coacervation was de- gastric, and intestinal) of these same co-microcapsules, checking
veloped in the work of Bungenberg de Jong (1929) using gelatin for probiotic bacteria survival and oil release. According to the
and GA. The first step is emulsification, where the core mate- authors, co-microencapsulation increased the viability of bacteria
rial is added to a solution of cationic polymer under stirring. during the digestion and retention of omega-3 fatty acids.
The second is coacervation, where the anionic biopolymer is Microencapsulation of probiotic bacteria (Lactobacillus paraca-
added, followed by pH adjustment to initiate the interactions. sei) using WPI and GA complexes was also proposed by Bosnea,
The obtained microcapsules have reduced mechanical resistance Moschakis, and Biliaderis (2017); the authors evaluated the addi-
due to the ionic nature of the interaction, which implies the tion of the microencapsules in yogurts. They observed that en-
use of a crosslinking agent (Wang et al., 2014; Zhang, Pan, & capsulated cells showed increased survival during storage and gas-
Chung, 2011). Finally, after obtaining a precipitate of microcap- tric simulation compared to yogurts containing nonencapsulated
sules, the material is dried by atomization or lyophilization to give a cells. In addition, it was demonstrated that the incorporation of
powder. coacervate complexes did not significantly affect the rheological
The complex coacervation microencapsulation process is illus- properties of the yogurts.
trated in Figure 2. The active ingredient (Oil) is added to the Coacervate complexes formed by WPI and GA were also used in
polymer 1 solution and, subsequently, the previously solubilized the encapsulation of pigments. Jain et al. (2015) microencapsulated
polymer 2 is added and the pH is adjusted, similar to Jain, Thakur, β-carotene by complex coacervation at an optimal pH of 4.2 and
Ghoshal, Katare, and Shivhare (2015), who used WPI and GA as WPI: GA ratio of 2:1, obtaining more than 70% encapsulation
a wall material in the encapsulation of β-carotene and glutaralde- efficiency. The microcapsules showed better stability than free β-
hyde as crosslinking agent. carotene, and strong antioxidant activity even after 3 months of
Table 2 lists the most-recent studies that used coacervate com- storage.
plexes formed between whey proteins and biopolymers as a wall In addition to WPI, recent studies have demonstrated not only
material for microencapsulation of different active ingredients in the formation of coacervate complexes with whey proteins but
addition to presenting the main analyses and parameters studied also their use in the encapsulation of compounds. Among them,
and the efficiency of encapsulation obtained. BSA was studied in the encapsulation of pigments from the inter-
The use of complexes formed by whey protein and polysac- action with GA. Shahgholian and Rajabzadeh (2015) promoted
charide as anionic biopolymer in the microencapsulation of hy- the encapsulation of curcumin using BSA and GA. For the forma-
drophilic and hydrophobic bioactive compounds has been pro- tion of the complexes, ideal conditions were found in a BSA: GA
posed by numerous authors. Among the bioactive compounds of ratio of 2:1 and pH of 3.7 with maximum encapsulation efficiency
hydrophobic character microencapsulated by these complexes are (92%) and particle size of 40 to 80 μm. According to the authors,
the oils, probiotic bacteria, pigments, flavoring and vitamins. the technique is efficient and promising for its high-efficiency
With regard to the biopolymers used in the encapsulations of encapsulation and ease of the method.
these bioactives, the pair WPI and GA had been the most studied. Carboxymethylcellulose (CMC) is an anionic polymer derived
Weinbreck, Minor, and de Kruif (2004b) studied the microencap- from cellulose studied by some authors in the formation of coac-
sulation of sunflower, lemon, and orange oils by complex coac- ervate complexes. However, studies of microencapsulation are still
ervation using WPI and GA. The stability of the microcapsules limited. The encapsulation of a flavoring (β-pyrene) by com-
prepared via emulsion, followed by complex coacervation, was plex coacervation was proposed by Koupantsis, Pavlidou, and
strongly dependent on the pH of the medium. The capsules were Paraskevopoulou (2014) using WPI and CMC. Coacervates pre-
formed at pH 4.0, where the strength of the electrostatic inter- pared in the highest proportion of WPI: CMC and β-pyrene mass
action and viscosity were maximal, and the encapsulated oil yield (6.99 g) were most effective. The authors achieved a maximum en-
was greater than 80%. Lemon oil capsules were added to cheese capsulation yield of 42% and maximum encapsulation efficiency
to assess flavor release. The authors observed that the size of the of 37%. The morphological characteristics were also evaluated,
capsules and the crosslinking process influenced the release of oil. concluding that the formation of microcapsules with compact or
Larger and uncrosslinked capsules showed greater flavor release spongy structures depended on the concentration of biopolymer
than smaller and crosslinked capsules. and β-pyrene. The spongy structures formed at higher WPI:CMC
Using WPI, Eratte, Wang, Dowling, Barrow, and Adhikari ratios showed a lower encapsulation efficiency and yield than com-
(2014) produced omega-3 rich tuna oil microcapsules by com- pact structures. The authors suggested that the void spaces of the
plex coacervation. The authors found that the ideal pH and WPI: spongy structures would facilitate evaporation of the volatile con-
GA ratio for complex coacervation were 3.7 and 3:1, respec- stituent.
tively. In addition, the authors confirmed that the drying method In addition to the hydrophobic compounds, the hydrophilic
also influenced oil stability, and they concluded that spray-dried character compounds microencapsulated by complex coacerva-
solid microcapsules showed better stability against oxidation than tion using whey proteins reported in the literature are pigments,
lyophilized microcapsules. proteins, and vitamins.
Eratte et al. (2015) evaluated the interaction between probiotic With positively charged biopolymer chitosan and negatively
bacteria (Lactobacillus casei) and omega-3 oils co-encapsulated in charged BSA, Kurukjia, Norton, and Spyropoulos (2016) pro-
WPI and GA, as well as the effect of the drying, temperature, posed the use of such complexes in the encapsulation of hy-
and time on the oxidative stability of omega-3 and viability of drophilic compounds, using as a model fluorescein, rhodamine
probiotic bacteria. Comparing the results of the microcapsules of B, and riboflavin. The results showed that the structure of the

8 Comprehensive Reviews in Food Science and Food Safety r Vol. 0, 2018 


C 2018 Institute of Food Technologists®
Complexes between WPI and biopolymers . . .

Figure 2–Microencapsulation process by complex coacervation.

Table 2–The most recent studies using coacervate complexes formed between whey proteins and biopolymers for microencapsulation of different
active ingredients.

Wall material Core material Tests Parameters EE (%) Reference


WPI and GA Sunflower / lemon ζ , MO, pH 4.0, r = 2:1, 5% oil – Weibreck et al. (2004b)
/ orange oil
PI and pectin Thiamine τ , ζ , DLS, MO, HPLC pH 3.5, r = 2:1 – Bédié et al. (2008)
WPI and CMC β Pyrene ζ , SEM, GC-FID pH 2.8, r = 6.99, β-Pinene 37 Koupantsis et al. (2014)
= 6.99 g
WPI and GA Tuna oil τ , ζ , CLSM, oxidative pH 3.75, r = 3:1 90 Eratte, Wang,
(ômega-3) estability, SEM, MO Dowling, Barrow, and
Adhikari (2014)
WPI and GA β Carotene DLS, ζ , CLSM, SEM, FTIR, in pH 4.2, r = 2:1 77 Jain et al. (2015)
vitro testing, estability
(time and T °C) DPPH,
BSA and GA Curcumin SEM, XRD, DSC, texture pH 3.7, r = 2:1, Curcumin 92 Shahgholian and Rajabzadeh
analysis = 5000 mg, (2015)
β-Lg and lysozyme Vitamin D3 Estability (time, UV), in vitro pH 7.5, r = 2:1, D3 110 90 Diarrassouba et al. (2015)
testing (SIF), in vivo testing μM,
WPI and GA L. casei and tuna Estability (time e T °C) pH 3.75 – Eratte et al. (2015)
oil
WPI and CMC β Pyrene FTIR, TGA, SEM pH 2.8, r = 6.99 36 Koupantsis et al. (2016)
BSA and chitosan Hydrophilic DLS, ζ , τ , MO, SEM, in vitro pH 5.0, r = 3:1, Ct = 1% 85 Kurukji, Norton, and
actives testing Spyropoulos (2016)
WPI and GA Lactobacillus Estability (time), in vitro pH 4.0, r = 2:1, Ct = 3% – Bosnea et al. (2016)
paracasei testing (SGF), rheology
β-Lg and LF Vitamin B9 ζ , DLS, MO, Estability pH 5.5, B9 0,25 mM, Cp 97 Chapeaou et al. (2017)
0.55 mM
WPI and pectin LF DLS, ζ , SEM, AFM, FTIR, DSC, pH 3.5, r = 2:1 25 Raei et al. (2017)
HPLC
WPI and pectin/GA Anthocyanin Estability (T °C), Confocal, pH 3.2, r = 1:1, 500 μL 99 Stănciuc et al. (2017)
colorimetry anthocyanin
WPI and GA L. casei and tuna In vitro testing pH 3.75 – Eratte et al. (2017)
oil
Definitions of abbreviations are shown in the Abbreviations section.

compounds influences the encapsulation efficiency. According to pectin ratio, and the acidification method, whether before or after
the authors, the encapsulation efficiency at pH 5.0 was similar for polymer mixing (pre-mixing and post-mixing acidification). They
rhodomine B and fluorescein (65%). However, it has not been concluded that the maximum formation of complexes occurred
possible to encapsulate riboflavin under the same conditions, as it at pH 3.5 with a 2:1 ratio, and the encapsulation efficiency was
is a more basic molecule. higher for the pre-acidification method.
Stănciuc et al. (2017) compared microcapsules formed by GA Recently, coacervate complexes formed through the interaction
and pectin together with WPI in the encapsulation of antho- between two different proteins have been the subject of studies for
cyanins. Significantly, the amount of encapsulated flavonoids and the encapsulation of both hydrophobic and hydrophilic bioactive
the protective effect against anthocyanin degradation was higher compounds. Diarrassouba et al. (2015) used β-Lg and lysozyme to
in pectin-containing microcapsules compared to GA. encapsulate vitamin D3 using a ratio of 2:4:1 (β-Lg: D3 :lysozyme)
The use of complexes formed between pectin and WPI was at pH 7.5, according to previous studies. The data obtained in-
also examined in the encapsulation of vitamins and proteins. dicated that microspheres effectively improved long-term stability
Bédié, Turgeon, and Makhlouf (2008) studied the encapsulation (4 °C) and reduced the photochemical degradation of vitamin D3 .
of thiamine (vitamin B1 ) by low-methoxylation pectin, and Raei, In addition, simulating the intestinal conditions, the authors ver-
Shahidi, Farhoodi, Jafari, and Rafe (2017) studied encapsulation ified not only that the capsules not only retained the vitamin but
of LF by high-methoxylation pectin; both observed similar re- that they were also released in the intestine for later absorption.
sults, even though the degrees of methoxylation of the pectins Chapeau et al. (2017) microencapsulated vitamin B9 through
were different. The authors found that the formation of com- heteroprotein complexes formed between β-Lg and LF at pH 5.5.
plexes was highly dependent on the pH of the medium, the WPI: The authors compared the performance of aggregates formed at


C 2018 Institute of Food Technologists® Vol. 0, 2018 r Comprehensive Reviews in Food Science and Food Safety 9
Complexes between WPI and biopolymers . . .

low and high concentrations of protein and heteroprotein coac- Conclusion


ervates. These were formed in intermediary concentrations as Whey proteins are biopolymers with excellent nutritional, func-
models of biochargers and showed that the coacervates showed tional, and technological properties and are an attractive alternative
higher performance than the aggregates. In this study, the authors for the encapsulation of active ingredients and construction of de-
found that the optimal encapsulation occurred at 10 mg of vitamin livery systems. Due to their characteristics, they can form particles
per g of protein, with the formation of spherical or slightly ovoid with other biopolymers through complex coacervation and are
structures with an average diameter of 8 μm. used because they do not require chemical reagents or elevated
The use of crosslinking agents such as glutaraldehyde, glycerol, temperatures. Complexes between whey proteins and biopoly-
tannic acid, or transglutaminase has been investigated by some mers are formed mainly by electrostatic interactions and are there-
authors to increase the strength of capsules and, consequently, the fore governed by intrinsic and extrinsic factors, such as pH, ionic
retention time of the active ingredient. Jain et al. (2015) and Eratte strength, charge density, biopolymer ratio, and molecular mass, in
et al. (2017) studied the effect of crosslinking on capsules formed addition to the type of crosslinker and the characteristics of the
by WPI and GA. In the case of Jain et al. (2015), glutaraldehyde active ingredients and biopolymers that influence the final prop-
promoted the crosslinking of amide bonds between the amino erties of the produced capsules. The behavior of these complexes,
acid residues of WPI. The authors observed that microcapsule as well as their structure and kinetics of formation, have been
images using scanning electron microscopy (SEM) depended on studied through different analytical techniques, presenting appli-
crosslinking density and that the size and shape of the particles cability in food products. Despite these numerous works, specific
were maintained, which was successful in crosslinking because of conditions are required for certain active ingredients to obtain an
the agent. encapsulation efficiency and appropriate delivery system. Their
Transglutaminase is the most commonly used crosslinker in use as an encapsulating agent has become promising for a vari-
foods due to the greater potential for cross-linking in protein- ety of active ingredients and applications in different products and
based coacervates and its safety and commercial availability without formulations.
causing sensory alteration in the final product (Bakry et al., 2016).
According to Eratte et al. (2017), transglutaminase crosslinking Conflicts of Interest
favored retention of the active ingredient during salivary and gas- The authors declare that there are no conflicts of interest.
tric simulation. These authors reported that the smaller amount
of oil released in simulated salivary fluid (SSF) and SGF (simu- Acknowledgments
lated gastric fluid) compared to SIF (simulated intestinal fluid) can The authors thank the Nat. Council for Scientific and Techno-
be attributed to lower degradation of the WPI-GA complex by logical Development (CNPq) and the Foundation for Research
salivary and gastric enzymes. The stability of these complexes at Support of the State of Rio de Janeiro (FAPERJ) for financial
low pH (3.0) was reported to be due to the crosslinking process support.
occurring at this pH. The literature indicates that the chemical
crosslinking of protein in coacervated complexes reduces the pro- Abbreviations
teolytic activity of gastric enzymes, resulting in a slower release of AFM Atomic force microscopy
oil under simulated gastric conditions (Kosaraju, Weerakkody, & α-Lg α-lactoglobulin
Augustin, 2009). Apo-LF Apo-lactoferrin
Different crosslinkers, glycerol and tannic acid, were tested by BCAA Higher branched-chain amino acid
Koupantsis, Pavlidou, and Paraskevopoulou (2016) during the pro- β-Lg β-lactoglobulin
duction of β-pyrene microcapsules formed by WPI and CMC. BSA Bovine serum albumin
The authors verified that the morphological characteristics of the CaCl2 Calcium chloride
complexes varied as a function of the crosslinking agent used. The CD Circular dichroism
addition of glycerol also increased the encapsulation efficiency and CLSM Confocal laser scanning microscopy
amount of encapsulated β-pyrene. The authors concluded that CMC Carboxymethylcellulose
both the hydrogen and hydrophobic binding, as well as the differ- CNaCl Concentration of salt or ionic strength
ence in the molecular size of the crosslinkers, were responsible for CT Total concentration of biopolymers
the observed behaviors. DLS Dynamic light scattering
In addition to reports in the literature, a variety of patents address DPPH Antioxidant capacity
the encapsulation of ingredients from complex coacervation. The DSC Calorimetric differential scanning
microencapsulation of hydrophobic oil particles was patented by EE Encapsulation efficiency
Green and Lowell (1957) using gelatin, casein, or GA, which are FTIR Fourier transform infrared spectroscopy
the primary materials approved for application in food products. G ‘and G’ Dynamic rheometry
Rosenberg (1997) discovered the functionality of whey proteins GA Gum acacia
to microencapsulate volatile compounds from the formation of GC-FID Flame ionization gas chromatography
emulsions. A patent was published on the formation of these simple GPC Gel permeation chromatography
microcapsules, which were easy to handle, had high stability, and Holo-LF Holo-lactoferrin
protected against deterioration of the encapsulated compound. HMP High-methoxylation pectin
Microencapsulation from the complex coacervation of whey HPLC High-performance liquid chromatography
proteins was patented by Weinbreck, Kruif, and Schrooyen (2003a) HSA Human serum albumin
as an alternative to gelatin. The core materials may be vitamins, IFS Intrinsic fluorescence spectroscopy
oils, fats, medicines, or pesticides. Weak polyelectrolytes include Ig Immunoglobulins
anionic polysaccharides, such as GA, pectin, gellan gum, alginate, IUPAC International Union of Pure and Applied
CMC, among others, or cationic polysaccharides like chitosan. Chemistry

10 Comprehensive Reviews in Food Science and Food Safety r Vol. 0, 2018 


C 2018 Institute of Food Technologists®
Complexes between WPI and biopolymers . . .

ITC Isothermal titration calorimetry Anema, S. G., & de Kruif, C. G. (2016). Phase separation and composition of
LF Lactoferrin coacervates of lactoferrin and caseins. Food Hydrocolloids, 52, 670–77.
https://doi.org/10.1016/j.foodhyd.2015.08.011.
LMP Low-methoxylation pectin
Bakry, A. M., Abbas, S., Ali, B., Majeed, H., Abouelwafa, M. Y., Mousa, A.,
MO Optical microscopy & Liang, L. (2016). Microencapsulation of oils: A comprehensive review of
n Binding stoichiometry benefits, techniques, and applications. Comprehensive Reviews in Food Science
NaCl Sodium chloride and Food Safety, 15, 143–182. http://10.1111/1541-4337.12179.
pHc Region of slight increase in turbidity Balcão, V. M., Costa, C. I., Matos, C. M., Moutinho, C. G., Amorim, M.,
Pintado, M. E., . . . Teixeira, J. A. (2013). Nanoencapsulation of bovine
pHθ 1 Region of abrupt turbidity increase lactoferrin for food and biopharmaceutical applications. Food Hydrocolloids,
pHθ 2 Region of dissociation of the complexes 32, 425–431. https://doi.org/10.1016/j.foodhyd.2013.02.004.
pHmax Maximum turbidity point Bastos, D. S., Barreto, B. N., Souza, H. K. S., Bastos, M., Rocha-Leão, M.
pI Isoelectric point H. M., Andrade, C. T., & Gonçalves, M. P. (2010). Characterization of a
pKa (−log Ka ) chitosan sample extracted from Brazilian shrimps and its application to
obtain insoluble complexes with a commercial whey protein isolate. Food
r Reason Hydrocolloids, 24, 709–718. https://doi.org/10.1016/j.foodhyd.2010.03.008.
SALS Small angle light scattering Bédié, G. K., Turgeon, S. L., & Makhlouf, J. (2008). Formation of native
SANS Small-angle neutron scattering whey protein isolate–low methoxyl pectin complexes as a matrix for
SAXS Small angle X-ray scattering; hydro-soluble food ingredient entrapment in acidic foods. Food
SDS- PAGE Sodium dodecyl sulfate polyacrylamide gel elec- Hydrocolloids, 22, 836–844. https://doi.org/10.1016/j.foodhyd.2007.03.010.
Bengoechea, C., Jones, O. G., Guerrero, A., & McClements, D. J. (2011).
trophoresis Formation and characterization of lactoferrin/pectin electrostatic
SEC Size-exclusion chromatography complexes: Impact of composition, pH and thermal treatment. Food
SEM Scanning electron microscopy Hydrocolloids, 25, 1227–1232. https://doi.org/10.1016/j.foodhyd.2010.
SGF Simulated gastric fluid 11.010.
SIF Simulated intestinal fluid Benichou, A., Aserin, A., & Garti, N. (2002). Protein-polysaccharide
interactions for stabilization of food emulsions. Journal of Dispersion Science
SSF Simulated salivary fluid and Technology, 23, 93–123. https://doi.org/10.1080/01932690208984192.
T Temperature Bokkhim, H., Bansal, N., Grøndahl, L., & Bhandarim, B. (2015).
TEM Transmission electron microscopy Interactions between different forms of bovine lactoferrin and sodium
TGA Thermogravimetric analysis alginate affect the properties of their mixtures. Food Hydrocolloids, 48, 38–46.
https://doi.org/10.1016/j.foodhyd.2014.12.036.
TMP Transmembrane pressure
Boland, M., Singh, H., & Thompson, A. (2014). Milk proteins: From expression
USANS Ultra-small-angle scattering to food (2nd ed). San Diego: Academic Press.
WPG Water-soluble fraction of Persian gum Bosnea, L. A., Moschakis, T., & Biliaderis, C. G. (2017). Microencapsulated
WP Whey protein cells of Lactobacillus paracasei subsp. paracasei in biopolymer complex
WPC Whey protein concentrate coacervates and their function in a yogurt matrix. Food & Function, 8,
554–562. https://doi.org/10.1039/C6FO01019A.
WPI Whey protein isolate
Bujacz, A. (2012). Structures of bovine, equine and leporine serum albumin.
WPH Whey protein hydrolyzed Acta Acta Crystallographica. Section D, Biological Crystallography, 68,
XRD X-ray diffraction 1278–1289. https://doi.org/10.1107/S0907444912027047.
δ shear stress Bungenberg De Jong, H. B. (1949). Complex colloid systems. Colloid science,
Cp Heat capacity variation 2, 335–432. https://doi.org/10.2516/ogst/2009070.
H Enthalpy change Bungenberg De Jong, H. B., & Kruyt, H. R. (1929). Coacervation (partial
μ Electrophoretic mobility miscibility in colloid systems). Recueil des Travaux Chimiques des Pays-Bas, 32,
849–856.
η Intrinsic viscosity
Burgess, D. J. (1990). Practical analysis of complex coacervate systems. Journal
δ Shear stress of Colloid and Interface Science, 14, 227–238.
τ Turbidity https://doi.org/10.1016/0021-9797(90)90338-O.
ζ Zeta potential Caroli, A. M., Chessa, S., & Erhardt, G. J. (2009). Invited review: Milk
protein polymorphisms in cattle: Effect on animal breeding and human
nutrition. Journal of Dairy Science, 92, 5335–5352.
References
https://doi.org/10.3168/jds.2009-2461.
Aberkane, L., Jasniewski, J., Gaiani, C., Scher, J., & Sanchez, C. (2010). Carter, D. C., & Ho, J. X. (1994). Structure of serum-albumin. Advances in
Thermodynamic characterization of acacia gum−β-lactoglobulin complex Protein Chemistry, 45, 153–203.
coacervation. Langmuir, 26, 12523–12533. http://10.1021/la100705d. https://doi.org/10.1016/S0065-3233(08)60640-3.
Aberkane, L., Jasniewski, J., Gaiani, C., Hussain, R., Scher, J., & Sanchez, C.
Carunchia, W. M. E., Croissant, A. E., & Drake, M. A. (2005).
(2012). Structuration mechanism of β-lactoglobulin – acacia gum assemblies
Characterization of dried whey protein concentrate and isolate flavor.
in presence of quercetin. Food Hydrocolloids, 29, 9–20.
Journal of Dairy Science, 88, 3826–39.
https://doi.org/10.1016/j.foodhyd.2012.01.010. https://doi.org/10.3168/jds.S0022-0302(05)73068-X.
Ach, D., Briançon, S., Dugas, V., Pelletier, J., Broze, G., & Chevalier, Y.
Chai, C., Lee, J., & Huang, Q. (2014). The effect of ionic strength on the
(2015). Influence of main whey protein components on the mechanism of
rheology of pH-induced bovine serum albumin/κ-carrageenan coacervates.
complex coacervation with Acacia gum. Colloid Surface A, 481, 367–374.
LWT-Food Science and Technology, 59, 356–360.
https://doi.org/10.1016/j.colsurfa.2015.06.006.
https://doi.org/10.1016/j.lwt.2014.05.024.
Adal, E., Sadeghpour, A., Connell, S., Rappolt, M., Ibanoglu, E., & Sarkar,
A. (2017). Heteroprotein complex formation of bovine lactoferrin and pea Chapeau, A. L., Hamon, P., Rousseau, F., Croguennec, T., Poncelet, D., &
protein isolate: A multiscale structural snalysis. Biomacromolecules, 18, Bouhallab, S. (2017). Scale-up production of vitamin loaded heteroprotein
625–635. http://10.1021/acs.biomac.6b01857. coacervates and their protective property. Journal of Food Engineering, 206,
67–76. https://doi.org/10.1016/j.jfoodeng.2017.03.005.
Almeida, C. C., Conte-Júnior, C. A., Silva, A. C. O., & Alvares, T. S. (2013).
Whey protein: Composition and functional properties. Enciclopédia Biosfera, Croguennec, T., Tavares, G. M., & Bouhallab, S. (2017). Heteroprotein
9, 1840–1854. complex coacervation: A generic process. Advences in Colloid and Interface
Science, 239, 115–126. https://doi.org/10.1016/j.cis.2016.06.009.
Anema, S. G., & de Kruif, C. G. K. (2014). Complex coacervates of
lactotransferrin and β-lactoglobulin. Journal of Colloid and Interface Science, Codex Alimentarius. (2011). Codex standart for fermented milk. Codex
430, 214–220. http://10.1016/j.jcis.2014.05.036. Alimentarius Comission. Milk and milk products, (2nd ed). Codex Stand
243–2003.


C 2018 Institute of Food Technologists® Vol. 0, 2018 r Comprehensive Reviews in Food Science and Food Safety 11
Complexes between WPI and biopolymers . . .

Comunian, T. A., Thomazini, M., Alves, A. J. G., de Matos Junior, F. E., de Antimicrobial Agents, 33, 301, e1–301.e8.
Carvalho Balieiro, J. C., & Favaro-Trindade, C. S. (2013). https://doi.org/10.1016/j.ijantimicag.2008.07.020.
Microencapsulation of ascorbic acid by complex coacervation: Protection González-Martı́nez, D. A., Carrillo-Navas, H., Barrera-Dı́az, C. E.,
and controlled release. Food Research International, 52, 373–379. Martı́nez-Vargas, S. L., Alvarez-Ramı́rez, J., & Pérez-Alonso, C. (2017).
https://doi.org/10.1016/j.foodres.2013.03.028. Characterization of a novel complex coacervate based on whey protein
Damodaran, P. K. L., & Fennema, O. R. (2010). Quı́mica de Alimentos de isolate-tamarind seed mucilage. Food Hydrocolloids, 72, 115–126.
Fennema (4th ed). Porto Alegre: Editora Artmed SA. https://doi.org/10.1016/j.foodhyd.2017.05.037.
Desfougeres, Y., Croguennec, T., Lechevalier, V., Bouhallab, S., & Nau, F. Green, B. K., & Lowell, S. (1957). Oil-containing microscopic capsules and method
(2010). Charge and size drive spontaneous self-assembly of oppositely of making them. Google Patents
charged globular proteins into microspheres. The Journal of Physical Chemistry Guimont, C., Marchall, E., Girardet, J. M., & Linden, G. (1997). Biologically
B, 114, 4138–4144. https://doi.org/10.1021/jp9090427. active factors in bovine milk and dairy byproducts: Influence on cell culture.
de Kruif, C. G., & Tuinier, R. (2001). Polysaccharide protein interactions. Critical Reviews in Food Science and Nutrition, 37, 393–410.
Food Hydrocolloids, 15, 555–563. https://doi.org/10.1080/10408399709527780.
https://doi.org/10.1016/S0268-005X(01)00076-5. Gulão, E. S., de Souza, C. J. F., da Silva, F. A. S., Coimbra, J. S. R., &
de Kruif, C. G., Weinbreck, F., & de Vries, R. (2004). Complex Garcia-Roja, E. E. (2014). Complex coacervates obtained from lactoferrin
coacervation of proteins and anionic polysaccharides. Current Opinion in and gum arabic: Formation and characterization. Food Research International,
Colloid & Interface Science, 9, 340–349. https://doi.org/10.1016/j.cocis. 65, 367–374. https://doi.org/10.1016/j.foodres.2014.08.024.
2004.09.006. Gurov, A. N., Gurova, N. V., Leontiev, A. L., & Tolstoguzov, V. B.
Diarrassouba, F., Remondetto, G., Garrait, G., Alvarez, P., Beyssac, E., & (1988). Equilibrium and non-equilibrium complexes between bovine
Subirade, M. (2015). Self-assembly of β-lactoglobulin and egg white serum albumin and dextran sulfate—I. Complexing conditions and
lysozyme as a potential carrier for nutraceuticals. Food Chemintry, 173, composition of non-equilibrium complexes. Food Hydrocolloids, 2,
203–209. https://doi.org/10.1016/j.foodchem.2014.10.009. 267–283. https://doi.org/10.1016/S0268-005X(88)80025-0.
Dong, L., Chen, X., & Hu, Z. (2007). Study of the effect of cal-red on the Hadian, M., Hosseini, S. M. H., Farahnaky, A., Mesbahi, G. R., Yousefi,
secondary structure of human serm albumin by spectroscopic techniques. G. H., & Saboury, A. A. (2016). Isothermal titration calorimetric and
Journal of Molecular Structure, 846, 112–118. spectroscopic studies of β-lactoglobulin-water-soluble fraction of Persian
https://doi.org/10.1016/j.molstruc.2007.01.034. gum interaction in aqueous solution. Food Hydrocolloids, 55, 108–118.
Doublier, J. L., Garnier, C., Renard, D., & Sanchez, C. (2000). https://doi.org/10.1016/j.foodhyd.2015.11.006.
Protein–polysaccharide interactions. Current Opinion in Colloid Haraguchi, F. K., Abreu, W. C., & Paula, H. (2006). Proteı́nas do soro do
& Interface Science, 5, 202–214. leite: Composição, propriedades nutricionais, aplicações no esporte e
https://doi.org/10.1016/S1359-0294(00)00054-6. benefı́cios para a saúde humana. Revista de Nutrição, 19, 479–488.
Du, X., Dubin, P. L., Hoagland, D. A., & Sun, L. (2014). Protein-selective https://doi.org/10.1590/S1415-52732006000400007.
coacervation with hyaluronic acid. Biomacromolecules, 15, 726–734. Harnsilawat, T., Pongsawatmanit, R., & McClements, D. J. (2006).
https://doi.org/10.1021/bm500041a. Characterization of β-lactoglobulin–sodium alginate interactions in aqueous
Du, J., Reuhs, B. L., & Jones, O. G. (2016). Influence of PEGylation on the solutions: A calorimetry, light scattering, electrophoretic mobility and
ability of carboxymethyl-dextran to form complexes with α-lactalbumin. solubility study. Food Hydrocolloids, 20, 577–585.
Food Chemintry, 196, 853–859. https://doi.org/10.1016/j.foodhyd.2005.05.005.
https://doi.org/10.1016/j.foodchem.2015.10.021. Hirt, S., & Jones, O. G. (2014). Effects of chloride, thiocyanate and sulphate
Dubin, P. L., Gao, J., & Mattison, K. (1994). Protein purification by selective salts on β-lactoglobulin–pectin associative complexes. International Journal of
phase separation with solyelectrolytes. Separation & Purification Reviews, 23, Food Science and Technology, 49, 2391–2398.
1–16. https://doi.org/10.1080/03602549408001288. https://doi.org/10.1111/ijfs.12560.
Emerson, T. E. (1989). Unique features of albumin - A brief review. Critical IUPAC. (1997). Compendium of chemical terminology (2rd ed). Oxford:
Care Medicine, 17, 690–694. Blackwell Scientific Publications.
Eratte, D., Wang, B., Dowling, K., Barrow, C. J., & Adhikari, B. P. (2014). Jain, A., Thakur, D., Ghoshal, G., Katare, O. P., & Shivhare, U. S. (2015).
Complex coacervation with whey protein isolate and gum arabic for the Microencapsulation by complex coacervation using whey protein isolates
microencapsulation of omega-3 rich tuna oil. Food and Function, 5, and gum acacia: An approach to preserve the functionality and controlled
2743–2750. https://doi.org/10.1039/c4fo00296b. release of β-carotene. Food Bioprocess Technology, 8, 1635–1644.
https://doi.org/10.1007/s11947-015-1521-0.
Eratte, D., McKnight, S., Gengenbach, T. R., Dowling, K., Barrow, C. J., &
Adhikari, B. P. (2015). Co-encapsulation and characterisation of omega-3 Jones, O. G., Handschin, S., Adamcik, J., Harnau, L., Bolisetty, S., &
fatty acids and probiotic bacteria in whey protein isolate–gum Arabic Mezzenga, R. (2011). Complexation of β-lactoglobulin fibrils and sulfated
complex coacervates. Journal of Functional Food, 19, 882–892. polysaccharides. Biomacromolecules, 12, 3056–3065.
https://doi.org/10.1016/j.jff.2015.01.037. https://doi.org/10.1021/bm200686r.
Eratte, D., Dowling, K., Barrow, C. J., & Adhikari, B. P. (2017). In-vitro Jovanovi, S., Bara, M., & Ma, O. (2005). Whey proteins-properties and
digestion of probiotic bacteria and omega-3 oil co-microencapsulated in possibility of application. Mljekarstvo: Journal of Dairy Productionand Processing
whey protein isolate-gum Arabic complex coacervates. Food Chemistry, 227, Improvement, 55, 215–233.
129–136. https://doi.org/10.1016/j.foodchem.2017.01.080. Kayitmazer, A. B., Strand, S. P., Tribet, C., Jaeger, W., & Dubin, P. L. (2007).
Etzel, M. R. (2004). Manufacture and use of dairy protein fractions. The Effect of polyelectrolyte structure on protein−polyelectrolyte coacervates:
Journal of Nutrition, 134, 996–1002. Coacervates of bovine serum albumin with poly (diallyldimethylammonium
chloride) versus chitosan. Biomacromolecules, 8, 3568–3577.
Farrell, H. M., Jimenez-Flores, R., Bleck, G. T., Brown, E. M., Butler, J. E., https://doi.org/10.1021/bm700645t.
Creamer, L. K., . . . Swaisgood, H. E. (2004). Nomenclature of the
proteins of cows’ milk–sixth revision. Journal of Dairy Science, 87, 1641–74. Kayitmazer, A. B. (2017). Thermodynamics of complex coacervation.
https://doi.org/10.3168/jds.S0022-0302(04)73319-6. Advances in Colloid and Interface Science, 239, 169–177.
https://doi.org/10.1016/j.cis.2016.07.006.
Flanagan, S. E., Malanowski, A. J., Kizilay, E., Seeman, D., Dubin, P. L.,
Donato-Capel, L., . . . Schmitt, C. (2015). Complex equilibria, speciation, Kinsella, J. E., & Whitehead, D. M. (1989). Proteins in whey: Chemical,
and heteroprotein coacervation of lactoferrin and β lactoglobulin. Langmuir, physical and functional properties. Advences in Foods and Nutrition Research,
31, 1776–1783. https://doi.org/10.1021/la504020e. 3, 343–438. https://doi.org/10.1016/S1043-4526(08)60130-8.
Girard, M., Turgeon, S. L., & Paquin, P. (2002a). Emulsifying properties of Kosaraju, S. L., Weerakkody, R., & Augustin, M. A. (2009). In-vitro
whey protein-carboxymethylcellulose complexes. Journal of Food Science, 67, evaluation of hydrocolloid–based encapsulated fish oil. Food Hydrocolloids,
113–119. https://doi.org/10.1111/j.1365-2621.2002.tb11369.x. 23, 1413–1419. https://doi.org/10.1016/j.foodhyd.2008.10.009.
Girard, M., Turgeon, S. L., & Gauthier, S. F. (2002b). Interpolymer Kontopidis, G., Holt, C., & Sawyer, L. (2004). Invited review:
complexing between β-lactoglobulin and low- and hight-methylated pectin Beta-lactoglobulin: Binding properties, structure, and function. Journal of
measured by potentiometric titration and ultrafiltration. Food Hydrocolloids, Dairy Science, 87, 785–96.
16, 585–591. https://doi.org/10.3168/jds.S0022-0302(04)73222-1.
González-Chávez, S. A., Arévalo-Gallegos, S., & Rascón-Cruz, Q. (2009). Koupantsis, T., Pavlidou, E., & Paraskevopoulou, A. (2014). Flavour
Lactoferrin: Structure, function and applications. International Journal of encapsulation in milk proteins – CMC coacervate-type complexes. Food

12 Comprehensive Reviews in Food Science and Food Safety r Vol. 0, 2018 


C 2018 Institute of Food Technologists®
Complexes between WPI and biopolymers . . .

Hydrocolloids, 37, 134–142. https://doi.org/10.1016/j.foodhyd.2013. Orme, C. A., & Giocondi, J. L. (2007). Model systems for formation and
10.031. dissolution of calcium phosphate minerals. In P. Behrens, E. Bauerlein, & S.
Koupantsis, T., Pavlidou, E., & Paraskevopoulou, A. (2016). Glycerol and Mann (Eds.), Handbook of biomineralization: Biomimetic and bio-inspired
tannic acid as applied in the preparation of milk proteins – CMC complex materials chemistry (pp. 135–157). Weinheim: Wiley-VCH.
coavervates for flavour encapsulation. Food Hydrocolloids, 57, 62–71. Overbeek, J. T. G., & Voorn, M. J. (1957). Phase separation in polyelectrolyte
https://doi.org/10.1016/j.foodhyd.2016.01.007. solutions. Theory of complex coacervation. Journal of Cellular Physiology, 49,
Kurukji, D., Norton, I., & Spyropoulos, F. (2016). Fabrication of sub-micron 7–26. https://doi.org/10.1002/jcp.1030490404.
protein-chitosan electrostatic complexes for encapsulation and Öztaş, Yeşim E. R., & Ozguneş, N. (2005). Lactoferrin: A multifunctional
pH-Modulated delivery of model hydrophilic active compounds. Food protein. Advances in Molecular Medicine, 1, 149–154.
Hydrocolloids, 53, 249–260. https://doi.org/10.1016/j.foodhyd.2015.02.021. Park, Y. W., Juárez, M., Ramos, M., & Haenlein, G. F. W. (2007).
Laos, K., Brownsey, G. J., & Ring, S. G. (2007). Interactions between Physico-chemical characteristics of goat and sheep milk. Small Ruminant
furcellaran and the globular proteins bovine serum albumin and Research, 68, 88–113. https://doi.org/10.1016/j.smallrumres.2006.09.013.
β-lactoglobulin. Carbohydrate Polymers, 67, 116–123. Pathak, J., Rawat, K., Aswal, V. K., & Bohidar, H. B. (2016). Hierarchical
https://doi.org/10.1016/j.carbpol.2006.04.021. internal structures in gelatin–bovine serum albumin/β-lactoglobulin gels
Lee, A. C., & Hong, Y. H. (2009). Coacervate formation of and coacervates. Journal of Physical Chemistry B, 120, 9506–12.
α-lactalbumin–chitosan and β-lactoglobulin–chitosan complexes. Food https://doi.org/10.1021/acs.jpcb.6b05378.
Reserach International, 42, 733–738. Pelegrine, D. H. G., & Gasparetto, C. A. (2003). Estudo da solubilidade das
https://doi.org/10.1016/j.foodres.2009.02.022. proteı́nas presentes no soro de leite e na clara de ovo. Revista Brasileira de
Lemon, P. W., Berardi, J. M., & Noreen, E. E. (2002). The role of protein Produtos Agroindustriais, 5, 57–65.
and amino acid supplements in the athlete’s diet: Does type or timing of http://10.15871/1517-8595/rbpa.v5n1p57-65.
ingestion matter? Current Sports Medicine Reports, 4, 214–221. Perozzo, R., Folkers, G., & Scapozza, L. (2004). Thermodynamics of
Lenormand, H., Deschrevel, B., Tranchepain, F., & Vincent, J. C. (2008). protein–ligand interactions: History, presence, and future aspects. Journal of
Electrostatic interactions between hyaluronan and proteins at pH 4: How do Receptors and Signal Transduction, 24, 1–52.
they modulate hyaluronidase activity. Biopolymers, 89, 1088–1103. https://doi.org/10.1081/RRS-120037896.
https://doi.org/10.1002/bip.21061. Phillips, G. O., & Williams, P. A. (2009). Handbook of hydrocolloids.
Li, X., Fang, Y., Al-Assaf, S., Phillips, G. O., Yao, X., Zhang, Y., . . . Jiang, Cambridge, U.K.: Woodhead Publishing Ltd.
F. (2012). Complexation of bovine serum albumin and sugar beet pectin: Raei, M., Shahidi, F., Farhoodi, M., Jafari, S. M., & Rafe, A. (2017).
Structural transitions and phase diagram. Langmuir, 28, 10164–10176. Application of whey protein-pectin nanocomplex carriers for loading of
https://doi.org/10.1021/la302063u. lactoferrin. International Journal of Biological Macromolecules, 105, 281–291.
Liu, J., Shim, Y. Y., Wang, Y., & Reaney, M. J. T. (2015). Intermolecular https://doi.org/10.1016/j.ijbiomac.2017.07.037.
interaction and complex coacervation between bovine serum albumin and Rosenberg, M. (1997). Food treatment forming individual capsule. Google
gum from whole flaxseed (Linum usitatissimum L.). Food Hydrocolloids, 49, Patents
95–103. https://doi.org/10.1016/j.foodhyd.2015.02.035.
Ru, Q., Wang, Y., Lee, J., Ding, Y., & Huang, Q. (2012). Turbidity and
Liu, J., Shim, Y. Y., Shen, J., Wang, Y., & Reaney, M. J. T. (2017). Whey rheological properties of bovine serum albumin/pectin coacervates: Effect
protein isolate and flaxseed (Linum usitatissimum L.) gum electrostatic of salt concentration and initial protein/polysaccharide ratio. Carbohydrate
coacervates: Turbidity and rheology. Food Hydrocolloids, 64, 18–27. Polymers, 88, 838–846. https://doi.org/10.1016/j.carbpol.2012.01.019.
https://doi.org/10.1016/j.foodhyd.2016.10.006.
Santos, M. J., Teixeira, J. A., & Rodrigues, L. R. (2011). Fractionation and
Lowery, L., Edel, J. F., & Mcbride, I. M. (2012). Dietary protein and strength recovery of whey proteins by hydrophobic interaction chromatography.
athletes. Strength & Conditioning Journal, 34, 26–32. Journal of Chromatography B, 879, 475–479.
https://doi.org/10.1519/SSC.0b013e31826284d9. https://doi.org/10.1016/j.jchromb.2011.01.003.
Marshall, K. (2004). Therapeutic applications of whey protein. Alternative Santos, M. B., de Carvalho, C. W. P., & Garcia-Rojas, E. E. (2018).
Medicine Review, 9, 136–156. Heteroprotein complex formation of bovine serum albumin and lysozyme:
McClements, D. J. (2006). Non-covalent interactions between proteins and Structure and thermal stability. Food Hydrocolloids, 74, 267–274.
polysaccharides. Biotechnology Advances, 24, 621–625. https://doi.org/10.1016/j.foodhyd.2017.08.016.
https://doi.org/10.1016/j.biotechadv.2006.07.003. Schmitt, C., Sanchez, C., Desobry-Banon, S., & Hardy, J. (1998). Structure
McClements, D. J., Decker, E. A., Park, Y., & Weiss, J. (2009). Structural and technofunctional properties of protein-polysaccharide complexes: A
design principles for delivery of bioactive components in nutraceuticals and review. Critical Reviews in Food Science and Nutrition, 38, 689–753.
functional foods. Critical Reviews in Food Science and Nutrition, 49, 577–606. https://doi.org/10.1080/10408699891274354.
https://doi.org/10.1080/10408390902841529. Schmitt, C., Aberkane, L., & Sanchez, C. (2009). Protein-polyssacharide
McClements, D. J. (2014). Nanoparticle- and microparticle-based delivery systems: complexes and coacervates In G. O. Phillips & P. A. Williams (Eds.),
Encapsulation, protection and release of active compounds (1st ed). Boca Raton: Handbook of Hydrocolloids Woodhead (pp. 420–476). Cambridge, U.K.:
CRC Press. Publishing Ltd.
Mellema, M. (2004). Complex coacervate encapsulate comprising lipophilic core. Schmitt, C., & Turgeon, S. L. (2011). Protein/polysaccharide complexes and
Google Patents. coacervates in food systems. Advances in Colloid and Interface Science, 167,
Mezzenga, R., & Fischer, P. (2013). The self-assembly, aggregation and phase 63–70. http://org/10.1016/j.cis.2010.10.001.
transitions of food protein systems in one, two, and three dimensions. Report Sgarbieri, V. C. (2004). Propriedades fisiológicas-funcionais das proteı́nas do
on Progress in Physics, 76, 1–43. soro de leite. Revista de Nutrição, 17, 397–409.
https://doi.org/10.1088/0034-4885/76/4/046601. https://doi.org/10.1590/S1415-52732004000400001.
Neemann, F., Rosenberger, S., Jefferson, B., & Mcadam, E. J. (2013). Shahgholian, N., & Rajabzadeh, G. (2015). Fabrication and characterization
Non-covalent protein–polysaccharide interactions and their influence on of curcumin-loaded albumin/gum arabic coacervate. Food Hydrocolloids, 53,
membrane fouling. Journal of Membrane Science, 446, 310–317. 249–260. https://doi.org/10.1016/j.foodhyd.2015.11.031.
https://doi.org/10.1016/j.memsci.2013.06.054. Sinha, R., Radha, C., Prakash, J., & Kaul, P. (2007). Whey protein
Ngarize, S., Adams, A., & Howell, N. A. (2005). Comparative study of heat hydrolysate: Functional properties, nutritional quality and utilization in
and high pressure induced gels of whey and egg albumen proteins and their beverage formulation. Food Chemistry, 101, 1484–1491,
binary mixtures. Food Hydrocolloids, 19, 984–996. https://doi.org/10.1016/j.foodchem.2006.04.021.
https://doi.org/10.1016/j.foodhyd.2004.12.008. Souza, C. J. F., Ramos, A. V., Câmara, P. B. S., Gulão, E. S., de Campos, M.
Nigen, M., Croguennec, T., Renard, D., & Bouhallab, S. (2007). F., & Garcia-Rojas, E. E. (2015). Polymeric complexes obtained from the
Temperature affects the supramolecular structures resulting from interaction of bovine serum albumin and κ-carrageenan. Food Hydrocolloids,
α-lactalbumin−lysozyme interaction. Biochemistry, 46, 1248–1255. 45, 286–290. https://doi.org/10.1016/j.foodhyd.2014.10.023.
https://doi.org/10.1021/bi062129c. Sperber, B. L. H. M., Stuart, M. A. C., Schols, H. A., Voragen, A. G. J., &
Norton, J. E., Espinosa, Y. G., Watson, R. L., Spyropoulos, F., & Norton, Norde, W. (2010). Overall charge and local charge density of pectin
I. T. (2015). Functional food microstructures for macronutrient release and determines the enthalpic and entropic contributions to complexation with
delivery. Food & Function, 6, 663–678. β-lactoglobulin. Biomacromolecules, 11(12)3578–3583.
https://doi.org/10.1039/C4FO00965G. https://doi.org/10.1021/bm1010432.


C 2018 Institute of Food Technologists® Vol. 0, 2018 r Comprehensive Reviews in Food Science and Food Safety 13
Complexes between WPI and biopolymers . . .

Stănciuc, N., Turturică, M., Oancea, A. M., Barbu, V., Ioniţă, E., Aprodu, I., Wally, J., & Buchnan, S. K. (2007). A structural comparison of human serum
& Râpeanu, G. (2017). Microencapsulation of anthocyanins from grape transferrin and human lactoferrin. Biometals, 20, 249–262.
skins by whey protein isolates and different polymers. Food and Bioprocess https://doi.org/10.1007/s10534-006-9062-7.
Technology, 10, 1715–1726. https://doi.org/10.1007/s11947-017- Walzem, R. L., Dillard, C. J., & German, J. B. (2002). Whey components:
1938-8. Millennia of evolution create functionalities for mammalian nutrition: What
Steijns, J. M., & Hooijdonk, A. C. (2000). Occurrence, structure, we know and what we may be overlooking. Critical Reviews in Food Science
biochemical properties and technological characteristics of lactoferrin. and Nutrition, 42, 353–375. https://doi.org/10.1080/10408690290825574.
British Journal of Nutrition, 84, S11–7. Wang, X., Wang, Y. W., Ruengruglikit, C., & Huang, Q. (2007). Effects of
https://doi.org/10.1017/S0007114500002191. salt concentration on formation and dissociation of β-lactoglobulin/pectin
Stojadinovic, M., Burazer, L., Ercili-Cura, D., Sancho, A., Buchert, J., complexes. Journal of Agricultural and Food Chemistry, 55, 10432–10436.
Cirkovic Velickovic, T., & Stanic-Vucinic, D. (2012). One-step method for https://doi.org/10.1021/jf071787g.
isolation and purification of native β-lactoglobulin from bovine whey. Wang, Q., Lei, H., Jiang, L., Fu, J., Liu, Y., Wen, Q., & Zhong, Y. (2014).
Journal of the Science of Food and Agriculture, 92, 1432–1440. Optimization and evaluation of microencapsulation of star anise oleoresin.
https://doi.org/10.1002/jsfa.4722. Journal of Food Processing and Preservation, 38, 2129–2136.
Tainaka, K. I. (1980). Effect of counterions on complex coacervation. https://doi.org/10.1111/jfpp.12193.
Biopolymers, 19, 1289–1298. https://doi.org/10.1002/bip.1980.360190705. Wang, B., Blanch, E., Barro, C. J., & Adhikari, B. (2017). Preparation and
Tainaka, K. (1979). Study of complex coacervation in low concentration by study of digestion behavior of lactoferrin-sodium alginate complex
virial expansion method. i. salt free systems. Journal of the Physical Society of coacervates. Journal of Functional Foods, 37, 97–106.
Japan, 46, 1899–1906. https://doi.org/10.1143/JPSJ.46.1899. https://doi.org/10.1016/j.jff.2017.07.044.
Tang, J. E., Moore, D. R., Kujbida, G. W., Tarnopolsky, M. A., & Phillips, S. Wee, M. S. M., Nurhazwani, S., Tan, K. W. J., Goh, K. K. T., Sims, I. M., &
M. (2009). Ingestion of whey hydrolysate, casein, or soy protein isolate: Matia-Merino, L. (2014). Complex coacervation of an
Effects on mixed muscle protein synthesis at rest and following resistance arabinogalactan-protein extracted from the Meryta sinclarii tree (puka gum)
exercise in young men. Journal of Applied Physiology, 107, 987–992. and whey protein isolate. Food Hydrocolloids, 42, 130–138.
https://doi.org/10.1152/japplphysiol.00076.2009. https://doi.org/10.1016/j.foodhyd.2014.03.005.
Tavares, G. M., Croguennec, T., Hamon, P., Carvalho, A. F., & Bouhallab, S. Weinbreck, F. C. J., Kruif, C. G., & Schrooyen, P. (2003a). Complex coacervates
(2015). Selective coacervation between lactoferrin and the two isoforms of containing whey proteins. Patents WO03106014 and EP1371410.
β-lactoglobulin. Food Hydrocolloids, 48, 238–247. Weinbreck, F., de Vries, R., Schrooyen, P., & de Kruif, C. G. (2003b).
https://doi.org/10.1016/j.foodhyd.2015.02.027. Complex coacervation of whey proteins and gum arabic. Biomacromolecules,
Tiebackx, F. W. Z. (1911). Gleichzeitige Ausflockung zweier Kolloide. 4, 293–303. https://doi.org/10.1021/bm025667n.
Kolloide Zeitschrift, 8, 198–201. Weinbreck, F., Tromp, R. H., & de Kruif, C. G. (2004a). Composition and
Thongkaew, C., Hinrichs, J., Gibis, M., & Weiss, J. (2015). Sequential structure of whey protein/gum arabic coacervates. Biomacromolecules, 5,
modulation of pH and ionic strength in phase separated whey protein isolate 1437–1445. https://doi.org/10.1021/bm049970v.
– Pectin dispersions: Effect on structural organization. Food Hydrocolloids, Weinbreck, F., Minor, M., & de Kruif, C. G. (2004b). Microencapsulation of
47, 21–31. https://doi.org/10.1016/j.foodhyd.2014.11.006. oils using whey protein/gum arabic coacervates. Journal of Microencapsulation,
Turgeon, S. L., Beaulieu, M., Schmitt, C., & Sanchez, C. (2003). 21, 667–679. https://doi.org/10.1080/02652040400008499.
Protein–polysaccharide interactions: Phase-ordering kinetics, Xu, A. Y., Melton, L. D., Jameson, G. B., Williams, M. A. K., &
thermodynamic and structural aspects. Current Opinion in Colloid & Interface McGillivray, D. J. (2015). Structural mechanism of complex assemblies:
Science, 8, 401–414. https://doi.org/10.1016/S1359-0294(03)00093-1. Characterisation of beta-lactoglobulin and pectin interactions. Soft Matter,
Turgeon, S. L., Schmitt, C., & Sanchez, C. (2007). Protein–polysaccharide 11, 6790–6799. https://doi.org/10.1039/C5SM01378J.
complexes and coacervates. Current Opinion in Colloid & Interface Science, 12, Yada, R. Y. (2004). Protein in food processing (1st ed). New York: CRC Press.
166–178.
Yan, Y., Kizilay, E., Seeman, D., Flanagan, S., Dubin, P. L., Bovetto, L., . . .
Tolstoguzov, V. B. (1991). Functional properties of food proteins and role of Schmitt, C. (2013). Heteroprotein complex coacervation: Bovine
protein-polysaccharide interaction. Food Hydrocolloids, 4, 429–468. β-lactoglobulin and lactoferrin. Langmuir, 29, 15614–15623.
https://doi.org/10.1016/S0268-005X(09)80196-3. https://doi.org/10.1021/la4027464.
Tolstoguzov, V. B. (2007). Ingredients interactions in complex foods: Zeeb, B., Mi-Yeon, L., Gibis, M., & Weiss, J. (2018). Growth phenomena in
Aggregation and phase separation. In D. J. McClements (Ed.), Understanding biopolymer complexes composed of heated WPI and pectin. Food
and controlling the microstructure of complex foods (pp. 185–206). Cambridge: Hydrocolloids, 74, 53–61. https://doi.org/10.1016/j.foodhyd.2017.07.026.
Wood Head Publishing.
Zhang, Z. Q., Pan, C. H., & Chung, D. (2011). Tannic acid cross-linked
Urista, M. C., Álvarez, F. R., Riera, R. F., Cuenca, A. A., & Téllez, J. A. gelatin–gum arabic coacervate microspheres for sustained release of allyl
(2011). Review: Production and functionality of active peptides from milk. isothiocyanate: Characterization and in vitro release study. Food Research
Food Science and Technology, 4, 293–317. International, 44, 1000–1007.
https://doi.org/10.1177/1082013211398801. https://doi.org/10.1016/j.foodres.2011.02.044.
Vidigal, M. C. T. R., Minim, V. P. R., Ramos, A. M., Ceresino, E. B., Zhang, N., & Mutilangi, W. (2013). Complex coacervates, methods and food
Diniz, M. D. M. S., Camilloto, G. P., & Minim, L. A. (2012). Effect of products. US patent 20130004640 A1.
whey protein concentrate on texture of fat-free desserts: Sensory and
instrumental measurements. Food Science and Technology, 32, 412–418. Zhang, S., Zhong, Z., & Vardhanabhuti, B. (2014). Effect of charge density
https://doi.org/10.1590/S0101-20612012005000047. of polysaccharides on self-assembled intragastric gelation of whey
protein/polysaccharide under simulated gastric conditions. Food & Function,
Voswinkel, L., & Kulozik, U. (2014). Fractionation of all major and minor 5, 1829–1838. http://10.1039/C4FO00019F.
whey proteins with radial flow membrane adsorption chromatography at lab
and pilot scale. International Dairy Journal, 39, 209–214. Zuvanov, V. C., Garcia-Rojas, E. E., Souza, C. J., de Gulão, E. S., & Pereira,
https://doi.org/10.1016/j.idairyj.2014.06.012. L. J. B. (2014). Obtaining process of interpolymeric complexes from
lactalbumin, xanthan gum and pectin. Brazilian Journal of Food Technology,
Walstra, P., Wouters, J. T. M., & Geurts, T. J. (2006). Dairy science and 17, 213–220. https://doi.org/10.1590/1981-6723.0814.
technology (2nd ed). Taylor & Francis Group; New York, EUA

14 Comprehensive Reviews in Food Science and Food Safety r Vol. 0, 2018 


C 2018 Institute of Food Technologists®

You might also like