You are on page 1of 17

Journal of Functional Foods 42 (2018) 58–74

Contents lists available at ScienceDirect

Journal of Functional Foods


journal homepage: www.elsevier.com/locate/jff

Whey protein hydrolysates as a source of bioactive peptides for functional T


foods – Biotechnological facilitation of industrial scale-up

Anja Dulliusa, Márcia Inês Goettertb, Claucia Fernanda Volken de Souzaa,
a
Laboratório de Biotecnologia de Alimentos, Programa de Pós-Graduação em Biotecnologia, Universidade do Vale do Taquari - Univates, Lajeado, RS, Brazil
b
Laboratório de Cultura de Células, Programa de Pós-Graduação em Biotecnologia, Universidade do Vale do Taquari - Univates, Lajeado, RS, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: Whey proteins, which possess the highest nutritional quality of all food proteins, are an optimal source of
Functional foods functional food ingredients. Enzymatic hydrolysis of whey proteins liberates fragments that can promote health
Bioactive peptides benefits in the immune, cardiovascular, nervous and gastrointestinal systems. Industrial production of peptide-
Whey proteins based food ingredients requires overcoming several challenges in product development to achieve economically
Downstream process
viable downstream processes. We suggest hybrid strategies currently utilized in biopharmacy that are based on
Industrial scale-up
computational modelling combined with heuristics and mechanistic modelling, thus minimizing the time- and
Peptide enrichment techniques
cost-intensive trial and error approach. Additionally, we propose an application of a cost performance indicator
during early decisions at the laboratory scale, which can lead to optimal downstream processing of bioactive,
peptide-based food ingredients. This review summarizes the requirements of industrial processes regarding
peptide release and stability, depending on several process parameters, and considers some enrichment tech-
niques for whey-derived peptides that are potentially applicable to industry.

1. Introduction Maeda-Yamamoto (in press) provided circumstantial information about


the necessary documentation for the approval of products as FOSHU by
Functional foods (also called nutraceuticals, pharmafoods and de- the Japanese Office of Health Policy on Newly Developed Foods,
signer foods) are made from natural ingredients and offer, in addition to highlighting that legal recognition as FOSHU requires a detailed review
their nutritional value, a specific health advantage due to the functional of scientific evidence, including human clinical studies, animal studies,
in vivo activity of a supplemented food additive (McIntosh et al., 1998). and in vitro metabolic and biochemical data.
Japan was the first country to introduce so-called Foods for Specified Food proteins are suggested to be a high-quality source of protein
Health Uses (FOSHU) into the market in 1991 (Yamada, Sato-Mito, because they are a natural ingredient in functional foods. Typical food
Nagata, & Umegaki, 2008). These types of functional foods are divided proteins derived from plant and animals are meat, milk, egg, fish, soy,
into three categories: (1) standardized FOSHU, (2) reduction of disease and wheat, amongst others. Whey proteins have the highest nutritional
risk FOSHU, and (3) qualified FOSHU. Currently, these categories are quality (Protein Digestibility-Corrected Amino Acid Score
based on evidence of eleven different health claims, e.g., gastro- (PDCAAS) = 1) of the dietary protein sources and even greater biolo-
intestinal regulation, lower blood cholesterol, glucose level reduction or gical value than the milk protein casein. Whey proteins are rich in
dental health (Maeda-Yamamoto, in press). Yamada et al. (2008) and human-essential amino acids (AAs), branched-chain amino acids

Abbreviations: FOSHU, Foods for Specified Health Uses; PDCAAS, protein digestibility corrected amino acid score; BCAAs, branched-chain amino acids; β-LG, β-lactoglobulin; TMP, total
milk protein; α-LA, α-lactalbumin; BSA, bovine serum albumin; BLF, bovine lactoferrin; GMP, glycomacropeptides; IgG, immunoglobulin G; BOD, biological oxygen demand; COD,
chemical oxygen demand; ACE, angiotensin-converting enzyme; DH, degree of hydrolysis; WPI, whey protein isolate; WPC, whey protein concentrate; BLP, Bacillus licheniformis protease;
UHPH, ultrahigh-pressure homogenization; STTT, short-time thermal treatment; US, ultrasonication; MW, microwave; VIP, virtual intermediate peptide; GIT, gastrointestinal tract; CPP,
critical process parameters; DPP-IV, dipeptidylpeptidase IV; BU, binding unit; SU, stimulating unit; SUMO, small ubiquitin-related modifier; LAB, lactic acid bacteria; NSLAB, non-starter
lactic acid bacteria; CCP, corn cob powder; PES, polyethersulfone; HPLC, high-pressure liquid chromatography; MS, mass spectrometry; FDA, Food and Drug Administration; AAs, amino
acids; ACPs, anticancer peptides; NK cells, natural killer cells; DoE, Design of Experiments; QSAR, quantitative structureactivity relationships; AMP, antimicrobial peptide; SVM, support
vector machines; RF, random forest; ANN, artificial neuronal networks; DA, discriminant analysis; MBPD, Milk Bioactive Peptide Database; CAMP, Collection of Anti-Microbial Peptides;
DPP-4, dipeptidyl peptidase-4; PPI, purification performance index; SCI, separation cost indicator; QSPR, quantitative structureproperty relations; SDS, sodium dodecyl sulphate; AC,
activated carbon; EDUF, electrodialysis-ultrafiltration; CFEMF, cross-flow electro-membrane filtration; MC, membrane chromatography; MAC, membrane adsorption chromatography;
MWCO, molecular weight cut-off; IMER, immobilized enzyme reactors; EOPO Co-polymer (UCON)/Phosphate, poly(ethylene glycol-ran-propylene glycol) monobutyl ether/phosphate

Corresponding author.
E-mail address: claucia@univates.br (C.F.V. de Souza).

https://doi.org/10.1016/j.jff.2017.12.063
Received 16 August 2017; Received in revised form 27 December 2017; Accepted 28 December 2017
Available online 10 January 2018
1756-4646/ © 2017 Elsevier Ltd. All rights reserved.
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

(BCAAs) and sulfur-containing AAs, and they promote metabolic reg- contains only 30% of the energy content of sucrose), could result in the
ulation and protein folding (Smithers, 2015). high-grade transformation of the effluent whey (Jayamuthunagai,
The health benefits of whey proteins may have been known for Srisowmeya, Chakravarthy, & Gautam, 2017).
hundreds of years or even longer, as whey was part of treatments for The elevated nutritional value of bovine whey proteins, together
illness in folk medicine (Hoffmann, 1961) and is still used by various with their technological properties and distinguished health benefits in
societies for food preservation and illness treatment (Gupta, 2012; the form of hydrolysed peptides, have transformed whey into a high-
Pieroni & Giusti, 2008; Pieroni & Gray, 2008; Sikarwar & Kaushik, value raw material for the food industry with particular applications in
1993). In the period from the 17th to the early 19th century, whey the production of functional foods (Yadav et al., 2015). Several recent
became popular as a medicinal drink and an inexpensive alternative studies have confirmed the applicability of health-promoting peptides
bath ingredient (Smithers, 2015). to functional foodstuffs: Chatterjee, Kanawjia, and Khetra (2016)
The precise molecular components of whey are known. In addition showed that the addition of a tryptic whey protein hydrolysate to In-
to lactose, minerals and vitamins, the health benefits of whey are re- dian sweetened yogurt significantly increased its angiotensin-con-
lated to whey proteins, including enzymes and glycomacropeptide verting enzyme (ACE) inhibition and antioxidant activity, indicating
(GMP) (Anand, Som Nath, & Chenchaiah, 2013). The main components the maintenance of peptide stability during food processing and within
of bovine whey proteins are β-lactoglobulin (β-LG, 3.3 g/L total milk the food matrix. Yu, Amorim, Marques, Calhau, and Pintado (2016)
protein (TMP)) and α-lactalbumin (α-LA, 1.2 g/L TMP), and compo- reported as a lead for peptide functionality and stability that a low-
nents in minor concentrations include bovine serum albumin (BSA, molecular-weight (< 1 kDa) whey peptide extract of a mixture of cow,
0.3 g/L TMP), bovine lactoferrin (BLF, 0.1 g/L TMP), immunoglobulins goat and ewe whey stimulated the growth of probiotic bacteria (Lac-
(0.5–1 g/L TMP), lactoperoxidase (0.03 g/L TMP), and proteinaceous tobacillus acidophilus and Bifidobacterium animalis) in Wistar rats. Graves
GMP (1.2 g/L TMP) (Korhonen, 2009a; Korhonen, 2012; Madureira, et al. (2016) found that an anticancer pentapeptide derived from rice
Tavares, Gomes, Pintado, & Malcata, 2010). The enzymatic action of bran was stable for 6 months when stored in spray-dried orange juice
chymosin releases proteinaceous GMP from milk κ-casein during and received acceptance from consumers regarding its sensory prop-
cheese-making, cleaving the C-terminal region of κ-casein between AA erties. Hafeez et al. (2014) proposed the use of bioactive peptides from
residues 105 and 106 and therefore leading to the precipitation of milk proteins, including whey proteins, to increase the functionality of
caseins, thus forming cheese (Thomä-Worringer, Sørensen, & López- fermented milk products and reviewed the three known production
Fandiño, 2006). GMP is a glycophosphopeptide that is 64 AAs in length process strategies: (a) peptide release directly into fermented milk
and lacks the aromatic AAs Phe, Tyr and Trp; for this reason, it is re- products, (b) supplementation of fermented milk products with pep-
commended in medical food design (Ney & Etzel, 2017) and showed tides produced outside of the product and (c) production of bioactive
prebiotic and anti-inflammatory effects in mice (Sawin et al., 2015). peptides using recombinant DNA technology.
Yadav et al. (2015) noted that in addition to their chemical function Generally, natural digestion of whey proteins in the gastrointestinal
(e.g., aiding emulsification or gelling) and nutraceutical characteristics, tract via the consumption of whey protein-containing products is not
whey proteins have known physiological benefits, including vitamin sufficient to achieve a positive health effect; as a consequence, the
binding, antibacterial activity and immunomodulatory effects, amongst bioactive peptides responsible for the health benefits must be enriched
others. by industrial processing of hydrolysed whey (Gauthier, Pouliot, & Saint-
Whey proteins with in vivo biological functions, such as positive Sauveur, 2006). The ongoing development in the past two decades of
influences on the cardiovascular, digestive, endocrine, immune, and technologies for the production of whey bioactive peptide-containing
nervous systems, are therefore optimal for use as a functional food foods on the industrial scale has resulted in several products on the
additive. However, many of the bioactivities of whey proteins are en- international market (Hettiarachchy, Sato, Marshall, & Kannan, 2011;
crypted within their native protein sequences and can thus be liberated Korhonen, 2009b). Table 1 lists bioactive peptides that are derived from
only by protein fragmentation, e.g., through enzymatic hydrolysis whey proteins and are used as ingredients in functional foods. Re-
(Pihlanto-Leppälä, 2000). The natural production of these bioactive garding the industrial production of peptide-based food ingredients,
whey peptides occurs via (I) gastrointestinal digestion of milk proteins membrane filtration techniques (ultrafiltration and nanofiltration) are
by digestive enzymes and (II) food processes, such as the fermentation the only methods that enrich food ingredients, such as whey hydro-
of milk or the ripening of cheese by lactic acid bacteria (LAB) lysates, with bioactive peptides in an economically viable way (Agyei,
(Korhonen, 2009b). Ongkudon, Wei, Chan, & Danquah, 2016; Korhonen, 2009b). More ef-
A problematic aspect of whey is its increasing volume as a by-pro- fective and specific enrichment of bioactive peptides, including pur-
duct, which became detrimental to the environment with in- ification, within downstream processes requires chromatography steps,
dustrialization and expansion of the manufacture of dairy products in which are cost intensive and time consuming when applied using
the 20th century (Smithers, 2015). For the production of 1 kg of cheese, conventional laboratory methods. These processes are expensive, dif-
approximately 9 kg of cheese whey accrues as an effluent (Prazeres, ficult to apply or affect the secondary structure of peptides, resulting in
Carvalho, & Rivas, 2012). Bovine whey contains 20% of the milk pro- the possible alteration or elimination of their bioactive characteristics;
tein content, the entire milk lactose content (∼5%), and traces of fat therefore, industrial scale-up is not economically viable for the food
(0.1%) and mineral salts (0.46–10%). Approximately 99% of this or- industry. Thus, the lack of economically profitable production processes
ganic matter is biodegradable, resulting in a high biochemical oxygen decelerates the development of products that contain bioactive peptides
demand (BOD; > 30.000 mg O2/L for sweet whey and 35.000 mg O2/L (Agyei & Danquah, 2011; Agyei et al., 2016). Furthermore, basic re-
for acid whey) and chemical oxygen demand (COD; > 60.000 mg O2/L search purification processes must be adapted to meet the requirements
for sweet whey and ∼80.000 mg O2/L for acid whey). Biological of an economically viable process at the industrial scale (Korhonen &
treatment of whey is difficult because of its acidic pH (3.8–6.5), low Marnila, 2013). New purification technologies in the food industry are
alkalinity, and sodium, free ammonia, potassium and volatile fatty acid less focused on products with high-grade, pure bioactive peptides and
contents (Prazeres et al., 2012; Smithers, 2015). Yadav et al. (2015) are more focused on the development of economically viable enrich-
reviewed possible methods to transform the proteinaceous part of ef- ment techniques (Agyei et al., 2016).
fluent whey into bioprotein, functional proteins and bioactive peptides; This review article provides a brief overview of the conventional
however, as mentioned before, nearly the entire lactose content of whey downstream process steps of bioactive whey peptide research devel-
remains in the whey permeate and thus requires further effluent post- oped under laboratory conditions compared to the actual flow path of
treatment steps. Innovative treatment technologies, such as the pro- industrial production. Furthermore, we discuss current critical process
duction of D-tagatose from lactose as an alternate sweetener (D-tagatose steps that can affect the release of bioactive peptide sequences and

59
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

Table 1
Commercially sold functional food ingredients based on bioactive peptides derived from whey proteins (Data from Ballard et al., 2009; Korhonen, 2009b; Lacroix et al., 2016).

Product name Manufacturer Type of product Peptide bioactivity Health claim

BioZate® product line Davisco, USA Hydrolysed whey protein isolate β-lactoglobulin fragments Reduction of blood pressure
BioPureGMP™ Davisco, USA Whey protein isolate (WPI) GMP f(106–169) Prevention of dental caries, blood clotting,
antibacterial, antivirus
Vivinal® ALPHA Borculo Domo Ingredients (BDI), Ingredient/hydrolysate Whey-derived peptide Relaxation and sleep
the Netherlands
Praventin™ DMV International, the Food supplement/capsule Lactoferrin enriched whey Acne reduction
Netherlands protein hydrolysate
Dermylex™ Advitec, Inc., Canada Food supplement/ tablet Whey protein extract XP-828L Reduces symptoms of psoriasis
Hilmar™8390 Hilmar Ingredients, USA Food supplement Whey protein hydrolysate ACE inhibition, DPP-IV enzyme inhibition
NOP-47™ Glanbia Nutritionals, USA Food supplement and Whey derived peptide Anti-inflammatory properties
pharmaceutical applications

Abbreviations: GMP: glycomacropeptide, ACE: angiotensin-converting enzyme, DPP-IV: dipeptidylpeptidase IV.

product quality and thus remain a challenge to industrial bioactive 2.1. Effects of pre-treatment on peptide bioactivity
peptide production. Additionally, we briefly review previously pro-
posed modified basic research processes that include in silico biopros- The main issue in protein hydrolysis is that well-defined secondary
pecting of peptide release and activity and have thus already led to a structures in proteins, such as α-helices, are less susceptible to pro-
more systematic research process. Within this review, we propose an teolysis and thus lead to a low degree of hydrolysis (DH) (Tsai, de
integration of complementary processes enabling the simultaneous Laureto, Fontana, & Nussinov, 2002). Pre-treatment of proteins with
production of bioactive peptide products, thus enhancing the efficiency heat or high pressure led α-helices to refold into β-sheets or random
of the industrial process to optimize the value of whey as a raw mate- coils (Ngarize, Herman, Adams, & Howell, 2004), which improves en-
rial. This technique requires a more methodical design of the processing doprotease access and hydrolysis (Fernández & Riera, 2013). It is well
pathway based on hybrid methods, including in silico methods, me- known that alterations in environmental conditions applied for pre-
chanistic modelling and heuristics, as currently applied in the bio- treatment, such as pH, salts, solvents, pressure and temperature, cause
pharmaceutical industry. Finally, we review enrichment methods with protein denaturation; moreover, temperature is a favoured tool for
a particular emphasis on the AA profiles of whey-derived bioactive protein unfolding (Tavano, 2013). Cheison and Kulozik (2017) de-
peptides, which are promising for economically viable industrial scale- scribed the heat-denaturation of whey proteins as a two-step process,
up and the consequent more targeted production of whey hydrolysate wherein unfolding firstly occurs at low denaturing temperatures
products. (characterized as the molten globule state), followed by the aggregation
step. Therefore, thermal protein treatment can enhance the exposure of
a peptide to the solvent, thus facilitating peptide release on the one
2. Conventional downstream process and its challenges in hand. On the other hand, peptide release can also be hindered, de-
bioactive peptide production pending on the position of the peptide within the protein, in-
tramolecular disulfide exchange reactions due to conformational
The common in vitro bottom-up proteomics approach is divided into changes in non-native monomer formation, or positioning of the pep-
several steps (Fig. 1A): (I) isolation of proteins or a protein mixture tide within aggregation sites. For example, the release of the ACE-in-
from a food matrix; (II) hydrolysis of proteins by different enzymatic hibiting, bioactive peptide f (Lys9-Lys14) derived from β-LG by trypsin
proteases, namely, (a) gastrointestinal enzymes such as pepsin, trypsin (E.C. 3.4.21.4) was hindered by the formation of aggregates at acidic
and chymotrypsin, (b) microorganism or plant proteases such as alca- pH (pH 5.1) and non-native monomers at neutral and alkaline pH (pH
lase, bromelain and ficin, and (c) the enzymes present in microbial 6.8 and 8.0) until a DH of 5% was reached (Leeb, Götz, Letzel, Cheison,
fermentation by starter LAB, non-starter LAB (NSLAB), or several yeast & Kulozik, 2015). Generally, the denaturation pattern is influenced by
strains used in cheese-making; (III) preliminary bioactivity screening; the composition of whey proteins (e.g., whey protein isolate (WPI),
(IV) separation and purification of bioactive peptides, combined with whey protein concentrate (WPC) or pure protein) and environmental
intermediate bioactivity tests; (V) identification of interesting bioactive factors, as ionic strength, pH or lactose concentration, as described and
peptides; and (VI) confirmation of bioactivity by de novo peptide reviewed in detail by Cheison and Kulozik (2017).
synthesis and bioactivity tests in vitro or/and in vivo (Agyei et al., 2016; Proteolysis may also be enhanced by high pressure. β-LG, which is
Capriotti, Cavaliere, Piovesana, Samperi, & Laganà, 2016; Li-Chan, resistant to peptic proteolysis under atmospheric pressure (0.1 MPa),
2015). released high levels of peptides under high-pressure treatment
The industrial production of whey protein hydrolysates enriched (300 MPa) (Chobert et al., 1996). Blayo, Vidcoq, Lazennec, and Dumay
with bioactive peptides (Fig. 1B) includes the following processing (2016) showed that high pressurization (UHPH, ultrahigh-pressure
steps: (1) fractionation of total solids to separate whey proteins from homogenization, i.e., 300 MPa of pressure followed by immediate
fat, lactose and minerals; (2) enzymatic hydrolysis of whey proteins by cooling) or thermal treatment (STTT, short-time thermal treatment,
(a) the use of commercial enzymes from animal sources, such as pan- exposure to 75 °C for 10 s) of whey proteins that were enriched in β-LG
creatin, trypsin, and pepsin, (b) bacterial and fungal proteases from and α-LA (∼68.5 and ∼21.5%, respectively, with other whey con-
microbial fermentation, or (c) plant enzymes such as papain and bro- stituents present in trace amounts) led to the highest proteolysis yields.
melain; (3) post-enzymatic treatment, e.g., filtration, centrifugation and In contrast, Bamdad, Bark, Kwon, Suh, and Sunwoo (2017) showed that
chromatographic steps; (4) spray-drying of the product; and (5) product high pressurization at 100 MPa led to a decrease in the DH to β-LG
packaging (Pasupuleti & Braun, 2008). These workflows (A and B in hydrolysis using trypsin, whereas the DH increased by the use of the
Fig. 1) include crucial industrial processing steps that can alter or, in enzymes neutrase, alcalase, savinase, elastase and thermolysin com-
the worst case, destroy the beneficial bioactivity of the peptide. These pared to hydrolysis under atmospheric pressure. Moreover, the anti-
steps are (a) the pre-treatment steps before protein hydrolysis, (b) the oxidant and immunomodulatory activities of alcalase-hydrolysed β-LG
protein hydrolysis per se, (c) the purification of bioactive peptides, and were improved. Hydrolysis of BSA with trypsin and chymotrypsin under
(d) the spray-drying step (Agyei et al., 2016). high pressure at 400 MPa for 25 min also resulted in an improved DH of

60
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

Fig. 1. (A) Conventional downstream process in bioactive peptide research with the “trial and error” approach. (B) Industrial process scheme of whey protein hydrolysates that are
enriched with bioactive peptides.

7% and 10%, respectively, whereas high pressurization above 400 MPa enzymatic hydrolysis. These examples indicate that different pre-
led to the formation of aggregates and thus a decrease in DH (De Maria, treatment methods have different effects on protein structure during
Ferrari, & Maresca, 2017). These results show that beyond improving denaturation. Pre-treatment methods can favour desired protein release
proteolysis activity, the use of high pressurization with simultaneous or prolong it, as demonstrated by Leeb, Kulozik, and Cheison (2011)
enzymatic hydrolysis can reduce industrial production costs as it and Leeb et al. (2015), who applied thermal heat at 80 °C as a pre-
eliminates protein pre-treatment as a separate process step. treatment tool. In these works, different pH values caused different
Other pre-treatment methods that do not employ heat or high particle sizes of β-LG, which signalled diverging alterations of its ter-
pressure include extrusion, sonication, microwave and pulsed electric tiary structure due to intramolecular disulfide bridge exchange reac-
field, as mentioned by Agyei et al. (2016) and Uluko et al. (2015), but tions or aggregation, resulting in decreased variability of the peptide
are not further discussed in this review. pattern due to lower trypsin (E.C. 3.4.21.4) susceptibility of some
Pre-treatment of proteins before their exposure to enzymatic hy- cleavage sites. These findings enable the design of a more controlled
drolysis does not necessarily destroy their functional activity. When α- process to release desired peptide sequences. Moreover, Delahaije,
lactalbumin was heated, its antibacterial activity increased (Expósito & Gruppen, van Eijk- van Boxtel, Cornacchia, and Wierenga (2016) re-
Recio, 2006). Arrutia, Puente, Riera, Menéndez, and González (2016) ported that the extent of aggregation could be directed by controlling
showed that heat pre-treatment of BSA for 1 h over a temperature range the level of protein unfolding. While increased protein concentration
of 65–75 °C increased the DH and thus resulted in a greater number of led to a higher concentration of aggregates, a shift from pH 7 to pH 8
bioactive peptides. The degree of BSA denaturation was not measured increased the native-like protein content after heat pre-treatment at
in that study, but Leeb et al. (2015) demonstrated the contrary results 80 °C. It was also found that non-native β-LG consists of two types of
that the hydrolysis of β-LG that had been denaturated close to 100% led conformations, one of which has an identical structure to native β-LG.
to reduced DH values, indicating that the degree of denaturation seems Increased temperature led to a shift from a native-like to the non-native
to be an important factor influencing the hydrolysis rate and reaction conformation, whereas pH, ionic strength and protein concentration
time. Uluko et al. (2015) compared the antioxidant activity of untreated had no influence on the ratio between native-like and non-native-like β-
milk protein hydrolysates derived from milk protein concentrates to LG states.
that of samples pre-treated with ultrasonication (US), microwave (MW) As a disadvantage of protein pre-treatments, when the secondary
or heat and combinations thereof (US + MW; US + heat and structure changes from α-helix to β-sheet, aggregates of whey peptides
US + MW + heat). US led to a higher release of antioxidant peptides are commonly expected to be formed because of their hydrophobic AA
compared to that of the control, whereas MW and heat pre-treatment content, leading to an abrogation of bioactivity (Bulone, Martorana, &
lowered the radical-scavenging activity of the hydrolysates. Combina- San Biagio, 2001). Hydrolysis of whey proteins can produce amphi-
tions of the pre-treatment methods did not significantly improve the philic peptides, whose self-assembly into micelles is guided by β-sheet
antioxidant activity compared to the single pre-treatments. Adjonu, interactions (Cui, Webber, & Stupp, 2010). Creusot and Gruppen (2008)
Doran, Torley, and Agboola (2013) also reported that heat pre-treat- showed that peptide aggregation within a WPI that had been hydro-
ment did not improve the antioxidant activity of whey protein hydro- lysed with a Bacillus licheniformis protease (BLP, E.C. 3.4.21.19) in-
lysates. Wali et al. (2017) reported higher ACE-inhibition activity using creased as DH increased. Under certain pH conditions, self-assembled
multifrequency ultrasound pre-treatment of rapeseed protein before peptide aggregation can be reversible as a result of different charge

61
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

distributions within the peptide (Aggeli et al., 2003). However, the that the mechanism of hydrolysis strongly depended on pH, probably as
extensive hydrolysis (> 18%) of whey proteins with Alcalase® 2.4L a result of changes in the charge state at the active site of the enzyme or
(E.C. 3.4.21.62) led to irreversible peptide aggregation and gel forma- the presence of negative charges in the P1′ position. Therefore, de-
tion. Peptide aggregates were generally formed by low-molecular- pending on the pH, the selectivity for some cleavage sites varied from
weight peptides (below 2 kDa) with a net zero charge in the pH range 0% to 27%. Vorob’ev, Butré, Sforza, Wierenga, and Gruppen (2017)
from 4 to 7 (Doucet, Otter, Gauthier, & Foegeding, 2003). Cheison, verified that the release of hardly accessible cleavage sites for the en-
Wang, and Xu (2007) reported that WPI hydrolysates generated with zyme BLP within intact β-LG could be described by a two-step de-
Alcalase® 2.4L and Proteinase N (IUB 3.4.24.28) were mostly composed masking model following a non-monotonous increasing function, which
of peptides with a molecular mass below 1 kDa that reached a DH of means the formation of a virtual intermediate peptide (VIP) from which
15% without any peptide aggregation. the final peptides can be formed. Thus, once a cleavage site is de-
masked, its hydrolysis rate is similar to other cleavage sites and follows
2.2. Roles of degree of hydrolysis and peptide distribution in the protein first-order reaction kinetics.
hydrolysis step Cheison, Leeb, Letzel et al. (2011) showed that buffer type and
concentration also influenced peptide release: Tris-HCl- or NaOH-buf-
The DH refers to the percentage of peptide bonds cleaved during the fered water resulted in similar peptide compositions derived from β-LG,
hydrolysis process (Adler-Nissen, 1979). Consequently, the DH de- whereas potassium phosphate-buffered hydrolysis released a different
termines the desired end product (specifically, the peptide chain peptide pattern at higher buffer concentrations, causing dimerisation of
length), which in turn influences peptide function. Thus, in the food the β-LG and peptide bond resistance in specific regions. Le Maux,
industry, a low DH (< 10%) is used to produce peptides with useful Nongonierma, and FitzGerald (2017) tried to reach defined DH values
physical functions, such as foaming, gelation or emulsification, whereas of 9% and 13% by hydrolysing β-LG with elastase at a constant tem-
peptides with biological functions require conditions with a high DH perature and pH (50 °C and pH 7) using different enzyme:substrate
(> 10%) (Agyei et al., 2016). In addition to DH, critical process para- ratios (0.5, 1 and 1.5%). Despite minimal differences in the molecular
meters (CPP) that influence product quality include the starting mate- mass distributions of the peptides in all DH 9% samples (62% common
rial quality (such as protein content, traces of other constituents, and peptides) and, accordingly, in all DH 13% samples (84% common
pH), the properties of the enzyme reaction mix (substrate specificity, peptides), there was no significant difference in the inhibitory activity
purity, number of reacting enzymes and their stability and optimal of DPP-IV (dipeptidylpeptidase IV, an enzyme involved in type 2 dia-
reaction conditions) and the process conditions (concentrations of re- betes), indicating that reaction kinetics can be the key advantage to
action constituents, enzyme/substrate ratio, reaction conditions and reduce production costs in this special case.
duration) (Li-Chan, 2015). Cheison and Kulozik (2017) thoroughly The bioactivity of a protein hydrolysate also depends on the applied
described the enzymatic reactions related to whey protein hydrolysis enzyme, which means that enzymes differ in peptide distributions and
and reviewed the environmental influences on DH, with a special focus the chemical characteristics of the protein hydrolysates that they pro-
on the trypsinisation of α-LA and β-LG. Under defined reaction condi- duce (Hudson, 1995). For instance, when fish protein from Selaroides
tions in WPC hydrolysis, enzymes show a preference for certain whey leptolepis was hydrolysed with alcalase, peptides with antioxidant ac-
proteins, whereas others are attacked only partially or not at all. tivities were produced at a low DH of 5%, whereas Flavourzyme 500L
Cheison, Leeb, Toro-Sierra, and Kulozik (2011) showed that WPI hy- hydrolysates worked better at a DH of 25% (Klompong, Benjakul,
drolysis beyond the optimal conditions for trypsin (37 °C and pH 7.8) Kantachote, & Shahidi, 2007). Additionally, bioactive peptides may
led to the recovery of α-LA (67.87% at 25 °C and pH 8.5), while β-LG behave differently in a peptide mixture than when they are pure due to
was completely hydrolysed after 2 h. In contrast, using acid protease A antagonistic and/or synergistic effects. As an example, tests of the in-
(37.5 °C and pH 3.0 for 120 min) allowed β-LG to be retained and α-LA hibition of individual hydrolysed WPIs showed that the α-LA hydro-
to be depleted (Cheison, Bor, Faraj, & Kulozik, 2012). Hence, tem- lysate had the highest DPP-IV-inhibiting activity, whereas after peptide
perature and pH are powerful tools to control the hydrolysis process purification, the more active peptides originated from β-LG (Lacroix &
and to enable targeted peptide release. For example, the use of pH > Li-Chan, 2013; Lacroix & Li-Chan, 2014). Li-Chan (2015) also men-
7.5 or < 2.0 or temperature above 40 °C caused the β-LG dimer to tioned that the coexistence of different proteins in WPIs and con-
dissociate to the monomeric form and the protein to undergo the formational changes in proteins during the industrial hydrolysis process
Tanford transition, improving its hydrolysis. In contrast, pH and tem- could cause different peptide activity profiles by comparing commer-
perature values out of these ranges led β-LG to self-associate to the cially produced β-LG hydrolysates and research samples. Therefore, it is
dimeric form, and neutral pH led to the octameric form (Cheison, Lai, critical to use industrial production conditions in the discovery of new
Leeb, & Kulozik, 2011). This study showed that different hydrolysis pH bioactive peptides in applied laboratory research.
and temperature values have effects beyond enzyme performance; they
change the enzyme susceptibility of the protein, which leads to an al- 2.2.1. Protein hydrolysis at the industrial scale
teration of the peptide release pattern as a result of cleavage-resistant Several methods make the industrial production of bioactive pep-
peptide bonds (Cheison, Brand, Leeb, & Kulozik, 2011; Cheison, Leeb, tides technically feasible. The pharmaceutical production of bioactive
Letzel, & Kulozik, 2011). In this regard, selectivity as the rate at which a peptides can occur via the use of aggregation-promoting carriers, acidic
protease cleaves each individual cleavage site compared to the total peptide-mediated production, de novo chemical synthesis, or the het-
rate of all cleavage sites of a protein was introduced (Butré, Sforza, erologous expression of the peptides fused to thioredoxin or a small
Wierenga, & Gruppen, 2015), and its dependence on pH was evidenced ubiquitin-related modifier (SUMO) in microbial strains (Aoki & Ueda,
using Bacillus licheniformis protease for WPI hydrolysis at 40 °C. Similar 2013). Application of the above-mentioned large-scale production
to the findings of Cheison, Leeb, Toro-Sierra et al. (2011), the results methods to the food industry is too cost-intensive and therefore not
showed that the susceptibility of β-LG within the WPI increased with economically viable (Aoki & Ueda, 2013; Capriotti et al., 2016; Lafarga
increasing pH and independently of DH, whereas the remaining intact & Hayes, 2017).
α-LA decreased almost linearly with increasing DH. These results As a result, the common method of production is large-scale enzy-
showed that the stability of α-LA was independent of pH. The pre- matic hydrolysis using food-grade enzymes (Lafarga & Hayes, 2017).
ference of β-LG was explained by its higher concentration in the WPI Agyei and Danquah (2011) suggested the use of proteases from selected
and a higher Glu content, belonging to the BLP specificity, compared to LAB strains in protein hydrolysis because of their advantages, which
α-LA. As the selectivity of the enzyme for specific cleavage sites de- include cheap cultivation and less laborious enzyme purification. The
termines the final composition of the hydrolysate, it could be shown proteolytic systems of LAB are composed of a high diversity of

62
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

proteases, thus offering a variety of enzymatic specificities to produce influence on the peptide profile (Carvalho, Pessato, Fernandes, de Lima
specific bioactive peptides (Hati, Patel, Sakure, & Mandal, 2017; Liu, Zollner, & Netto, 2017; Cheison, Brand et al., 2011; Cheison, Leeb,
Bayjanov, Renckens, Nauta, & Siezen, 2010). Toro-Sierra et al., 2011).
The simplest method to perform large-scale hydrolysis is to use free To achieve a high DH, which is favoured for the production of
enzymes in a continuous process that is coupled to an ultrafiltration physiologically active peptides, hydrolysis times of 4–100 h are re-
membrane separation module (Korhonen & Pihlanto, 2006). The filter quired, raising the issue of microbial contamination. While bacterio-
system in the membrane reactor hinders the enzyme’s removal and si- static or bactericidal preservatives were previously a common tech-
multaneously functions as the first purification step in the downstream nique to prevent microbial contamination in industrial production
process. The use of immobilized enzymes offers various advantages processes, newer technologies such as UV treatment, pulsed electric
over conventional soluble hydrolysis, e.g., better hydrolysis control, field applications, or ozone are currently used. New enzyme technolo-
avoidance of enzyme autolysis activity and recycling of the im- gies that withstand high temperatures or acidic conditions can help
mobilized enzyme after the hydrolysis step to make the process eco- eliminate optimal growth conditions for microbes, preventing process
nomically viable (Chmiel, 2011). A recent innovation in bioactive contamination (Pasupuleti & Braun, 2008). Microfiltration steps using a
peptide production is on-column-digestion of proteins on continuous- MWCO of 20 kDa reduced microbial load in WPC before hydrolysis,
flow immobilized enzyme reactors (IMER) as they enable an effective, whereas the peptide fractions after nanofiltration (3 kDa) were free
low cost and fast cleavage process, compared to conventional digestion from bacteria, yeasts and molds (Tavares et al., 2012). Bamdad et al.
processes. An efficient and rapid protein digestion, together with a low (2017) noted enzyme hydrolysis under high pressurization as a method
backpressure of this technique, while operating with a high flow rate, that has a reduced risk of microbial contamination while preserving the
subsequently enhance the productivity and yield of the bioactive pep- shelf-life stability of the product.
tide production system, rendering it suitable for large-scale operations. The DH during the industrial production of peptides is monitored by
Siliceous monoliths possessing a tortuous tri-modal flow-through in-process sampling and determination of the ratio of amine nitrogen to
channel structure with macropores of 20–30 µm pore size and meso- total nitrogen. Once the desired DH is reached, the process is stopped by
pores of ∼20 nm and ∼3 nm are ideal to function as an enzyme support deactivating the enzymes with heat (Pasteurization). The free and total
and to guarantee the access of bulky proteinaceous substrates to the AA contents and the molecular weight distribution of peptides in the
immobilized enzymes. Thus a back-pressure of 64 kPa/cm2 could be hydrolysate are also determined (Pasupuleti & Braun, 2008). Conesa
maintained with a flow rate of 750 mL/cm2 ∗ min, which eliminated the and FitzGerald (2013) investigated the impacts of DH and total solid
necessity of a HPLC pump, making the process safer and cost-efficiently content (5–30% (w/v)) of hydrolysed WPC on the thermal inactivation
(Szymańska et al., 2016). Mao, Černigoj, Zalokar, Štrancar, and Kulozik process of Corolase® PP at 60–75 °C. They found that the enzyme was
(2017) compared two flow-through systems with immobilized trypsin more heat sensitive at a higher DH (∼21%) than at a lower DH
on a polymethacrylate monolith with a pore size of 2.1 µm and 6 µm to (∼10%), indicating that the type of protein or peptide species present
free trypsin hydrolysis. The reactor system could be reused 18 times at a in the solution is important. Additionally, Corolase PP® was more stable
flow rate of 15 mL/min and reached a higher DH (∼10%) in 4 h than at a higher total solid content, which was probably related to the
the stirred solution (∼6%), although the contact time of the im- viscosity of the solution and the decreasing heat transfer coefficient in
mobilized enzyme to its substrate was much shorter due to the autolysis solutions with increasing viscosity. It was also conjectured that the
of trypsin in the free enzyme system. The peptide release pattern was bond of an enzyme to a cleavage site of intact protein has a thermo-
very similar, hence only the two peptides f(125–135) and f(61–69) were protective effect. This theory implies that a low total solid content
missed, whereas the peptides f(41–70), f(41–69), f(101–124) and f combined with a high DH results in more effective enzyme inactivation
(92–101) were only present in the hydrolysates from IMER. However, by the application of heat, which is an interesting factor for the food
the high costs of common immobilization supports often decrease the industry regarding energy costs and product stability. Mohan,
economic advantage of this method, whereas the use of agricultural Udechukwu, Rajendran, and Udenigwe (2015) mentioned that in-
wastes can be an attractive alternative (Ittrat, Chacho, Pholprayoon, activation of the protease is likely more efficient by application of heat
Suttiwarayanon, & Charoenpanich, 2014). Bassan et al. (2016) im- than cold, whereas heat inactivation can result in N-terminal cyclisation
mobilized the enzyme trypsin (E.C. 3.4.21.4) on the agro-industrial of the peptide. Nonetheless, cold-inactivated papain hydrolysate
lignocellulose-rich waste corn cob powder (CCP) to generate bioactive showed higher ferric-reducing capacity than heat-inactivated hydro-
peptides from whey proteins. The use of trypsin-glyoxyl-CCP resulted in lysates after 5 h.
an average DH of 23%, and more than 80% of the enzyme activity
remained after five reuse cycles. 2.3. Further demands on product quality
A more controlled enzymatic hydrolysis reaction requires a constant
pH value, which is achieved by the addition of an alkali or acidic so- The promised physiological effects of valuable peptide-based pro-
lution (usually sodium hydroxide or hydrochloric acid). This so-called ducts must be guaranteed; however, more challenges regarding peptide
pH-stat technique also controls the DH (Adler-Nissen, 1986). Non-pH- characteristics must still be overcome to deliver these effects. Thus,
controlled methods are interesting in industrial applications because another challenge in the large-scale production of bioactive peptides is
they avoid the addition of external additives to the reaction mixture. the stability of the peptides during downstream processing because
Fernández and Kelly (2016) compared the pH-stat technique with free- peptides must survive high pressures and temperatures as well as pH
fall pH hydrolysis and determined that the hydrolysis of WPI with the variations (Lafarga & Hayes, 2017). A change in processing conditions
Bacillus protease preparation Protamex® (containing subtilisin, E.C. can lead to structural modifications and alter or abrogate the physio-
3.4.21.62) via the free-fall pH technique is faster, whereas the pH-stat logical function of the peptide. Acidic treatments target Gln and Asn,
technique is more suitable for WPI hydrolysis with trypsin (E.C. whereas basic conditions affect Cys, Ser and Thr, leading to the pro-
3.4.21.4). The hydrolysis of WPI with papain comparing the pH-stat duction of lysinoalanine and D-AAs and decreasing the bioavailability
and pH free-fall methods resulted in a similar DH, but the peptide of Lys, which undergoes a Maillard reaction (Korhonen, Pihlanto-
profile and physicochemical properties differed. At a constant pH of Leppäla, Rantamäki, & Tupasela, 1998). Mohan, Udechukwu et al.
7.0, the antioxidant activity of the whey protein hydrolysate was (2015) observed that 1% remaining lactose and other sugars under
higher, whereas the DPP-IV activity was not influenced by pH regula- hydrolysis conditions led to a Maillard reaction in papain and alcalase
tion. This finding indicates that bioactive peptide production should be reactions, which influenced the ferric-reducing capacity and SH content
controlled not only on a temporal basis but also by bioactivity tests (Le of peptides, whereas the latter had no influence on the antioxidant
Maux, Nongonierma, Barre, & FitzGerald, 2016), as pH has a decisive bioactivity. The increase in ferric-reducing capacity was related not

63
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

only to whey protein hydrolysis but also to Maillard reaction products, To address the challenge of bitterness, the encapsulation of whey
the occurrence of which depends on protease specificity and tempera- protein hydrolysates comprising bioactive peptides by spray-drying
ture/pH conditions during the reaction process. Deng, Wierenga, utilizing a carrier system of proteins, polysaccharides or lipids is a
Schols, Sforza, and Gruppen (2017) found that lysine/arginine-specific widely applied method. Through this method, other product quality
proteases such as trypsin or α-chymotrypsin were unable to hydrolyse challenges such as hygroscopicity, stability within the food matrix or
glycated cleavage sites or cleavage sites near glycation sites in α-LA, during GIT passage, or bioavailability can be resolved at the same time
whereas hydrolysis with BLP and subtilisin A was independent of gly- (Mohan, Rajendran, He, Bazinet, & Udenigwe, 2015). In bioactive
cation. This means that for Lys/Arg- dependent proteases, a modifica- peptide research or product development, it is crucial to demonstrate
tion of the protein primary structure has a direct influence on selectivity biostability, which can be affected by the spray-drying process. How-
and consequently on the extent of hydrolysis, but not on the rate of ever, studies regarding the bioactivity of spray-dried bioactive whey
hydrolysis. peptides are still scarce. Microencapsulation of an immunomodulatory
A further problem for peptide stability occurs when bioactive pep- whey protein hydrolysate with WPC or a WPC/sodium alginate mixture
tides have destinations inside the human body. The passage through the by spray-drying produced a significantly lower bitter taste. The results
gastrointestinal tract (GIT) can strongly hinder their effectiveness, of this study evidenced that spray-drying preserved the im-
particularly long-chain peptides with a certain conformation, because munomodulatory activity of the whey protein hydrolysate and there-
of the harsh pH conditions and presence of the GIT enzymes (Bouglé & fore is a good strategy for the industrial production of encapsulated
Bouhallab, 2017). Regarding short peptides, García-Mora et al. (2017) non-bitter peptides due to its low processing costs (Ma et al., 2014).
reported that the simulated gastrointestinal digestion of lentil-derived
peptides generally enhanced antioxidant and ACE-inhibiting activity, 3. Purification of bioactive peptides at the industrial scale and its
resulting in the additional release of smaller peptide fragments with challenges
synergistic biological effects. A study by García-Tejedor et al. (2014)
showed that lactoferrin-derived peptides were unstable under in vitro After the hydrolysis step, peptides with different functionalities
GIT conditions, but ACE-inhibiting activity in vivo remained in spon- must be separated and enriched to achieve high-yield products that may
taneously hypertensive rats. For the long bovine peptide lactoferricin, be saleable as bioactive food ingredients. The main challenges in that
stabilisation of its tertiary structure by a sulfur bridge is crucial for its process step are the elimination of so-called contaminant peptides,
antitumoral activity but not for its antibacterial effect (Eliassen et al., which can have antagonistic effects, and the enrichment of all the de-
2002). Naturally phosphorylated or glycosylated peptides such as sired peptides with special physiological features and synergistic ef-
casein-derived peptides or GMP are protected from hydrolysis by di- fects. Peptides can be distinguished by selective criteria, such as charge,
gestive proteases and have an improved binding capacity, thus pre- size, solubility, and hydropathicity (Fernández, Zhu, FitzGerald, &
serving their in vivo activity (Bouglé & Bouhallab, 2017). Recently, Riera, 2014). Whey proteins contain higher proportions of essential and
Sanchón et al. (2018) found that in vitro GIT-simulated digestion of sulfur-containing AAs than caseins (whey proteins: 609 mg/g essential
casein and whey proteins could be considered a good approximation of AAs, 52 mg/g sulfur-containing AAs; caseins: 511 mg/g essential AAs,
in vivo digestion, although the peptide release in vitro was much lower 32 mg/g sulfur-containing AAs) (Anand et al., 2013). Generally, caseins
than in vivo, which was probably related to suboptimal enzyme con- have a higher content of Gln, Pro and Arg, whereas whey proteins are
centration or a lack of enzyme adjuvants in the in vitro digestion. It was rich in Leu, Ile and Val (BCAAs) (Da Silva, Bigo, Barbier, & Rudkowska,
also highlighted that whey peptides containing proline or negative 2017). In recent decades, it has become clear how primary and sec-
charges were resistant to gastrointestinal enzyme digestion. Further- ondary structural characteristics are related to peptide bioactivity
more, the hypocholesterolemic peptides I71IAEK75 and G9LDIQK15 and (Eliassen et al., 2002; Udechukwu, Downey, & Udenigwe, 2018), al-
the DPP-IV-inhibiting peptide I78PAVF82 derived from β-LG were lowing bioactive peptides to be classified into species based on their
identified. These findings are an interesting basis for peptide ex- physiological activity and specific physicochemical characteristics.
amination using simulated GIT conditions. Peptides with Pro or hy-
droxyproline at their C-terminus are generally resistant to digestive 3.1. Bioactive whey peptide primary and secondary structural
degradation (Vermeirssen, Van Camp, & Verstraete, 2004). PEGylation characteristics
of the peptide sequence is an inexpensive and FDA (Food and Drug
Administration)-approved modification that does not influence peptide Many studies have shown that peptides from enzymatic hydrolysed
function and may be used to overcome these stability challenges in whey proteins possess a variety of physiological functions, such as an-
bioactive peptide production. Additionally, PEGylation reduces the tioxidant, antimicrobial (Sah, Vasiljevic, McKechnie, & Donkor, 2017),
antigenicity of α-LA and β-LG sequences (Roberts, Bentley, & Harris, antihypertensive (Bhat, Kumar, & Bhat, 2017), anticancer (Hernández-
2012; Zhong et al., 2016). Ledesma & Hsieh, 2017), opioid (Garg, Nurgali, & Kumar Mishra,
Another challenge is bitterness, which relates to a high DH as a 2016), and immunomodulatory functions (Santiago-López, Hernández-
desired tool for the production of bioactive peptides that provide health Mendoza, Vallejo-Cordoba, Mata-Haro, & González-Córdova, 2016),
benefits. Ney (1971) related peptide bitterness to peptide hydro- which makes them valuable for use in functional food production
phobicity, indicating that peptides with a Q value (a measure of the free (Table 2). The bioactivity of a given peptide sequence can also have
energy (ΔG) needed to transfer peptides from ethanol to water) greater dual or more functions as it is known for the well-researched multi-
than 1400 cal/mole were bitter, whereas peptides with a Q value lower functional peptide lactoferricin (Aoki & Ueda, 2013), and other food
than 1300 cal/mole were predicted to not be bitter. However, hydro- derived peptides (García-Mora et al., 2017; Lacroix, Meng, Cheung, &
phobicity alone cannot be used to predict peptide bitterness, as there Li-Chan, 2016; Pandey, Kapila, & Kapila, 2018).
are peptides with a Q value above 1400 cal/mole that not taste bitter, Bioactive peptides are typically 2–20 AAs in length and are classi-
thus Lafarga, Rai, O'Connor, and Hayes (2016) suggest the use of sen- fied as (a) short peptides, which have fewer than 7 AAs; (b) moderate-
sory evaluation panels. Maehashi and Huang (2009) reported that bitter length peptides (7–25 AAs) and (c) large peptides, which have more
peptides have up to eight AA residues that form a spherical shape and than 25 AAs (Panchaud, Affolter, & Kussmann, 2012). Pihlanto (2006)
contain a BU (binding unit for bitter taste receptor binding) site and an reported that dairy-derived peptides with antioxidant activity usually
SU (stimulating unit) site. AAs that can occupy the BU site are Arg, Phe, contain 5–11 AAs, including hydrophobic residues such as Pro, His, Tyr
His and Pro. The SU site needs a bulky, basic group or a hydrophobic or Trp. Whey protein fractions containing aromatic AAs have stronger
group at the N-terminus. Aluko (2017) reviewed in detail the structural antioxidant potential than simple AAs (Peña-Ramos, Xiong, & Arteaga,
characteristics of bitter peptides derived from food. 2004). Proper AA content and correct positioning within the sequence

64
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

Table 2
Bioactive peptide sequences derived from bovine whey proteins with antioxidant, antimicrobial, antihypertensive, anticancer, opioid or immunomodulatory activity.

Substrate Enzyme Amino acid sequence Activity Reference

β-LG Corolase PP W19YSLAMAASDI29 Antioxidant Hernández-Ledesma, Dávalos,


Bartolomé, and Amigo (2005)
β-LG Thermo-lysin L58QKW61 Antioxidant ACE inhibition del Mar Contreras, Hernández-
L59DTDYKK101 Ledesma, Amigo, Martín-Álvarez, and
Recio (2011)
β-LG Trypsin V15AGTWY20 Antibacterial against Gram- Pellegrini, Dettling, Thomas, and
A25ASDISLLDAQSAPLR40 positive bacteria Hunziker (2001)
I78PAVFK83
V92LVLDTDYK100
β-LG Chymo-trypsin H146IRL149 Opioid activity Pihlanto-Leppälä, Paakkari, Rinta-
Koski, and Antila (1997)
β-LG Trypsin L22AMA25 ACE inhibition Pihlanto-Leppälä et al. (2000) and
L32DAQSAPLR40 Mullally, Meisel, and FitzGerald
V81FK83 (1997)
A142LPMIHR148
β-LG Corolase® PP MAA ACE inhibition O'Keeffe et al. (2017)
VAGT
Lacto-ferrin Pepsin Lactoferricin: F17KCRRWQWRMKKLGAPSITCVRRAF41 Antimicrobial, Antitumoral Eliassen et al. (2002)
α-LA Protease from flower of K16GYGGVSLPEW26 ACE inhibition Tavares et al. (2011)
β-LG Cynara cardunculus D97KVGINYW104
D33AQSAPLRVY42
α-LA Synthetic f(18–20) Tyr-Gly-Tyr Lymphozyte proliferation Gauthier et al. (2006)
f(50–51) Tyr-Gly
BSA Proteinase K F221P222 ACE inhibition Hernández-Ledesma et al. (2011)
Whey protein Alcalase® 2.4L Trp-Tyr-Ser-Leu Antioxidant Zhang, Wu, Ling, and Lu (2013)
concentrate

Abbreviations: β-LG: β-lactoglobulin, ACE: angiotensin-converting enzyme, α-LA: α-lactalbumin, BSA: bovine serum albumin.

can be crucial to the antioxidant activity of peptides. Thus, Chen, prevalent membranolytic ACPs (∼20%) are α-helical with a kink in
Muramoto, and Yamauchi (1995) found a general sequence that pos- their helical axis. They generally have a length of 10–30 AA residues,
sesses high antioxidant characteristics; an N-terminal Val or Leu and are amphipathic, have a cationic net charge between +2 and +9 and
Pro, His and Tyr are included within the sequence. Similarly, Elias, have 40–60% hydrophobic AA content. ACPs are abundant in Arg and
Kellerby, and Decker (2008) concluded that antioxidant peptides often Lys, whereas Lys is overrepresented in toxic peptides. The high hy-
contain hydrophobic AAs, including Val or Leu, at the N-terminus. drophobicity on the apolar face of ACPs leads to greater peptide effi-
Sarmadi and Ismail (2010) also reported that antioxidant properties are cacy and haemolytic activity. Active α-ACPs are rich in Ala, Gly, Ile, Lys
related to composition, hydrophobicity and whether AA sequences and Leu residues, whereas Lys is the preferred positively charged AA
contain Tyr, Trp, Met, Lys, Cys, and His. In whey protein derived (Gabernet, Müller, Hiss, & Schneider, 2016). A well-described anti-
peptides, the Cys, His, Tyr, Lys, Val and Phe content was significantly tumoural whey-derived peptide is lactoferricin B (LfcinB, 25 AAs in
higher as in the WPI hydrolysate, indicating that these AA are crucial length, residues 17–41), which is obtained from bovine lactoferrin. The
for antioxidant activity (Jiang, Zhang, Yuan, Qu, & Feng, 2017). secondary structure consists of an amphipathic α-helix (residues
Antibacterial peptide sequences are very diverse, although most 17–28), a turn (residues 29 and 30) and a β-strand (residues 31–41).
sequences have similar AA composition, net charge and hydrophobicity Together, these form a cyclic structure that is fixed by a disulfide bridge
(Akalin, 2014). Most antimicrobial peptides display physicochemical between Cys19 and Cys36 (Eliassen et al., 2002).
similarities, including small molecular size (10–50 AAs) and cationic ACE (peptidyl dipeptide hydrolase, E.C. 3.4.15.1)-inhibiting pep-
and amphipathic characteristics, that are essential for activity tides are antihypertensive peptides obtained from whey proteins α-LA
(Madureira et al., 2010; Théolier, Fliss, Jean, & Hammami, 2014). In and β-LG and are known as lactokinins (Madureira et al., 2010). Gen-
addition, anionic peptides can contribute to the amphiphilic nature of a erally, ACE-inhibiting peptides have sequences between 2 and 12 AAs
molecule (Li et al., 2016). Mohanty et al. (2016) reported that anti- in length (Hernández-Ledesma, del Mar Contreras, & Recio, 2011).
bacterial peptides can have a length up to 100 AA with amphipathic Pihlanto-Leppälä, Koskinen, Piilola, Tupasela, and Korhonen (2000)
character, possessing a positive net charge of +2 to +9 and exist in reported that the low-molecular-weight (< 1 kDa) fraction from pepsin-
cationic or anionic forms. Dziuba and Dziuba (2014) and Sah et al. hydrolysed α-LA and β-LG inhibited ACE more strongly than fractions
(2017) described three subclasses of cationic antibacterial peptides: (1) containing peptides with higher molecular weights. When aromatic
linear cationic α-helical peptides, (2) cationic peptides that contain Pro AAs, such as Trp, Tyr, Phe and Pro, are located in the C-terminus, they
and Arg and are mainly linear, although some may show extended coils can interact with many subsites of ACE and thus inhibit the enzyme
and (3) peptides containing Cys residues that form disulfide linkages activity (García-Mora et al., 2017). AA residues, such as Leu, Ile, Val,
and stable β-sheets. Milk protein-derived antimicrobial peptides con- Lys and Arg, have also been evidenced to play an important role in
tain higher Lys, Leu, Val, and Pro contents (Dziuba & Dziuba, 2014), significantly increasing ACE inhibition (Pan, Cao, Guo, & Zhao, 2012;
whereas Arg, Lys, Ala, Cys, and Leu dominate various positions in an- Pina & Roque, 2009). Peptide conformation is expected to play a sig-
tibacterial peptides obtained from whey proteins (Sah et al., 2017). nificant role in ACE inhibition by long-chain peptides (Hernández-
Anticancer peptides (ACPs) have a length of 5–30 AAs and a cationic Ledesma et al., 2011).
charge (Chen, Ding, Feng, Lin, & Chou, 2016). Specifically, membra- Peptides with opioid activities, have an affinity for opiate receptors
nolytic ACPs are very similar to antimicrobial peptides and thus share a because the presence of the aromatic AA Tyr at the N-terminus and a
related mode of action. They may act using a variety of models: the Phe at the third or fourth position leads to a structural motif that fits
barrel-stave pore, the toroidal pore, the carpet, the detergent-like and into the binding site of opioid receptors (Meisel, 1997). Well-known
the Shai-Huang-Matzusaki models have been discussed. The most opioid-like sequences generated from bovine whey proteins include the

65
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

α- and β-lactorphins from α-LA f(50–53) and β-LG f(102–105), stimu- ultrafiltration with cut-off values of 0.14 µm and a 5 kDa. Further la-
lating gene expression and mucin secretion in intestinal cells (Martínez- boratory-scale membrane processing using a 1-kDa and, subsequently, a
Maqueda, Miralles, Ramos, & Recio, 2013; Martínez-Maqueda et al., 0.65-kDa membrane resulted in a 4.4-fold increase in antioxidant ac-
2012). Typical opioid peptides possess the N-terminal sequence Tyr- tivity and a 1.3-fold increase in ACE-inhibitory activity. As the most
Gly-Gly-Phe, whereas atypical opioid peptides have only a conserved N- potent novel peptides could be identified the sequences MAA and VAGY
terminal Tyr (sequences: Tyr-X-Phe; Tyr-X1-X2-Phe) (Pihlanto-Leppälä, showing the highest ACE inhibition and MPI with the highest anti-
2000). oxidant activity (O'Keeffe et al., 2017).
Casein and whey proteins are the most important protein sources of High levels of peptide transmission can be achieved by operating at
immunomodulatory proteins (Santiago-López et al., 2016). The im- a pH near the peptide’s isoelectric point, highlighting the importance of
munomodulatory effects whey derived peptides are related to the AA the net charge and charge distribution of the peptide during filtration
content, sequence and length of the peptides. Thus, the im- process (Arrutia, Rubio, & Riera, 2016). However, the use of ultra-
munomodulatory effect of the peptide lactoferricin is related to its filtration is not strongly recommended for the purification of small
positively charged region (Vogel et al., 2002). Acidic and large peptides peptides because some apolar peptides, even though their passage is
derived from a tryptic β-LG hydrolysate increased the secretion of IFN-γ theoretically possible, are unable to cross the membrane and thus are
(a TH1 cytonkine), whereas short peptides of the hydrolysate enhanced lost (Capriotti et al., 2015). Arrutia, Rubio et al. (2016) reported that
TNF-α production of monocytes (Rodríguez-Carrio, Fernández, Riera, & polyethersulfone (PES) membranes with the smallest exclusion pores
Suárez, 2014). It was reported that GMP also has immunomodulatory (1–5 kDa) can have very similar sieve effects by forming a polarized
effects that can be enhanced by digestion with pepsin (E.C. 3.4.23.1) (Li layer over the membrane surface. Thus a flux reduction could be ob-
& Mine, 2004). Short, hydrophobic peptide sequences of 2–12 AA served over time and almost the same peptides crossed both mem-
containing aromatic and charged AA are able to interact with the im- branes, whereas the whey peptide sequences ALPMHIR, TKIPAVFK and
mune system; however the exact mechanism of interaction is still un- SLAMAASDISLLDAQSAPLR were absent in both permeates.
clear, but may be by binding to receptors on the surface of immune To prevent a deposite layer over the membrane, and to overcome the
cells, thus causing immune responses (Wu et al., 2016). Bamdad et al. low selectivity of pressure-driven membrane filtration, charged and low-
(2017) found that 38% of anti-inflammatory peptide sequences derived molecular weight peptides can be separated using electro membrane
from β-LG hydrolysis with alcalase consisted of aromatic or hydro- technologies, as for instance electrodialysis-ultrafiltration (EDUF), which
phobic residues. All of these sequences had a Leu, Phe, or both at the C- uses electric charge and molecular mass to effect peptide separation in
terminus. one step. Compared with conventional ultrafiltration, EDUF uses an
electric field rather than pressure to drive peptides through the mem-
3.2. Process development challenges: up-scaling downstream processes of brane, thus the peptide separation occurs by different electrophoretic
whey peptides mobility or different molecular size of the peptides. Stacking ultrafiltra-
tion membranes of different sizes (20 and 50 kDa) enabled peptide
Current industrial peptide separation techniques are membrane se- fractionation with a higher resolution and grouping into the specific
paration methods, such as ultrafiltration and nanofiltration. bioactivities ACE inhibition and glucose uptake stimulation (Doyen et al.,
Chromatography methods are mainly applied in the pharmaceutical 2014). In a recent study, the tryptic hydrolysation process of WPI rea-
industry (Agyei & Danquah, 2011; Capriotti et al., 2016; Korhonen, lized in separately to the EDUF step (ex situ) was compared to a hydro-
2012; Lafarga & Hayes, 2017; Li-Chan, 2015). Peptide separation via lysis process with simultaneous EDUF processing (in situ) (Suwal, Rozoy,
membrane technologies is a low-cost method that preserves the func- Manenda, Doyen, & Bazinet, 2017). It could be observed, that in situ
tional properties of the peptide and is easy to scale-up. Thus it could be digestions resulted in a higher migration of anionic peptides to their
shown at pilot-scale, that hydrolysis of a WPC with less conventional recovery compartment, while ex situ hydrolysis led to a higher cationic
Cynara cardunculus enzymes and subsequent fractionation using ul- peptide migration. Nonetheless, independent of the digestion process,
trafiltration (20 kDa) and further fractionation of the permeate with bioactive sequences such as IDALNENK (antimicrobial) and VY-
nanofiltration (3 kDa) led to a peptide concentrate below 3 kDa (43% VEELKPTPEGDLEILLQK (hypocholesterolemic) could be recovered at the
w/w protein), which possessed ACE inhibition activity (Tavares et al., anionic recovery compartment and ALPMHIR (antihypertensive), TKI-
2012). Mundi and Aluko (2014) recommended the use of a series- PAVFK (hypocholesterolemic), VLVLDTDYKK (antimicrobial), and
connected membrane ultrafiltration and nanofiltration system, e.g., VAGTWY (antimicrobial) at the cationic recovery compartment. With a
fractionating the hydrolysate with a series of 1–10 kDa cut-off mem- regard to membrane fouling in electrodialysis methods, Persico et al.
branes. O'Loughlin et al. (2014) tested fractionation of heat-pre-treated (2016) showed, that this process is driven by charge distribution of the
WPI (80 °C for 10 min) to separate ACE inhibition and ferrous chelating peptides, which means that only specific sequences are responsible for
activity peptides after enzymatic hydrolysis with Corolase® PP on a this undesirable effects.
pilot scale and compared the process to fractionation without pre- A further method that applies an electric field to membrane filtra-
treated, hydrolysed WPI. Therefore, a series of membrane filtration tion to minimize particle deposition on the membrane is cross-flow
with a molecular weight cut-off (MWCO) range of 1 kDa, 5 kDa, 10 kDa electro-membrane filtration (CFEMF). For CFEMF, an electrical field is
and 30 kDa was used, showing the highest bioactivity in the 1-kDa superimposed upon pressure driven membrane filtration causing se-
permeate fraction of pre-heated WPI. However, to fractionate the heat paration effects as electroosmosis, electrophoresis, and electrolysis.
pre-treated hydrolysate, the 30-kDa membrane had to be changed to a Holder, Scholz, Kulozik, and Hinrichs (2013) verified, that cathode or
0.14-µm membrane due to the high level of insoluble aggregated ma- anode located on the permeate side has an influence on fractionation
terial, but in spite of everything, the work findings of laboratory ex- efficiency, which was tested using six casein derived peptides, which
periments could be validated at pilot scale. Further investigations of the are known to possess antihypertensive activity, but differ in charge,
1-kDa permeate of heat pre-treated WPI hydrolysate and the 30-kDa length and hydrophobicity. With the anode on the permeate side, a
retentate of unheated WPI hydrolysate resulted in an increased iron higher permeation rate for three of these six peptides was observed,
solubility of 72% in the presence of the 1-kDa permeate (high content of whereas those, for which a retention was expected, also permeated.
the hydrophilic amino acids Arg, Lys, and His), which was maintained Further studies showed that in this case negatively charged peptides
to about 98% during simulated GIT digestion (O’Loughlin, Kelly, were mainly affected by the electrical field, while higher transmem-
Murray, FitzGerald, & Brodkorb, 2015). brane pressures increase the convective flow and therefore permeation
At a semi-pilot scale ACE inhibiting and antioxidant peptides from is increased. Positively charged peptides were retained only when a
WPC hydrolysed with Corolase® PP could be obtained after lower transmembrane pressure was applied, but utilization of high

66
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

transmembrane pressures led the drag forces to overcome the electrical radial flow MAC is interesting by the fact, that yields and purities are
effects resulting in an increased peptide transmission to the permeate independent of the flow rate, which means that maximal possible flow
(Holder, Merath, Kulozik, & Hinrichs, 2015). rates can be used to achieve a reduction of process time and costs
Compared to filtration methods, chromatographic methods separate (Brand, Dachmann, Pichler, Lotz, & Kulozik, 2016). Brand, Voigt,
peptides at higher resolution, guaranteeing high-quality (very pure) Zochowski, and Kulozik (2016) showed that loading in recirculation
products. On the other hand, the costs of chromatographic technologies mode instead of a single pass mode increased the binding capacity of
at large scales are exorbitant and are therefore a limiting factor in the the sample. Furthermore this fluid contacting mode prevents pore
commercialization of food ingredients based on bioactive peptides. plugging, due to a mix of convective flow close to the flow channel and
Purification processes resulting in high-purity products, as in the diffusive mass transport in the membrane pores compared to convective
pharmaceutical industry, account for 90% of the downstream process flow alone in radial flow MAC. Additionally, mass transfer of tangential
costs, and even biologically produced chemicals can require more than flow MAC was tenfold higher than for batch adsorption, as it was found
50% of the production costs (Agyei & Danquah, 2011). To scale la- by comparing the effect of temperature to the binding capacity of β-LG
boratorial and pilot processes to industrial levels, parameters such as in batch adsorption, radial and tangential flow using anion exchange
resin material, particle size, linear flow rate, bed height and buffer MAC. Increasing the temperature up to 50 °C led to higher mass transfer
composition must remain constant. In the case of gel filtration methods, but lowered the binding capacity. Thus, working at lower temperatures,
zone broadening relative to the column length must not change, for instance, 10 °C was recommended to additionally prevent microbial
whereas adsorptive chromatography techniques require a constant ratio growth during protein processing (Voswinkel, Etzel, & Kulozik, 2017).
of product mass to adsorber mass. Furthermore, step gradients, rather In this way, Leeb et al. (2014) obtained the ACE inhibitor peptide β-LG
than linear gradients, are usually used in the laboratory to facilitate (9–14) with a yield of 52% from a complex tryptic β-LG hydrolysate by
elution. Thus, an upscaling of the developed laboratory process is often ion exchange MAC in a two-step fractionation process using a strong
economically not viable, and subsequent modifications of the operating anion exchange and cation exchange adsorber module in series. While
process generate highly out-of-scale costs (Chmiel, 2011). From an CFEMF enables the selective fractionation of a complex peptides mix-
economic point of view, manufacture of biotechnological products for ture into two fractions, the fractionation capacity of ion exchange MAC
the food industry require different chromatography processes than the is much higher, thus it produces a high amount of fractions, carrying
conventional using bead-based columns due to the fact, that low flow different properties. However, in contrast to CFEMF the ion exchange
rates and a high backpressure lead to suboptimal enrichment processes MAC requires further treatment after peptide enrichment because of the
(Leeb et al., 2014). Table 3 gives several examples of enrichment high salt content in the elution buffer.
methods (explained in the text below) that are applicable for whey- As a further peptide enrichment process, column adsorption chro-
derived peptides and may also operate at the industrial scale in the matography using activated carbon (AC) as an adsorbent enables an
future. enrichment of Trp-containing peptides such as Ile-Trp or Ala-Trp, which
Thus, membrane chromatography (MC) seems to be an interesting show ACE-inhibiting activity (Hippauf, Lunow, Borchardt, Henle, &
purification process, as it is already used in industry for protein se- Kaskel, 2014). For an economic up-scale and to allow higher through-
paration. This technique combines ultrafiltration with chromatography puts in a continuous system Hippauf et al. (2016) used adsorbent col-
by functionalizing the filtration membrane with grafted ligands that umns loaded with granulated AC instead of powdered AC. This column
have special chemical features (e.g. ion exchangers) that select for the arrangement can replace expensive preparative HPLC equipment and
desired proteins (Bhut, Weaver, Carter, Wickramasinghe, & Husson, therefore enables large-scale production. AC is a non-toxic material that
2011). Advantages of MC over conventional chromatographic methods can be produced from natural bioresources, including algae, wood
are its low backpressure, high flow rates and ease of scale-up (Chik & sawdust or agricultural waste such as cow manure or pulp-mill sludge.
Saufi, 2010). To this effect the use of adsorbent materials as alternatives Hydrophobic, π-π and other intermolecular interactions with the AC are
in peptide purification is mentioned in the literature. Membrane ad- responsible for peptide adsorption.
sorption chromatography (MAC) can be applied as a high throughput As a chromatography free, recently developed separation method,
ion exchange chromatography method, which is suited for selection an aqueous two-phase system which consist of poly(ethylene glycol-
between different net charges of proteins or peptides (Voswinkel & ran-propylene glycol) monobutyl ether/phosphate (EOPO Co-polymer
Kulozik, 2011). The application of a radial flow MAC, which has a spiral (UCON)/Phosphate) was applied for enrichment of antioxidant pep-
wound membrane instead of stacked layers with axial flow, resulted in tides derived from a pepsin hydrolysate of WPI. This technique was
more homogenous flow properties. At 50-fold pilot scale, Voswinkel developed to facilitate and accelerate the enrichment step for bioactive
and Kulozik (2014) showed that a linear scale up by maintaining the peptides, as antioxidant and hydrophobic AA moved to the top phase of
bed height is possible as the dynamic binding capacity was close to that the two-phase system at acidic conditions (pH 4), which consequently
of lab scale and high flow rates of 5 bed volumes min−1 could be showed a higher antioxidant activity as peptides in the bottom phase.
maintained without cleaning at 5 repeated cycles. The scalability of The antioxidant sequences DIQKVAGTWYSL, NENKVLVLDTDYKKY and

Table 3
Methods to enrich bioactive whey peptides that may be amenable to industrial scale-up.

Enrichment step Enriched peptide class Advantages of the method Method source

Molecular weight cut-off membrane Depends on molecular weight of peptides Simple and rapid method with large-scale application Mundi and Aluko
filtration (2014)
Cross-flow electro membrane filtration Electrostatically charged peptides with Reduce membrane fouling High permeation flux can be maintained Holder et al.
different molecular weight Low costs and easy to scale up (2013)
Ion exchange adsorber membrane Discriminates between charge High separation selectivity High flow rates applicable, while Leeb et al. (2014)
chromatography backpressure remains low
Adsorption with activated carbon Short Trp-containing peptides, BCAA-rich Non-toxic material, can be produced from bioresources, enables Hippauf et al.
peptides large-scale production (2016)
Extraction with EOPO Co-polymer Peptides with hydrophobic and Rapid and inexpensive method Can be specifically separate Jiang et al. (2017)
(UCON)/Phosphate antioxidative AA peptides with different properties Polymer can be reused Up-scaling
possible

Abbreviations: SDS: sodium dodecyl sulphate, BCAA: branched-chain amino acid, AAs: amino acids.

67
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

CAQKKIIAEKTKIPAVF could be identified and under optimum condi- methodologies (Tong, Li, Bai, & Li, 2017; Udenigwe, 2014). Fig. 2 lists
tions (pH 4, 4:4 vol ratio of 40% (w/w) UCON solution and 15.5% (w/ the possible applications of bioinformatics to support the classical ap-
w) KH2PO4 solution, 2 mL of 49.35 mg/mL WPI hydrolysate and 0.40 g/ proach and shows research tendencies in the field of bioactive peptides
10 mL NaCl) the extraction efficiency was 38.95%. Because of structural derived from whey protein hydrolysates, which are explained in the
changes of EOPO, due to high sensitivity to external conditions, the text.
polymer can be separated from the peptides and reused, which ad- A food protein source can be selected using random sequence si-
ditionally turns the technique viable for industrial application (Jiang milarity. As sequences up to 20 AA in length are too short to be eval-
et al., 2017). uated without risk of error, the programs must be carefully chosen. As
recommended by Minkiewicz, Dziuba, Iwaniak, Dziuba, and Darewicz
4. In silico strategies to develop economically feasible (2008), MS BLAST can find short peptide precursors. Recently, a novel
downstream processing of bioactive peptides platform for bioactive peptides from milk accessible under the URL
http://mbpdb.nws.oregonstate.edu, was published (Nielsen, Beverly,
Experimental trial and error approaches, which reflect the conven- Qu, & Dallas, 2017). With the help of this database, named Milk
tional process stream at the laboratory scale (described above), are cost- Bioactive Peptide Database (MBPD), it will be possible to screen protein
intensive and time-consuming (Agyei et al., 2016). Application of dif- sources for bioactive peptides, based on peptide structure activity re-
ferent proteases results in hydrolysates with diverse peptide profiles lationship data. Only peptides whose sequence, related to the bioac-
and subsequent bioactivities (Welsh et al., 2017). Processing conditions tivity, is exactly known were included to the database. Additionally, as
in pre-treatment and hydrolysis for instance: time, temperature, pH, a new feature, these bioactive peptides were visually mapped on pro-
pressure, buffer type or enzyme-substrate-ratio also have a crucial in- tein sequences derived from human or cow proteins.
fluence on the enzyme specificity resulting in varying peptide profiles Bioinformatic approaches furthermore enable the prediction of
(Adjonu et al., 2013; Cheison, Brand et al., 2011; Cheison, Lai et al., peptide sequences by simulation of hydrolysis in silico. The bioactive
2011; Cheison, Leeb, Letzel et al., 2011; Cheison, Leeb, Toro-Sierra behaviour of the resulting peptide sequences can be subsequently ex-
et al., 2011; Leeb, Kulozik, & Cheison, 2011). Thus, as mentioned by Li- amined using physicochemical or structural properties, such as electric
Chan (2015) it is important to keep process parameters constant during charge or hydrophobicity, searching for sequence motifs and predicting
up-scaling to guarantee a stable peptide profile and underlines the allergenic, sensory, toxic properties or the ability to permeate the in-
importance to use industrial production conditions in the discovery of testines (Fu & Lin, 2017; Iwaniak, Minkiewicz, Darewicz, Sieniawski, &
new bioactive peptides in applied laboratory research. Starowicz, 2016; Jung et al., 2007; Minkiewicz et al., 2008). Today, it is
Bioinformatic approaches are innovative and helpful tools to over- well known, that peptides possessing certain length, and physico-
come expensive and time-consuming experiment trials, which can lead chemical properties, defined sterical conformation, specific AA pat-
to a more systematic process design and invention of new terns, or given AAs at defined positions within the peptide, can be

Fig. 2. Initial approach for predictive application of bioinformatics (marked with blue stars) to the conventional research downstream process. The design of the downstream process
involves hybrid methods, like algorithmic methods, knowledge-based approaches (heuristics) and mechanistic models. These important tools, commonly used in biopharmaceutical
industry, can be crucial for peptide enrichment process development, with the possibility of industrial upscaling. CQA, critical quality attribute, for example measurable product
properties within product quality requirements. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

68
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

grouped into various bioactivities, as treated above. To explore the complex nature of protein-enzyme interactions was also mentioned by
bioactive profiles and physicochemical properties of peptides, Udenigwe (2014). Gu and Wu (2013) named factors such as pH, tem-
Minkiewicz et al. (2008) reviewed the potential application of internet- perature, the formation of complex protein structures, and the catalytic
available computational peptide science tools and still provides a efficiency of proteases. However, in silico analysis is time-saving and,
comprehensive list of currently 62 bioactive peptide databases, avail- compared to the traditional in vitro experimental approaches, a more
able online with URL http://www.uwm.edu.pl/biochemia/index.php/ economical strategy for bioactive peptide investigations (Fu et al.,
pl/biopep/32-bioactive-peptide-databases. 2016).
A platform that is widely used to implement the above-mentioned During peptide processing it is most important not to affect the
computational approaches is the BIOPEP database (http://www.uw- functional regions of the peptides. Molecular docking studies are useful
m.edu.pl/biochemia), which was developed at the Chair of Food to explain the mode of action of involved AAs in certain peptide
Biochemistry at the University of Warmia and Mazury in Olsztyn, bioactivities (Tu et al., 2017). García-Mora et al. (2017) identified the
Poland (Minkiewicz et al., 2008; Nongonierma, Le Maux, Hamayon, & molecular mechanism of dual antioxidant/ACE inhibiting activity, de-
FitzGerald, 2016; Tulipano, Faggi, Nardone, Cocchi, & Caroli, 2015). rived from lentil proteins, which were hydrolysed with Savinase®. Thus,
The platform can currently simulate an in silico hydrolysis of the desired the C-terminal tripeptide RYL of lentil derived peptides plays a key role
protein with a library of 31 proteolytic enzymes, from which two en- in ACE inhibiting activity, while interacting strongly through a network
zymes can be tested at a different pH. Various strategies for enzyme of hydrogen bonds with the enzyme. Furthermore, other ACE inhibiting
selection were applied until today. Recent studies tried to overcome this peptides listed in the BIOPEP database, also possess this C-terminal
hurdle using a ranking of A-values (defined as a ratio of the number of tripeptide, which underlines its importance. These results imply the
peptides with a given activity to the number of AA residues in a protein necessity to preserve such key regions during downstream processing,
sequence) or an additional appliance of the PeptideRanker (http:// to prevent the loss of the peptides bioactivity.
bioware.ucd.ie/∼compass/biowareweb/Server_pages/peptider- Regarding the enzymatic hydrolysis step, several working groups
anker.php), developed by Mooney, Haslam, Pollastri, and Shields tried to embody the complex reaction kinetics into mathematical
(2012), beside a comparison with literature data (Agirbasli & Cavas, modelling. Valencia, Espinoza, Ceballos, Pinto, and Almonacid (2015)
2017; Lafarga, Aluko, Rai, O'Connor, & Hayes, 2016; Wang et al., proposed a new methodology to predict hydrolysis time under different
2017). operation conditions for protein concentration, enzyme concentration
Evaluation of potential biological and physiological activities, al- and temperature, usable for different protein sources and enzymes.
lergenicity, sensory activity, and peptide release also can be performed Maresca and Ferrari (2017) worked out a model for hydrolysis of BSA
with the BIOPEP platform (Iwaniak et al., 2016; Minkiewicz et al., α-chymotrypsin and trypsin under high pressure conditions
2008). Minkiewicz et al. (2008) mentioned that predicting the sec- (100–600 MPa, 37 °C, hydrolysis time: 0–25 min), which, after inclusion
ondary structures and physicochemical properties of short sequences is in future enzyme-substrate ratios, different processing conditions and
more promising than searching for sequence homology and re- types of enzyme is thought to be usable for hydrolysis process devel-
commended the parallel use of other databases and programs that opment in the food, ingredient or pharmaceutical industry. To predict
process and evaluate peptide sequences, such as CutDB for protein the impact of Maillard reaction in protein structure to enzyme specifi-
hydrolysis, ProtParam for physicochemical properties or REMUS to city, selectivity and binding site sensitivity in vitro reaction kinetics
evaluate differences in peptide fragments (Gasteiger et al., 2005; were modelled (Deng et al., 2017).
Igarashi et al., 2007; Pai et al., 2006). A useful program to predict To use the full capacity of whey derived proteins, in a bioeconomic
secondary structures for linear peptides is PEP-FOLD3 or can be found sense, it is fundamental to achieve an optimal design of the enrichment
at the UniProt server (Lamiable et al., 2016; UniProt Consortium, process in order to separate and enrich peptide groups with different
2008). biofunctions. This can be carried out in series by concatenate unit op-
Dziuba and Dziuba (2014) indicated, as is the case for antimicrobial erations or by simultaneous fractionation, as it was done by Suwal et al.
peptides (AMP), that the simple prediction of physicochemical prop- (2017). Therefore, computational tools, as mentioned above, facilitate
erties and values of biological activity indicators is inadequate to effect the investigation of a peptide’s bioactivity and its physicochemical
a final selection, and must be complemented with multidimensional properties. Combining possible hydrolysis results with in silico bio-
statistical analysis using machine learning algorithms, such as support prospecting can enable organizing the peptides into bioactivity cate-
vector machines (SVM), random forest (RF), artificial neuronal net- gories whose members share special physicochemical properties. For
works (ANN) and discriminant analysis (DA), which are available in the example, hydrophobic sequences with sulfhydryl groups can act as
CAMP (Collection of Anti-Microbial Peptides) database. Quantitative antioxidant peptides, whereas hydrophobic peptides with cationic
structure-activity relationship (QSAR) modelling has been recently charges often have antimicrobial properties (Dziuba & Dziuba, 2014;
applied to investigate DPP-IV activity (Nongonierma & FitzGerald, Tian et al., 2015).
2016). Gabere and Noble (2017) verified on the basis of AMP prediction Additionally, the use of a hybrid methods model that incorporates
tools that the quality of the predictions is difficult to quantify and tested experimental, heuristic and in silico processes can lead to a successful
two sets of AMP (antimicrobial, antibacterial and bacteriocins) and downstream operating process that can be industrially upscaled (Li-
non-AMP on ten publicly available prediction tools. The CAMPR3 Chan, 2015; Nfor, Verhaert, van der Wielen, Hubbuch, & Ottens, 2009).
(random forest) model was evaluated to have the best prediction per- As an example from the biopharmaceutical industry, Nfor et al. (2009)
formance for AMPs. proposed an in silico process design to model a protein purification
Capriotti et al. (2016) advised that the use of computational tools system based on the simulated behaviour of biomolecules. Transferring
strongly depends on their ability to recognize bioactive sequences. this computational modelling approach to the field of peptide enrich-
However, Singh, Sharma, Kumari, and Korpole (2014) reported, that a ment and combining it with knowledge-based procedures could help to
new bacteriocin, recognized by the prediction tools for antibacterial select an economically effective and time-saving operating process,
activity CAMPR3 and BACTIBASE, had no significant similarity with which would enable upscaling to an industrial level. Hanke and Ottens
known sequences available in that databases. Chatterjee, Kanawjia, (2014) listed the commonly required parameters that are needed to
Khetra, and Saini (2015) criticized that many peptides obtained in vitro describe a mechanistic model. Appropriate amino acid descriptors from
were not predicted in silico, and thus proposed prediction tools that are QSAR can be used to rate a peptide’s hydrophilicity/hydrophobicity,
more flexible, including variables such as pH, time, temperature, and molecular size/bulkiness and electronic properties/charge (Kim & Li-
enzyme:substrate ratio. The fact that in vitro production of peptides that Chan, 2006). Notably, VHSE (a set of vectors describing hydrophobic,
were predicted in in silico digestion may not occur as a result of the steric and electronic properties) has better predictive power than other

69
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

named descriptors because it uses a broader range of physicochemical prospection, determination of physicochemical properties, mathema-
properties (Xie et al., 2013). 3D characterization (molecular weight, tical simulation of hydrolysis and “omics”-integrated techniques, such
isoelectric point and hydrophobicity) coupled to a random forest cali- as molecular docking studies can help to overcome the challenges of
bration or quantitative structure-property relations (QSPR) are helpful time-consuming and cost-intensive lab-scale peptide discoveries and
to evaluate peptide separation and column performance to optimize an characterization. The use of new viable peptide separation methods
enrichment step. Further important parameters are fluid flow and mass (e.g. EDUF, CFEMF or simultaneous hydrolysis/enrichment) in labora-
transfer in the column, as well as solute mixture composition, its mo- tory research, facilitated by mechanistic modelling can bridge up-
lecular properties, possible sample interaction, and the resin- in the scaling challenges in an economical way, as they relate to their in-
form of absorption isotherms. These in silico based predictions can also dustrial application in the functional food industry. Early monitoring of
be applied for similar solutes or to determine characteristics for con- the downstream process costs during development at the lab scale fa-
taminants to eliminate them, namely heuristic flowsheeting (Hanke & cilitates deciding which methods apply and contributes to the afford-
Ottens, 2014). ability of the scale-up. Considering these systematic process design
The economic feasibility of an industrial, large-scale process de- factors (such a proposed processing pathway), applicable for whey
pends strongly on early planning-phase decisions that are based on the derived peptides, is likely to produce various targeted whey hydrolysate
laboratorial scale process (Cziner, Virkki-Hatakka, Hurme, & Turunen, products, with a high content of the desired bioactive peptides.
2005). Therefore, the decision to introduce alternative methods to
purification downstream processes always requires a cost evaluation. Acknowledgements
Winkelnkemper and Schembecker (2010) proposed a new performance
indicator for rating purifications and their cost-efficiency, based on We would like to thank Conselho Nacional de Desenvolvimento
purity improvements, yields and specific costs from single steps, which Científico e Tecnológico (CNPq) and Coordenação de Aperfeiçoamento
can be extracted from the beginning of the experimental investigation. de Pessoal de Nível Superior (CAPES) for scholarships. We would also
Purification processes can be divided into (A) an initial purification like to thank Universidade do Vale do Taquari - Univates and Secretaria
step, which results in moderate purity; (B) an intermediate purification do Desenvolvimento Econômico, Ciência e Tecnologia (SDECT) of Rio
step, which generates a moderately pure to high purity sample; and (C) Grande do Sul state for financial support.
a final purification step, which requires a great deal of effort but
reaches very high purity. Conflicts of interest
High to very high purity in food industry peptide-based ingredients
is not an essential target (Lafarga & Hayes, 2017). The purification The authors declare they do not have any conflicts of interest.
processes can be rated by using the classical parameters of purity dif-
ferences, purification factors and clearance factors (a measure of the References
reduction in the level of impurities), as well as by normalizing indices,
such as the purification index, the logarithmic purification index, and Adjonu, R., Doran, G., Torley, P., & Agboola, S. (2013). Screening of whey protein isolate
the purification performance index (PPI). The PPI, defined as the ratio hydrolysates for their dual functionality: Influence of heat pre-treatment and enzyme
specificity. Food Chemistry, 136(3), 1435–1443.
of the logarithmized clearance factors of the step to those of the total Adler-Nissen, J. (1979). Determination of the degree of hydrolysis of food protein hy-
process, is the factor that is most balanced in describing the purification drolysates by trinitrobenzenesulfonic acid. Journal of Agricultural and Food Chemistry,
process and can thus be used to link a purification performance with its 27(6), 1256–1262.
Adler-Nissen, J. (1986). Enzymic hydrolysis of food proteins. Elsevier Applied Science
expected effort. The separation cost indicator (SCI) depends on the Publishers.
normalized purification rating, yield and specific cost of each single Aggeli, A., Bell, M., Carrick, L. M., Fishwick, C. W., Harding, R., Mawer, P. J., ... Boden, N.
step, and allows a quantitative rating of cost-efficiency to be made. (2003). PH as a trigger of peptide β-sheet self-assembly and reversible switching
between nematic and isotropic phases. Journal of the American Chemical Society,
Furthermore, SCI can be used at the beginning of an experiment-based 125(32), 9619–9628.
process to direct purification development towards an economically Agirbasli, Z., & Cavas, L. (2017). In silico evaluation of bioactive peptides from the green
viable procedure (Winkelnkemper & Schembecker, 2010). algae Caulerpa. Journal of Applied Phycology, 29(3), 1635–1646.
Agyei, D., & Danquah, M. K. (2011). Industrial-scale manufacturing of pharmaceutical-
grade bioactive peptides. Biotechnology Advances, 29(3), 272–277.
5. Conclusion Agyei, D., Ongkudon, C. M., Wei, C. Y., Chan, A. S., & Danquah, M. K. (2016). Bioprocess
challenges to the isolation and purification of bioactive peptides. Food and
There is an increasing interest to integrate bioactive whey peptides Bioproducts Processing, 98, 244–256.
Akalın, A. S. (2014). Dairy-derived antimicrobial peptides: Action mechanisms, phar-
as ingredients in functional foods because of their physiological activ- maceutical uses and production proposals. Trends in Food Science & Technology, 36(2),
ities such as providing health-promoting benefits to the corporal im- 79–95.
mune, cardiovascular, nervous and gastrointestinal system. However, Aluko, R. E. (2017). Structural characteristics of food protein-derived bitter peptides.
Bitterness: Perception, Chemistry and Food Processing105–129.
production at the lab scale and a subsequent scale-up to the industrial Anand, S., Som Nath, K., & Chenchaiah, M. (2013). Whey and whey products. Milk and
scale requires surmounting many challenges. Production processes, Dairy Products in Human Nutrition: Production, Composition and Health477–497.
such as pre-treatment and controlled hydrolysis, increase the specific Aoki, W., & Ueda, M. (2013). Characterization of antimicrobial peptides toward the de-
velopment of novel antibiotics. Pharmaceuticals, 6(8), 1055–1081.
release of particular bioactive peptides. Hydrolysis and enrichment Arrutia, F., Puente, Á., Riera, F. A., Menéndez, C., & González, U. A. (2016). Influence of
must be performed under mild conditions to avoid affecting the struc- heat pre-treatment on BSA tryptic hydrolysis and peptide release. Food Chemistry,
ture and function of the peptides. Overcoming the bitterness and 202, 40–48.
Arrutia, F., Rubio, R., & Riera, F. A. (2016). Production and membrane fractionation of
maintaining the stability of these bioactive peptides are other chal- bioactive peptides from a whey protein concentrate. Journal of Food Engineering,
lenges, and they must be resolved while maintaining cost-friendly 184, 1–9.
production conditions. Achieving a more systematic process flow of Ballard, K. D., Bruno, R. S., Seip, R. L., Quann, E. E., Volk, B. M., Freidenreich, D. J., ...
Volek, JS. (2009). Acute ingestion of a novel whey-derived peptide improves vascular
product development that can be cost-efficiently up-scaled to an in-
endothelial responses in healthy individuals: A randomized, placebo controlled trial.
dustrial production are the main challenges in this field, and passing Nutrition Journal, 8(1), 34.
these obstacles has led to a reorganization of conventional downstream Bamdad, F., Bark, S., Kwon, C. H., Suh, J. W., & Sunwoo, H. (2017). Anti-inflammatory
processes that requires the integration of hybrid methods. In contrast to and antioxidant properties of peptides released from β-lactoglobulin by high hy-
drostatic pressure-assisted enzymatic hydrolysis. Molecules, 22(6), 949.
bioactive peptide production in the pharmaceutical industry, purifica- Bassan, J. C., de Souza Bezerra, T. M., Peixoto, G., da Cruz, C. Z. P., Galán, J. P. M., Vaz,
tion of bioactive peptides by chromatographic methods is cost-prohi- A. B. D. S., ... Monti, R. (2016). Immobilization of trypsin in lignocellulosic waste
bitive for peptide enrichment in the food industry, preventing their use material to produce peptides with bioactive potential from whey protein. Materials,
9(5), 357.
in large-scale production. In silico approaches like peptide activity

70
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

Bhat, Z. F., Kumar, S., & Bhat, H. F. (2017). Antihypertensive peptides of animal origin: A Da Silva, M. S., Bigo, C., Barbier, O., & Rudkowska, I. (2017). Whey protein hydrolysate
review. Critical reviews in food science and nutrition, 57(3), 566–578. and branched-chain amino acids downregulate inflammation-related genes in vas-
Bhut, B. V., Weaver, J., Carter, A. R., Wickramasinghe, S. R., & Husson, S. M. (2011). The cular endothelial cells. Nutrition Research, 38, 43–51.
role of polymer nanolayer architecture on the separation performance of anion-ex- De Maria, S., Ferrari, G., & Maresca, P. (2017). Effect of high hydrostatic pressure on the
change membrane adsorbers: I. Protein separations. Biotechnology and Bioengineering, enzymatic hydrolysis of bovine serum albumin. Journal of the Science of Food and
108(11), 2645–2653. Agriculture, 97(10), 3151–3158.
Blayo, C., Vidcoq, O., Lazennec, F., & Dumay, E. (2016). Effects of high pressure pro- del Mar Contreras, M., Hernández-Ledesma, B., Amigo, L., Martín-Álvarez, P. J., & Recio,
cessing (hydrostatic high pressure and ultra-high pressure homogenisation) on whey I. (2011). Production of antioxidant hydrolyzates from a whey protein concentrate
protein native state and susceptibility to tryptic hydrolysis at atmospheric pressure. with thermolysin: Optimization by response surface methodology. LWT-Food Science
Food Research International, 79, 40–53. and Technology, 44(1), 9–15.
Bouglé, D., & Bouhallab, S. (2017). Dietary bioactive peptides: Human studies. Critical Delahaije, R. J., Gruppen, H., van Eijk- van Boxtel, E. L., Cornacchia, L., & Wierenga, P. A.
Reviews in Food Science and Nutrition, 57(2), 335–343. (2016). Controlling the ratio between native-like, non-native-like, and aggregated β-
Brand, J., Dachmann, E., Pichler, M., Lotz, S., & Kulozik, U. (2016). A novel approach for lactoglobulin after heat treatment. Journal of Agricultural and Food Chemistry, 64(21),
lysozyme and ovotransferrin fractionation from egg white by radial flow membrane 4362–4370.
adsorption chromatography: Impact of product and process variables. Separation and Deng, Y., Wierenga, P. A., Schols, H. A., Sforza, S., & Gruppen, H. (2017). Effect of
Purification Technology, 161, 44–52. Maillard induced glycation on protein hydrolysis by lysine/arginine and non-lysine/
Brand, J., Voigt, K., Zochowski, B., & Kulozik, U. (2016). Lysozyme fractionation from egg arginine specific proteases. Food Hydrocolloids, 69, 210–219.
white at pilot scale by means of tangential flow membrane adsorbers: Investigation of Doucet, D., Otter, D. E., Gauthier, S. F., & Foegeding, E. A. (2003). Enzyme-induced ge-
the flow conditions. Journal of Chromatography A, 1438, 143–149. lation of extensively hydrolyzed whey proteins by Alcalase: Peptide identification
Bulone, D., Martorana, V., & San Biagio, P. L. (2001). Effects of intermediates on ag- and determination of enzyme specificity. Journal of Agricultural and Food Chemistry,
gregation of native bovine serum albumin. Biophysical Chemistry, 91(1), 61–69. 51(21), 6300–6308.
Butré, C. I., Sforza, S., Wierenga, P. A., & Gruppen, H. (2015). Determination of the in- Doyen, A., Udenigwe, C. C., Mitchell, P. L., Marette, A., Aluko, R. E., & Bazinet, L. (2014).
fluence of the pH of hydrolysis on enzyme selectivity of Bacillus licheniformis protease Anti-diabetic and antihypertensive activities of two flaxseed protein hydrolysate
towards whey protein isolate. International Dairy Journal, 44, 44–53. fractions revealed following their simultaneous separation by electrodialysis with
Capriotti, A. L., Cavaliere, C., Foglia, P., Piovesana, S., Samperi, R., Chiozzi, R. Z., & ultrafiltration membranes. Food Chemistry, 145, 66–76.
Laganà, A. (2015). Development of an analytical strategy for the identification of Dziuba, B., & Dziuba, M. (2014). New milk protein-derived peptides with potential an-
potential bioactive peptides generated by in vitro tryptic digestion of fish muscle timicrobial activity: An approach based on bioinformatic studies. International Journal
proteins. Analytical and Bioanalytical Chemistry, 407(3), 845–854. of Molecular Sciences, 15(8), 14531–14545.
Capriotti, A. L., Cavaliere, C., Piovesana, S., Samperi, R., & Laganà, A. (2016). Recent Elias, R. J., Kellerby, S. S., & Decker, E. A. (2008). Antioxidant activity of proteins and
trends in the analysis of bioactive peptides in milk and dairy products. Analytical and peptides. Critical Reviews in Food Science and Nutrition, 48(5), 430–441.
Bioanalytical Chemistry, 408(11), 2677–2685. Eliassen, L. T., Berge, G., Sveinbjørnsson, B., Svendsen, J. S., Vorland, L. H., & Rekdal, Ø.
Carvalho, N. C.d., Pessato, TB., Fernandes, LGR., de Lima Zollner, R., & Netto, FM. (2002). Evidence for a direct antitumor mechanism of action of bovine lactoferricin.
(2017). Physicochemical characteristics and antigenicity of whey protein hydro- Anticancer Research, 22, 2703–2710.
lysates obtained with and without pH control. International Dairy Journal, 71, 24–34. Expósito, I. L., & Recio, I. (2006). Antibacterial activity of peptides and folding variants
Chatterjee, A., Kanawjia, S. K., & Khetra, Y. (2016). Properties of sweetened Indian yogurt from milk proteins. International Dairy Journal, 16(11), 1294–1305.
(mishti dohi) as affected by added tryptic whey protein hydrolysate. Journal of Food Fernández, A., & Kelly, P. (2016). PH-stat vs. free-fall pH techniques in the enzymatic
Science and Technology, 53(1), 824–831. hydrolysis of whey proteins. Food Chemistry, 199, 409–415.
Chatterjee, A., Kanawjia, S. K., Khetra, Y., & Saini, P. (2015). Discordance between in Fernández, A., & Riera, F. (2013). Β-Lactoglobulin tryptic digestion: A model approach
silico & in vitro analyses of ACE inhibitory & antioxidative peptides from mixed milk for peptide release. Biochemical Engineering Journal, 70, 88–96.
tryptic whey protein hydrolysate. Journal of Food Science and Technology, 52(9), Fernández, A., Zhu, Y., FitzGerald, R. J., & Riera, F. A. (2014). Membrane fractionation of
5621–5630. a β-lactoglobulin tryptic digest: Effect of the membrane characteristics. Journal of
Cheison, S. C., Bor, E. K., Faraj, A. K., & Kulozik, U. (2012). Selective hydrolysis of α- Chemical Technology and Biotechnology, 89(4), 508–515.
lactalbumin by Acid Protease A offers potential for β-lactoglobulin purification in Fu, Z., & Lin, J. (2017). An overview of bioinformatics tools and resources in allergy. Food
whey proteins. LWT-Food Science and Technology, 49(1), 117–122. Allergens: Methods and Protocols223–245.
Cheison, S. C., Brand, J., Leeb, E., & Kulozik, U. (2011). Analysis of the effect of tem- Fu, Y., Young, J. F., Løkke, M. M., Lametsch, R., Aluko, R. E., & Therkildsen, M. (2016).
perature changes combined with different alkaline pH on the β-lactoglobulin trypsin Revalorisation of bovine collagen as a potential precursor of angiotensin I-converting
hydrolysis pattern using MALDI-TOF-MS/MS. Journal of Agricultural and Food enzyme (ACE) inhibitory peptides based on in silico and in vitro protein digestions.
Chemistry, 59(5), 1572–1581. Journal of Functional Foods, 24, 196–206.
Cheison, S. C., & Kulozik, U. (2017). Impact of the environmental conditions and sub- Gabere, M. N., & Noble, W. S. (2017). Empirical comparison of web-based antimicrobial
strate pre-treatment on whey protein hydrolysis: A review. Critical Reviews in Food peptide prediction tools. Bioinformatics, 33(13), 1921–1929.
Science and Nutrition, 57(2), 418–453. Gabernet, G., Müller, A. T., Hiss, J. A., & Schneider, G. (2016). Membranolytic anticancer
Cheison, S. C., Lai, M. Y., Leeb, E., & Kulozik, U. (2011). Hydrolysis of β-lactoglobulin by peptides. Medicinal Chemistry Communications, 7(12), 2232–2245.
trypsin under acidic pH and analysis of the hydrolysates with MALDI–TOF–MS/MS. García-Mora, P., Martín-Martínez, M., Bonache, M. A., González-Múniz, R., Peñas, E.,
Food Chemistry, 125(4), 1241–1248. Frias, J., & Martinez-Villaluenga, C. (2017). Identification, functional gastrointestinal
Cheison, S. C., Leeb, E., Letzel, T., & Kulozik, U. (2011). Influence of buffer type and stability and molecular docking studies of Lentil peptides with dual antioxidant and
concentration on the peptide composition of trypsin hydrolysates of β-lactoglobulin. Angiotensin I converting enzyme inhibitory activities. Food Chemistry, 221, 464–472.
Food Chemistry, 125(1), 121–127. García-Tejedor, A., Sánchez-Rivera, L., Castelló-Ruiz, M., Recio, I., Salom, J. B., &
Cheison, S. C., Leeb, E., Toro-Sierra, J., & Kulozik, U. (2011). Influence of hydrolysis Manzanares, P. (2014). Novel antihypertensive lactoferrin-derived peptides produced
temperature and pH on the selective hydrolysis of whey proteins by trypsin and by Kluyveromyces marxianus: Gastrointestinal stability profile and in vivo angio-
potential recovery of native alpha-lactalbumin. International Dairy Journal, 21(3), tensin I-converting enzyme (ACE) inhibition. Journal of Agricultural and Food
166–171. Chemistry, 62(7), 1609–1616.
Cheison, S. C., Wang, Z., & Xu, S. Y. (2007). Multivariate strategy in screening of enzymes Garg, S., Nurgali, K., & Kumar Mishra, V. (2016). Food proteins as source of opioid
to be used for whey protein hydrolysis in an enzymatic membrane reactor. peptides-a review. Current Medicinal Chemistry, 23(9), 893–910.
International Dairy Journal, 17(4), 393–402. Gasteiger, E., Hoogland, C., Gattiker, A., Duvaud, S., Wilkins, M. R., Appel, R. D., &
Chen, W., Ding, H., Feng, P., Lin, H., & Chou, K. C. (2016). IACP: A sequence-based tool Bairoch, A. (2005). The proteomics protocols handbook. In J.M. Walker.
for identifying anticancer peptides. Oncotarget, 7(13), 16895. Gauthier, S. F., Pouliot, Y., & Saint-Sauveur, D. (2006). Immunomodulatory peptides
Chen, H. M., Muramoto, K., & Yamauchi, F. (1995). Structural analysis of antioxidative obtained by the enzymatic hydrolysis of whey proteins. International Dairy Journal,
peptides from soybean β-conglycinin. Journal of Agricultural and Food Chemistry, 16(11), 1315–1323.
43(3), 574–578. Graves, A. M., Hettiarachchy, N., Rayaprolu, S., Li, R., Horax, R., & Seo, H. S. (2016).
Chik, T., & Saufi, S. M. (2010). Mixed matrix membrane chromatography for bovine whey Bioactivity of a rice bran–derived peptide and its sensory evaluation and storage
protein fractionation. Doctoral Thesis available at < http://hdl.handle.net/10092/ stability in orange juice. Journal of Food Science, 81(4), H1010–H1015.
3647 > . Accessed June, 2017. Gu, Y., & Wu, J. (2013). LC–MS/MS coupled with QSAR modeling in characterising of
Chmiel, H. (Ed.). (2011). Bioprozesstechnik (pp. 356–372). Springer-Verlag. angiotensin I-converting enzyme inhibitory peptides from soybean proteins. Food
Chobert, J. M., Briand, L., Dufour, E., Dib, R., Dalgalarrondo, M., & Haertle, T. (1996). Chemistry, 141(3), 2682–2690.
How to increase β-lactoglobulin susceptibility to peptic hydrolysis. Journal of Food Gupta, C. (2012). Whey proteins: A novel source of bioceuticals. Middle-East Journal of
Biochemistry, 20(4), 439–462. Scientific Research, 12(3), 365–375.
Conesa, C., & FitzGerald, R. J. (2013). Total solids content and degree of hydrolysis in- Hafeez, Z., Cakir-Kiefer, C., Roux, E., Perrin, C., Miclo, L., & Dary-Mourot, A. (2014).
fluence proteolytic inactivation kinetics following whey protein hydrolysate manu- Strategies of producing bioactive peptides from milk proteins to functionalize fer-
facture. Journal of Agricultural and Food Chemistry, 61(42), 10135–10144. mented milk products. Food Research International, 63, 71–80.
Creusot, N., & Gruppen, H. (2008). Hydrolysis of whey protein isolate with Bacillus li- Hanke, A. T., & Ottens, M. (2014). Purifying biopharmaceuticals: Knowledge-based
cheniformis protease: Aggregating capacities of peptide fractions. Journal of chromatographic process development. Trends in Biotechnology, 32(4), 210–220.
Agricultural and Food Chemistry, 56(21), 10332–10339. Hati, S., Patel, N., Sakure, A., & Mandal, S. (2017). Influence of whey protein concentrate
Cui, H., Webber, M. J., & Stupp, S. I. (2010). Self-assembly of peptide amphiphiles: From on the production of antibacterial peptides derived from fermented milk by lactic
molecules to nanostructures to biomaterials. Peptide Science, 94(1), 1–18. acid bacteria. International Journal of Peptide Research and Therapeutics, 1–12.
Cziner, K., Virkki-Hatakka, T., Hurme, M., & Turunen, I. (2005). Evaluative approach for Hernández-Ledesma, B., Dávalos, A., Bartolomé, B., & Amigo, L. (2005). Preparation of
process development. Chemical Engineering & Technology, 28(12), 1490–1499. antioxidant enzymatic hydrolysates from α-lactalbumin and β-lactoglobulin.

71
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

Identification of active peptides by HPLC-MS/MS. Journal of Agricultural and Food Le Maux, S., Nongonierma, A. B., Barre, C., & FitzGerald, R. J. (2016). Enzymatic gen-
Chemistry, 53(3), 588–593. eration of whey protein hydrolysates under pH-controlled and non pH-controlled
Hernández-Ledesma, B., del Mar Contreras, M., & Recio, I. (2011). Antihypertensive conditions: Impact on physicochemical and bioactive properties. Food Chemistry, 199,
peptides: Production, bioavailability and incorporation into foods. Advances in Colloid 246–251.
and Interface Science, 165(1), 23–35. Le Maux, S., Nongonierma, A. B., & FitzGerald, R. J. (2017). Peptide composition and
Hernández-Ledesma, B., & Hsieh, C. C. (2017). Chemopreventive role of food-derived dipeptidyl peptidase IV inhibitory properties of β-lactoglobulin hydrolysates having
proteins and peptides: A review. Critical reviews in Food Science and Nutrition, 57(11), similar extents of hydrolysis while generated using different enzyme-to-substrate
2358–2376. ratios. Food Research International. http://dx.doi.org/10.1016/j.foodres.2017.05.012.
Hettiarachchy, N. S., Sato, K., Marshall, M. R., & Kannan, A. (Eds.). (2011). Bioactive food Leeb, E., Götz, A., Letzel, T., Cheison, S. C., & Kulozik, U. (2015). Influence of dena-
proteins and peptides: Applications in human health. CRC Press. turation and aggregation of β-lactoglobulin on its tryptic hydrolysis and the release of
Hippauf, F., Huettner, C., Lunow, D., Borchardt, L., Henle, T., & Kaskel, S. (2016). functional peptides. Food Chemistry, 187, 545–554.
Towards a continuous adsorption process for the enrichment of ACE-inhibiting Leeb, E., Holder, A., Letzel, T., Cheison, S. C., Kulozik, U., & Hinrichs, J. (2014).
peptides from food protein hydrolysates. Carbon, 107, 116–123. Fractionation of dairy based functional peptides using ion-exchange membrane ad-
Hippauf, F., Lunow, D., Borchardt, L., Henle, T., & Kaskel, S. (2014). Extraction of ACE- sorption chromatography and cross-flow electro membrane filtration. International
inhibiting dipeptides from protein hydrolysates using porous carbon materials. Dairy Journal, 38(2), 116–123.
Carbon, 77, 191–198. Leeb, E., Kulozik, U., & Cheison, S. (2011). Thermal pre-treatment of β-Lactoglobulin as a
Hoffmann, K. F. (1961). On the history of whey cures, especially in the 17th, 18th, and tool to steer enzymatic hydrolysis and control the release of peptides. Procedia Food
19th centuries. Medizinische Monatsschrift, 15, 411. Science, 1, 1540–1546.
Holder, A., Merath, C., Kulozik, U., & Hinrichs, J. (2015). Impact of diffusion, trans- Li, S., Hao, L., Bao, W., Zhang, P., Su, D., Cheng, Y., ... Yang, Y. (2016). A novel short
membrane pressure and the electrical field on peptide fractionation using cross-flow anionic antibacterial peptide isolated from the skin of Xenopus laevis. Archives of
electro membrane filtration. International Dairy Journal, 46, 31–38. Microbiology, 198(5), 473–482.
Holder, A., Scholz, S., Kulozik, U., & Hinrichs, J. (2013). Cross-flow electro membrane Li, E. W., & Mine, Y. (2004). Immunoenhancing effects of bovine glycomacropeptide and
filtration: Theory and application in the dairy industry. Chemie Ingenieur Technik, its derivatives on the proliferative response and phagocytic activities of human
85(8), 1193–1200. macrophagelike cells, U937. Journal of Agricultural and Food Chemistry, 52(9),
Hudson, M. J. (1995). Product development horizons-a view from industry. European 2704–2708.
Journal of Clinical Nutrition, 49, S64–S70. Li-Chan, E. C. (2015). Bioactive peptides and protein hydrolysates: Research trends and
Igarashi, Y., Eroshkin, A., Gramatikova, S., Gramatikoff, K., Zhang, Y., Smith, J. W., ... challenges for application as nutraceuticals and functional food ingredients. Current
Godzik, A. (2007). CutDB: A proteolytic event database. Nucleic Acids Research, Opinion in Food Science, 1, 28–37.
35(suppl_1), D546–D549. Liu, M., Bayjanov, J. R., Renckens, B., Nauta, A., & Siezen, R. J. (2010). The proteolytic
Ittrat, P., Chacho, T., Pholprayoon, J., Suttiwarayanon, N., & Charoenpanich, J. (2014). system of lactic acid bacteria revisited: A genomic comparison. BMC Genomics,
Application of agriculture waste as a support for lipase immobilization. Biocatalysis 11(1), 36.
and Agricultural Biotechnology, 3(3), 77–82. Ma, J. J., Mao, X. Y., Wang, Q., Yang, S., Zhang, D., Chen, S. W., & Li, Y. H. (2014). Effect
Iwaniak, A., Minkiewicz, P., Darewicz, M., Sieniawski, K., & Starowicz, P. (2016). BIOPEP of spray drying and freeze drying on the immunomodulatory activity, bitter taste and
database of sensory peptides and amino acids. Food Research International, 85, hygroscopicity of hydrolysate derived from whey protein concentrate. LWT-Food
155–161. Science and Technology, 56(2), 296–302.
Jayamuthunagai, J., Srisowmeya, G., Chakravarthy, M., & Gautam, P. (2017). D-Tagatose Madureira, A. R., Tavares, T., Gomes, A. M. P., Pintado, M. E., & Malcata, F. X. (2010).
production by permeabilized and immobilized Lactobacillus plantarum using whey Invited review: Physiological properties of bioactive peptides obtained from whey
permeate. Bioresource Technology, 235, 250–255. proteins. Journal of Dairy Science, 93(2), 437–455.
Jiang, B., Zhang, X., Yuan, Y., Qu, Y., & Feng, Z. (2017). Separation of antioxidant Maeda-Yamamoto, M. (in press). Development of functional agricultural products and use
peptides from pepsin hydrolysate of whey protein isolate by ATPS of EOPO co- of a new health claim system in Japan. Trends in Food Science & Technology. http://dx.
polymer (UCON)/Phosphate. Scientific Reports, 7(1), 13320. doi.org/10.1016/j.tifs.2017.08.011, available online 22 August 2017.
Jung, E., Kim, J., Kim, M., Jung, D. H., Rhee, H., Shin, J. M., ... Choi, YJ. (2007). Artificial Maehashi, K., & Huang, L. (2009). Bitter peptides and bitter taste receptors. Cellular and
neural network models for prediction of intestinal permeability of oligopeptides. Molecular Life Sciences, 66(10), 1661–1671.
BMC Bioinformatics, 8(1), 245. Mao, Y., Černigoj, U., Zalokar, V., Štrancar, A., & Kulozik, U. (2017). Production of β-
Kim, H. O., & Li-Chan, E. C. (2006). Quantitative structure-activity relationship study of lactoglobulin hydrolysates by monolith based immobilized trypsin reactors.
bitter peptides. Journal of Agricultural and Food Chemistry, 54(26), 10102–10111. Electrophoresis. http://dx.doi.org/10.1002/elps.201700188.
Klompong, V., Benjakul, S., Kantachote, D., & Shahidi, F. (2007). Antioxidative activity Maresca, P., & Ferrari, G. (2017). Modelling of the kinetics of bovine serum albumin
and functional properties of protein hydrolysate of yellow stripe trevally (Selaroides enzymatic hydrolysis assisted by high hydrostatic pressure. Food and Bioproducts
leptolepis) as influenced by the degree of hydrolysis and enzyme type. Food Chemistry, Processing, 105, 1–11.
102(4), 1317–1327. Martínez-Maqueda, D., Miralles, B., De Pascual-Teresa, S., Reverón, I., Muñoz, R., &
Korhonen, H. J. (2009a). Bioactive components in bovine milk. Bioactive Components in Recio, I. (2012). Food-derived peptides stimulate mucin secretion and gene expres-
Milk and Dairy Products, 15–42. sion in intestinal cells. Journal of Agricultural and Food Chemistry, 60(35), 8600–8605.
Korhonen, H. (2009b). Milk-derived bioactive peptides: From science to applications. Martínez-Maqueda, D., Miralles, B., Ramos, M., & Recio, I. (2013). Effect of β-lactoglo-
Journal of Functional Foods, 1(2), 177–187. bulin hydrolysate and β-lactorphin on intestinal mucin secretion and gene expression
Korhonen, H. J. (2012). Production and properties of health-promoting proteins and in human goblet cells. Food Research International, 54(1), 1287–1291.
peptides from bovine colostrum and milk. Cellular & Molecular Biology, 58(1), 26–38. McIntosh, G. H., Royle, P. J., Le Leu, R. K., Regester, G. O., Johnson, M. A., Grinsted, R. L.,
Korhonen, H. J., & Marnila, P. (2013). Milk bioactive proteins and peptides. Milk and Dairy ... Smithers, GW. (1998). Whey proteins as functional food ingredients? International
Products in Human Nutrition: Production, Composition and Health148–171. Dairy Journal, 8(5–6), 425–434.
Korhonen, H., & Pihlanto, A. (2006). Bioactive peptides: Production and functionality. Meisel, H. (1997). Biochemical properties of regulatory peptides derived from mil pro-
International Dairy Journal, 16(9), 945–960. teins. Peptide Science, 43(2), 119–128.
Korhonen, H., Pihlanto-Leppäla, A., Rantamäki, P., & Tupasela, T. (1998). Impact of Minkiewicz, P., Dziuba, J., Iwaniak, A., Dziuba, M., & Darewicz, M. (2008). BIOPEP
processing on bioactive proteins and peptides. Trends in Food Science & Technology, database and other programs for processing bioactive peptide sequences. Journal of
9(8), 307–319. AOAC International, 91(4), 965–980.
Lacroix, I. M., & Li-Chan, E. C. (2013). Inhibition of dipeptidyl peptidase (DPP)-IV and α- Mohan, A., Rajendran, S. R., He, Q. S., Bazinet, L., & Udenigwe, C. C. (2015).
glucosidase activities by pepsin-treated whey proteins. Journal of Agricultural and Encapsulation of food protein hydrolysates and peptides: A review. RSC Advances,
Food Chemistry, 61(31), 7500–7506. 5(97), 79270–79278.
Lacroix, I. M., & Li-Chan, E. C. (2014). Isolation and characterization of peptides with Mohan, A., Udechukwu, M. C., Rajendran, S. R., & Udenigwe, C. C. (2015). Modification
dipeptidyl peptidase-IV inhibitory activity from pepsin-treated bovine whey proteins. of peptide functionality during enzymatic hydrolysis of whey proteins. RSC Advances,
Peptides, 54, 39–48. 5(118), 97400–97407.
Lacroix, I. M., Meng, G., Cheung, I. W., & Li-Chan, E. C. (2016). Do whey protein-derived Mohanty, D., Jena, R., Choudhury, P. K., Pattnaik, R., Mohapatra, S., & Saini, M. R.
peptides have dual dipeptidyl-peptidase IV and angiotensin I-converting enzyme in- (2016). Milk derived antimicrobial bioactive peptides: A review. International Journal
hibitory activities? Journal of Functional Foods, 21, 87–96. of Food Properties, 19(4), 837–846.
Lafarga, T., Aluko, R. E., Rai, D. K., O'Connor, P., & Hayes, M. (2016). Identification of Mooney, C., Haslam, N. J., Pollastri, G., & Shields, D. C. (2012). Towards the improved
bioactive peptides from a papain hydrolysate of bovine serum albumin and assess- discovery and design of functional peptides: Common features of diverse classes
ment of an antihypertensive effect in spontaneously hypertensive rats. Food Research permit generalized prediction of bioactivity. PLoS ONE, 7(10), e45012.
International, 81, 91–99. Mullally, M. M., Meisel, H., & FitzGerald, R. J. (1997). Identification of a novel angio-
Lafarga, T., & Hayes, M. (2017). Bioactive protein hydrolysates in the functional food tensin-I-converting enzyme inhibitory peptide corresponding to a tryptic fragment of
ingredient industry: Overcoming current challenges. Food Reviews International, bovine β-lactoglobulin. FEBS Letters, 402(2–3), 99–101.
33(3), 217–246. Mundi, S., & Aluko, R. E. (2014). Inhibitory properties of kidney bean protein hydrolysate
Lafarga, T., Rai, D. K., O'Connor, P., & Hayes, M. (2016). Generation of bioactive hy- and its membrane fractions against renin, angiotensin converting enzyme, and free
drolysates and peptides from bovine hemoglobin with in vitro renin, angiotensin-I- radicals. Austin Journal of Nutrition and Food Sciences, 2, 1008.
converting enzyme and dipeptidyl peptidase-IV inhibitory activities. Journal of Food Ney, K. H. (1971). Prediction of bitterness of peptides from their amino acid composition.
Biochemistry, 40(5), 673–685. Zeitschrift für Lebensmittel-Untersuchung und -Forschung, 147, 64–68.
Lamiable, A., Thévenet, P., Rey, J., Vavrusa, M., Derreumaux, P., & Tufféry, P. (2016). Ney, D. M., & Etzel, M. R. (2017). Designing medical foods for inherited metabolic dis-
PEP-FOLD3: Faster de novo structure prediction for linear peptides in solution and in orders: Why intact protein is superior to amino acids. Current Opinion in Biotechnology,
complex. Nucleic Acids Research, 44(W1), W449–W454. 44, 39–45.

72
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

Nfor, B. K., Verhaert, P. D., van der Wielen, L. A., Hubbuch, J., & Ottens, M. (2009). Santiago-López, L., Hernández-Mendoza, A., Vallejo-Cordoba, B., Mata-Haro, V., &
Rational and systematic protein purification process development: The next genera- González-Córdova, A. F. (2016). Food-derived immunomodulatory peptides. Journal
tion. Trends in Biotechnology, 27(12), 673–679. of the Science of Food and Agriculture, 96(11), 3631–3641.
Ngarize, S., Herman, H., Adams, A., & Howell, N. (2004). Comparison of changes in the Sarmadi, B. H., & Ismail, A. (2010). Antioxidative peptides from food proteins: A review.
secondary structure of unheated, heated, and high-pressure-treated β-lactoglobulin Peptides, 31(10), 1949–1956.
and ovalbumin proteins using Fourier transform Raman spectroscopy and self-de- Sawin, E. A., De Wolfe, T. J., Aktas, B., Stroup, B. M., Murali, S. G., Steele, J. L., & Ney, D.
convolution. Journal of Agricultural and Food Chemistry, 52(21), 6470–6477. M. (2015). Glycomacropeptide is a prebiotic that reduces Desulfovibrio bacteria,
Nielsen, S. D., Beverly, R. L., Qu, Y., & Dallas, D. C. (2017). Milk bioactive peptide da- increases cecal short-chain fatty acids, and is anti-inflammatory in mice. American
tabase: A comprehensive database of milk protein-derived bioactive peptides and Journal of Physiology-Gastrointestinal and Liver Physiology, 309(7), G590–G601.
novel visualization. Food Chemistry, 232, 673–682. Sikarwar, R. L. S., & Kaushik, J. P. (1993). Folk medicines of the Morena district, Madhya
Nongonierma, A. B., & FitzGerald, R. J. (2016). Structure activity relationship modelling Pradesh, India. International Journal of Pharmacognosy, 31(4), 283–287.
of milk protein-derived peptides with dipeptidyl peptidase IV (DPP-IV) inhibitory Singh, P. K., Sharma, S., Kumari, A., & Korpole, S. (2014). A non-pediocin low molecular
activity. Peptides, 79, 1–7. weight antimicrobial peptide produced by Pediococcus pentosaceus strain IE-3 shows
Nongonierma, A. B., Le Maux, S., Hamayon, J., & FitzGerald, R. J. (2016). Strategies for increased activity under reducing environment. BMC Microbiology, 14(1), 226.
the release of dipeptidyl peptidase IV (DPP-IV) inhibitory peptides in an enzymatic Smithers, G. W. (2015). Whey-ing up the options - Yesterday, today and tomorrow.
hydrolyzate of α-lactalbumin. Food & Function, 7(8), 3437–3443. International Dairy Journal, 48, 2–14.
O’Loughlin, I. B., Kelly, P. M., Murray, B. A., FitzGerald, R. J., & Brodkorb, A. (2015). Suwal, S., Rozoy, É., Manenda, M., Doyen, A., & Bazinet, L. (2017). Comparative study of
Molecular characterization of whey protein hydrolysate fractions with ferrous che- in-situ and ex-situ enzymatic hydrolysis of milk protein and separation of bioactive
lating and enhanced iron solubility capabilities. Journal of Agricultural and Food peptides in an electromembrane reactor. ACS Sustainable Chemistry & Engineering,
Cchemistry, 63(10), 2708–2714. 5(6), 5330–5340.
O'Keeffe, M. B., Conesa, C., & FitzGerald, R. J. (2017). Identification of angiotensin Szymańska, K., Pietrowska, M., Kocurek, J., Maresz, K., Koreniuk, A., Mrowiec-Białoń, J.,
converting enzyme inhibitory and antioxidant peptides in a whey protein concentrate ... Jarzębski, A. (2016). Low back-pressure hierarchically structured multichannel
hydrolysate produced at semi-pilot scale. International Journal of Food Science & microfluidic bioreactors for rapid protein digestion–proof of concept. Chemical
Technology, 52(8), 1751–1759. Engineering Journal, 287, 148–154.
O'Loughlin, I. B., Murray, B. A., FitzGerald, R. J., Brodkorb, A., & Kelly, P. M. (2014). Tavano, O. L. (2013). Protein hydrolysis using proteases: An important tool for food
Pilot-scale production of hydrolysates with altered bio-functionalities based on biotechnology. Journal of Molecular Catalysis B: Enzymatic, 90, 1–11.
thermally-denatured whey protein isolate. International Dairy Journal, 34(1), Tavares, T. G., Amorim, M., Gomes, D., Pintado, M. E., Pereira, C. D., & Malcata, F. X.
146–152. (2012). Manufacture of bioactive peptide-rich concentrates from whey:
Pai, T. W., Chang, M. D. T., Tzou, W. S., Su, B. H., Wu, P. C., Chang, H. T., & Chou, W. I. Characterization of pilot process. Journal of Food Engineering, 110(4), 547–552.
(2006). RE MUS: A tool for identification of unique peptide segments as epitopes. Tavares, T., del Mar Contreras, M., Amorim, M., Pintado, M., Recio, I., & Malcata, F. X.
Nucleic Acids Research, 34(suppl_2), W198–W201. (2011). Novel whey-derived peptides with inhibitory effect against angiotensin-
Pan, D., Cao, J., Guo, H., & Zhao, B. (2012). Studies on purification and the molecular converting enzyme: In vitro effect and stability to gastrointestinal enzymes. Peptides,
mechanism of a novel ACE inhibitory peptide from whey protein hydrolysate. Food 32(5), 1013–1019.
Chemistry, 130(1), 121–126. Théolier, J., Fliss, I., Jean, J., & Hammami, R. (2014). Antimicrobial peptides of dairy
Panchaud, A., Affolter, M., & Kussmann, M. (2012). Mass spectrometry for nutritional proteins: From fundamental to applications. Food Reviews International, 30(2),
peptidomics: How to analyze food bioactives and their health effects. Journal of 134–154.
Proteomics, 75(12), 3546–3559. Thomä-Worringer, C., Sørensen, J., & López-Fandiño, R. (2006). Health effects and
Pandey, M., Kapila, R., & Kapila, S. (2018). Osteoanabolic activity of whey-derived anti- technological features of caseinomacropeptide. International Dairy Journal, 16(11),
oxidative (MHIRL and YVEEL) and angiotensin-converting enzyme inhibitory (YLLF, 1324–1333.
ALPMHIR, IPA and WLAHK) bioactive peptides. Peptides, 99, 1–7. Tian, M., Fang, B., Jiang, L., Guo, H., Cui, J., & Ren, F. (2015). Structure-activity re-
Pasupuleti, V. K., & Braun, S. (2008). State of the art manufacturing of protein hydro- lationship of a series of antioxidant tripeptides derived from β-Lactoglobulin using
lysates. Protein hydrolysates in biotechnology (pp. 1–32). Netherlands: Springer. QSAR modeling. Dairy Science & Technology, 95(4), 451–463.
Pellegrini, A., Dettling, C., Thomas, U., & Hunziker, P. (2001). Isolation and character- Tong, J., Li, L., Bai, M., & Li, K. (2017). A new descriptor of amino acids-SVGER and its
ization of four bactericidal domains in the bovine β-lactoglobulin. Biochimica et applications in peptide QSAR. Molecular Informatics, 36(5–6).
Biophysica Acta (BBA)-General Subjects, 1526(2), 131–140. Tsai, C. J., de Laureto, P. P., Fontana, A., & Nussinov, R. (2002). Comparison of protein
Peña-Ramos, E. A., Xiong, Y. L., & Arteaga, G. E. (2004). Fractionation and character- fragments identified by limited proteolysis and by computational cutting of proteins.
isation for antioxidant activity of hydrolysed whey protein. Journal of the Science of Protein Science, 11(7), 1753–1770.
Food and Agriculture, 84(14), 1908–1918. Tu, M., Feng, L., Wang, Z., Qiao, M., Shahidi, F., Lu, W., & Du, M. (2017). Sequence
Persico, M., Mikhaylin, S., Doyen, A., Firdaous, L., Hammami, R., & Bazinet, L. (2016). analysis and molecular docking of antithrombotic peptides from casein hydrolysate
How peptide physicochemical and structural characteristics affect anion-exchange by trypsin digestion. Journal of Functional Foods, 32, 313–323.
membranes fouling by a tryptic whey protein hydrolysate. Journal of Membrane Tulipano, G., Faggi, L., Nardone, A., Cocchi, D., & Caroli, A. M. (2015). Characterisation
Science, 520, 914–923. of the potential of β-lactoglobulin and α-lactalbumin as sources of bioactive peptides
Pieroni, A., & Giusti, M. E. (2008). The remedies of the folk medicine of the Croatians affecting incretin function: In silico and in vitro comparative studies. International
living in Ćićarija, northern Istria. Collegium Antropologicum, 32(2), 623–627. Dairy Journal, 48, 66–72.
Pieroni, A., & Gray, C. (2008). Herbal and food folk medicines of the Russlanddeutschen Udechukwu, M. C., Downey, B., & Udenigwe, C. C. (2018). Influence of structural and
living in Künzelsau/Taläcker, South-Western Germany. Phytotherapy Research, 22(7), surface properties of whey-derived peptides on zinc-chelating capacity, and in vitro
889–901. gastric stability and bioaccessibility of the zinc-peptide complexes. Food Chemistry,
Pihlanto, A. (2006). Antioxidative peptides derived from milk proteins. International Dairy 240, 1227–1232.
Journal, 16(11), 1306–1314. Udenigwe, C. C. (2014). Bioinformatics approaches, prospects and challenges of food
Pihlanto-Leppälä, A. (2000). Bioactive peptides derived from bovine whey proteins: bioactive peptide research. Trends in Food Science & Technology, 36(2), 137–143.
Opioid and ace-inhibitory peptides. Trends in Food Science & Technology, 11(9), Uluko, H., Zhang, S., Liu, L., Tsakama, M., Lu, J., & Lv, J. (2015). Effects of thermal,
347–356. microwave, and ultrasound pretreatments on antioxidative capacity of enzymatic
Pihlanto-Leppälä, A., Koskinen, P., Piilola, K., Tupasela, T., & Korhonen, H. (2000). milk protein concentrate hydrolysates. Journal of Functional Foods, 18, 1138–1146.
Angiotensin I-converting enzyme inhibitory properties of whey protein digests: UniProt Consortium (2008). The universal protein resource (UniProt). Nucleic Acids
Concentration and characterization of active peptides. Journal of Dairy Research, Research, 36(suppl 1), D190–D195.
67(1), 53–64. Valencia, P., Espinoza, K., Ceballos, A., Pinto, M., & Almonacid, S. (2015). Novel mod-
Pihlanto-Leppälä, A., Paakkari, I., Rinta-Koski, M., & Antila, P. (1997). Bioactive peptide eling methodology for the characterization of enzymatic hydrolysis of proteins.
derived from in vitro proteolysis of bovine β-lactoglobulin and its effect on smooth Process Biochemistry, 50(4), 589–597.
muscle. Journal of Dairy Research, 64(1), 149–155. Vermeirssen, V., Van Camp, J., & Verstraete, W. (2004). Bioavailability of angiotensin I
Pina, A. S., & Roque, A. C. A. (2009). Studies on the molecular recognition between converting enzyme inhibitory peptides. British Journal of Nutrition, 92(3), 357–366.
bioactive peptides and angiotensin-converting enzyme. Journal of Molecular Vogel, H. J., Schibli, D. J., Jing, W., Lohmeier-Vogel, E. M., Epand, R. F., & Epand, R. M.
Recognition, 22(2), 162–168. (2002). Towards a structure-function analysis of bovine lactoferricin and related
Prazeres, A. R., Carvalho, F., & Rivas, J. (2012). Cheese whey management: A review. tryptophan-and arginine-containing peptides. Biochemistry and Cell Biology, 80(1),
Journal of Environmental Management, 110, 48–68. 49–63.
Roberts, M. J., Bentley, M. D., & Harris, J. M. (2012). Chemistry for peptide and protein Vorob’ev, M. M., Butré, C. I., Sforza, S., Wierenga, P. A., & Gruppen, H. (2017).
PEGylation. Advanced Drug Delivery Reviews, 64, 116–127. Demasking kinetics of peptide bond cleavage for whey protein isolate hydrolysed by
Rodríguez-Carrio, J., Fernández, A., Riera, F. A., & Suárez, A. (2014). Immunomodulatory Bacillus licheniformis protease. Journal of Molecular Catalysis B: Enzymatic, http://dx.
activities of whey β-lactoglobulin tryptic-digested fractions. International Dairy doi.org/10.1016/j.molcatb.2017.03.005.
Journal, 34(1), 65–73. Voswinkel, L., Etzel, M. R., & Kulozik, U. (2017). Adsorption of beta-lactoglobulin in
Sah, B. N. P., Vasiljevic, T., McKechnie, S., & Donkor, O. N. (2017). Antioxidative and anion exchange membrane chromatography versus the contacting mode and tem-
antibacterial peptides derived from bovine milk proteins. Critical Reviews in Food perature. LWT-Food Science and Technology, 79, 78–83.
Science and Nutrition, 1–15. Voswinkel, L., & Kulozik, U. (2011). Fractionation of whey proteins by means of mem-
Sanchón, J., Fernández-Tomé, S., Miralles, B., Hernández-Ledesma, B., Tomé, D., brane adsorption chromatography. Procedia Food Science, 1, 900–907.
Gaudichon, C., & Recio, I. (2018). Protein degradation and peptide release from milk Voswinkel, L., & Kulozik, U. (2014). Fractionation of all major and minor whey proteins
proteins in human jejunum. Comparison with in vitro gastrointestinal simulation. with radial flow membrane adsorption chromatography at lab and pilot scale.
Food Chemistry, 239, 486–494. International Dairy Journal, 39(1), 209–214.

73
A. Dullius et al. Journal of Functional Foods 42 (2018) 58–74

Wali, A., Ma, H., Aadil, R. M., Zhou, C., Rashid, M. T., & Liu, X. (2017). Effects of mul- 2556–2566.
tifrequency ultrasound pretreatment on the enzymolysis, ACE inhibitory activity, and Xie, J., Xu, Z., Zhou, S., Pan, X., Cai, S., Yang, L., & Mei, H. (2013). The VHSE-based
the structure characterization of rapeseed protein. Journal of Food Processing and prediction of proteasomal cleavage sites. PLoS ONE, 8(9), e74506.
Preservation. http://dx.doi.org/10.1111/jfpp.13413. Yadav, J. S. S., Yan, S., Pilli, S., Kumar, L., Tyagi, R. D., & Surampalli, R. Y. (2015). Cheese
Wang, T. Y., Hsieh, C. H., Hung, C. C., Jao, C. L., Lin, P. Y., Hsieh, Y. L., & Hsu, K. C. whey: A potential resource to transform into bioprotein, functional/nutritional pro-
(2017). A study to evaluate the potential of an in silico approach for predicting di- teins and bioactive peptides. Biotechnology Advances, 33(6), 756–774.
peptidyl peptidase-IV inhibitory activity in vitro of protein hydrolysates. Food Yamada, K., Sato-Mito, N., Nagata, J., & Umegaki, K. (2008). Health claim evidence re-
Chemistry, 234, 431–438. quirements in Japan. The Journal of Nutrition, 138(6), 1192S–1198S.
Welsh, G., Ryder, K., Brewster, J., Walker, C., Mros, S., Bekhit, A. E. D. A., ... Carne, A. Yu, Y. J., Amorim, M., Marques, C., Calhau, C., & Pintado, M. (2016). Effects of whey
(2017). Comparison of bioactive peptides prepared from sheep cheese whey using a peptide extract on the growth of probiotics and gut microbiota. Journal of Functional
food-grade bacterial and a fungal protease preparation. International Journal of Food Foods, 21, 507–516.
Science & Technology, 52(5), 1252–1259. Zhang, Q. X., Wu, H., Ling, Y. F., & Lu, R. R. (2013). Isolation and identification of an-
Winkelnkemper, T., & Schembecker, G. (2010). Purification performance index and se- tioxidant peptides derived from whey protein enzymatic hydrolysate by consecutive
paration cost indicator for experimentally based systematic downstream process chromatography and Q-TOF MS. Journal of Dairy Research, 80(3), 367–373.
development. Separation and Purification Technology, 72(1), 34–39. Zhong, J., Cai, X., Liu, C., Liu, W., Xu, Y., & Luo, S. (2016). Purification and conforma-
Wu, W., Zhang, M., Sun, C., Brennan, M., Li, H., Wang, G., ... Wu, H. (2016). Enzymatic tional changes of bovine PEGylated β-lactoglobulin related to antigenicity. Food
preparation of immunomodulatory hydrolysates from defatted wheat germ (Triticum Chemistry, 199, 387–392.
Vulgare) globulin. International Journal of Food Science & Technology, 51(12),

74

You might also like