Next Article in Journal
Unsteady Flow Characteristics of Rotating Stall and Surging in a Backward Centrifugal Fan at Low Flow-Rate Conditions
Next Article in Special Issue
Application of Detrended Fluctuation Analysis and Yield Stability Index to Evaluate Near Infrared Spectra of Green and Roasted Coffee Samples
Previous Article in Journal
Determination of Vitamins K1, K2 MK-4, MK-7, MK-9 and D3 in Pharmaceutical Products and Dietary Supplements by TLC-Densitometry
Previous Article in Special Issue
Magnetic Fields in Food Processing Perspectives, Applications and Action Models
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Production of Liquid Milk Protein Concentrate with Antioxidant Capacity, Angiotensin Converting Enzyme Inhibitory Activity, Antibacterial Activity, and Hypoallergenic Property by Membrane Filtration and Enzymatic Modification of Proteins

1
Department of Food Engineering, Faculty of Food Science, Szent István University, Ménesi st 44, HU-1118 Budapest, Hungary
2
Department of Refrigeration and Livestock Products Technology, Faculty of Food Science, Szent István University, Ménesi út 43-45, HU-1118 Budapest, Hungary
3
Department of Food Microbiology and Biotechnology, Faculty of Food Science, Szent István University, Somlói st 14–16, HU-1118 Budapest, Hungary
4
Department of Applied Chemistry, Faculty of Food Science, Szent István University, Villányi út 29-43, HU-1118 Budapest, Hungary
5
Division of Clinical Physiology, Department of Cardiology, Faculty of Medicine, University of Debrecen, Móricz Zsigmond Str 22, HU-4032 Debrecen, Hungary
6
Department of Biology, National Agricultural Research and Innovation Center, Food Science Research Institute, Herman Ottó út 15, HU-1022 Budapest, Hungary
7
Department of Physics and Control, Faculty of Food Science, Szent István University, Somlói street 14-16, HU-1118 Budapest, Hungary
*
Authors to whom correspondence should be addressed.
Processes 2020, 8(7), 871; https://doi.org/10.3390/pr8070871
Submission received: 3 June 2020 / Revised: 2 July 2020 / Accepted: 10 July 2020 / Published: 18 July 2020
(This article belongs to the Special Issue Processing Foods: Process Optimization and Quality Assessment)

Abstract

:
Liquid milk protein concentrate with different beneficial values was prepared by membrane filtration and enzymatic modification of proteins in a sequential way. In the first step, milk protein concentrate was produced from ultra-heat-treated skimmed milk by removing milk serum as permeate. A tubular ceramic-made membrane with filtration area 5 × 10−3 m2 and pore size 5 nm, placed in a cross-flow membrane house, was adopted. Superior operational strategy in filtration process was herein: trans-membrane pressure 3 bar, retention flow rate 100 L·h−1, and implementation of a static turbulence promoter within the tubular membrane. Milk with concentrated proteins from retentate side was treated with the different concentrations of trypsin, ranging from 0.008–0.064 g·L−1 in individual batch-mode operations at temperature 40 °C for 10 min. Subsequently, inactivation of trypsin in reaction was done at a temperature of 70 °C for 30 min of incubation. Antioxidant capacity in enzyme-treated liquid milk protein concentrate was measured with the Ferric reducing ability of plasma assay. The reduction of angiotensin converting enzyme activity by enzyme-treated liquid milk protein concentrate was measured with substrate (Abz-FRK(Dnp)-P) and recombinant angiotensin converting enzyme. The antibacterial activity of enzyme-treated liquid milk protein concentrate towards Bacillus cereus and Staphylococcus aureus was tested. Antioxidant capacity, anti-angiotensin converting enzyme activity, and antibacterial activity were increased with the increase of trypsin concentration in proteolytic reaction. Immune-reactive proteins in enzyme-treated liquid milk protein concentrate were identified with clinically proved milk positive pooled human serum and peroxidase-labelled anti-human Immunoglobulin E. The reduction of allergenicity in milk protein concentrate was enzyme dose-dependent.

1. Introduction

For years, different types of non-fermented and fermented dairy-based food formulas are received great attention among different communities. As time progressed, dairy industries tried their best to improve the quality of dairy-based formulas to fulfill the expectations of consumers [1]. For industrial production of fermented dairy products, milk with a standardized amount of protein is necessary to maintain the quality of products [2,3]. Different dairy-based protein concentrates, such as milk protein concentrate, milk protein isolate, whey protein concentrate, whey protein isolate, micellar casein concentrate, micellar casein isolate, whey concentrate, and selectively demineralized whey concentrate are widely used in the food and biopharmaceutical industries [4]. Milk protein concentrate is well accepted among all communities because it is an abundant source of the various kinds of proteins, including micellar casein, whey proteins, and glycomacropeptide, and has significantly lower amounts of lactose and fat compared to whey protein concentrate and whole milk protein concentrate, respectively [5]. Therefore, it is popularly used to prepare infant formula, protein bar, yogurt, recombined cheese, cultured product, frozen dessert, weight management products, and sports formulas [6]. To produce dairy-based protein concentrate, large-scale production plants with different unit operations are requirement. It may be noted that production of milk varies throughout the year. During the spring season, milk production is quite high compared to the fall season. To balance economic competitiveness, small-scale and medium-scale dairy industries avoid expensive processing steps, such as evaporation and drying to prepare dairy-based protein concentrate in powder form, and prefer to use liquid milk protein concentrate for manufacturing fresh cultured-food products [7,8].
In the dairy industry, application of membrane technology is noteworthy. Membrane technology is used for preparing concentrated milk proteins, fractionation of dairy proteins, demineralization of whey, and removal of microbial count in milk [9,10,11]. In some cases, ultrafiltration or nanofiltration operated with diafiltration mode was adopted to achieve high protein concentrate and avoid membrane fouling [12,13,14]. Some limitations in this context are reported. The limitations are (a) development of gel layer (concentration polarization) on the membrane surface and subsequent membrane fouling, and (b) high energy consumption. During filtration, due to deposition of solute molecules on the membrane surface, concentration polarization take place on the membrane surface. Because of this, permeate flux is reduced in drastic way [15,16,17]. However, increase of trans-membrane pressure (TMP) or fluid flow through a mechanical pump reduces the development of gel layer on the membrane surface; there is a debating issue about high energy consumption [18,19]. Therefore, it may feel that an efficient membrane separation process and its operational strategy are needed to explore to produce liquid milk protein concentrate.
However, milk has gained a great attention around the globe due to the presence of the different types of proteins (αS1-casein, αS2-casein, β-casein, κ-casein, γ-casein, immunoglobulin, bovine serum albumin, lactoferrin, α-lactalbumin, and β-lactoglobulin), lactose, vitamins (vitamin A, vitamin E, ascorbic acid, riboflavin, vitamin B6, nicotinic acid, pantothenic acid, and thiamin) and minerals (calcium, magnesium, phosphorus, potassium, selenium, and zinc) [20]. The milk sensitive community frequently experiences with the symptoms of immunoglobulin-mediated milk protein allergies, in some cases [21,22]. Due to the presence of Immunoglobulin E- and Immunoglobulin G- binding epitopes, milk proteins are listed among the “big 8” allergens [23,24]. Milk allergens provoke mild symptoms to life-threatening biochemical outcomes, including severe enterocolitis atopic eczema and immediate immunoglobulin-mediated systematic multisystem reactions [25]. Milk is not recommended in the diet chart due to the presence of saturated fatty acids—those contribute heart disease [26,27], type 2 diabetes, and Alzheimer’s disease [28,29]. Furthermore, due to the absence of lactase, a hydrolytic enzyme in brush border of epithelial cells in the small intestine, the milk sensitive community frequently suffers with symptoms of lactose maldigestion [20,21]. However, concentrations of lactose and fat in milk protein concentrate are significantly low; in some cases, food formulas fortified with milk protein concentrate offer immunoglobulin-mediated allergies among people of all ages [30,31]. A plethora of literature about thermal and non-thermal processing technologies have been adopted to combat milk protein allergens [32,33]. The reduction in protein allergenicity in the molecular basis is the destruction of structure of epitopes. Applications of high pressure- [34,35], heat- [36,37,38], microwave- [39], and membrane bioreactors [40,41] were implemented for the reduction of allergenic sequences in milk proteins. In some cases, physical- [42,43,44] and enzymatic- [45,46,47,48] modifications of proteins have been adopted for a similar objective. Furthermore, combined physical and biochemical technologies have been adopted for the reduction of milk protein allergens [49,50,51,52,53,54,55,56]. In some cases, new epitopes (neoepitopes) or hidden epitopes may even be produced during cow milk processing due to denaturation of native allergen (cryptotopes) [57]. Realizing advantages and disadvantages of mentioned technologies, it may feel that enzymatic hydrolysis of allergenic epitopes in protein sequences may be an effective attempt to reduce milk protein allergens. Besides the elimination of their allergenic potentiality, modification of milk proteins through enzymatic routes may alter their functional properties, because peptides with unique amino acids in C- and N- terminal positions are produced through enzymatic hydrolysis of peptide bonds in milk proteins. Furthermore, the enzymatic modification of milk proteins may generate new antigenic substances, which may offer immunomodulation, and provide extra health benefits [58]. However, lots of information about the reduction of allergenic epitopes in milk proteins through an enzymatic route are stored in scholarly databases; its production in industrial scale is limited [59]. The challenging issues in enzyme-mediated process are (a) high cost of enzymes, and (b) find out suitable operating process parameters in enzyme-mediated processes. Therefore, it can feel that an investigation is needed to find out the minimum amount of enzymes, which is responsible for reducing a significant amount (>99.9%) of the allergenic sequence and improve the functional activities of milk protein concentrate. Trypsin is an endopeptidase generally found in the pancreas of mammalians, and cleaves at the carboxyl terminal side of arginine and lysine amino acid residues, except arginyl-proline and lysyl-proline bonds. It is popularly used for preparation of dairy formulations with lower antigenic activities [60,61]. As the catalytic activity of trypsin is quite high (relative activity 99%) at pH 7 and may be able to change the biological activity of proteins and peptides [62,63,64], it was used in this investigation.
From the above discussion, one can realize that efforts are needed to reduce the limitations of milk protein concentrate production and dairy product consumption. In this investigation, an attempt was considered to develop liquid milk protein concentrate from ultra-heat-treated skimmed milk with antioxidant capacity, angiotensin converting enzyme inhibitory activity, antibacterial activity, and hypoallergenic property by membrane filtration and enzymatic modification of proteins in a sequential way. In the present investigation, membrane filtration process was adopted to increase the protein concentration in milk by reducing the milk serum as permeate and, subsequently, trypsin was adopted to hydrolyze the concentrated liquid milk proteins, obtained at the retentate side. Membrane filtration process itself cannot change the structural and biological activities of milk proteins. Peptides, produced by enzymatic hydrolysis of milk proteins with unique C- and N-terminal amino acids, peptide length, and amino acid sequence, offer distinguishing biological activities. Furthermore, allergenic activity of proteins is reduced due to enzymatic cleavage in allergenic epitopes within the amino acid sequence in protein.

2. Materials and Methods

2.1. Chemicals and Reagents

Lyophilized trypsin (≥27.78 units per mg of solid at temperature 25 °C) from bovine pancreas, Bradford reagent, bovine serum albumin, casein, α-lactalbumin and β-lactoglobulin from bovine milk, Abz-FRK(Dnp)-P, peroxidase-labelled anti-human Immunoglobulin E, 2,4,6-Tris(2-pyridyl)-s-triazine (≥98%), 4-chloronaphtol (≥98%), hydrogen peroxide (≥98%), ethanol (≥99%), and phosphate buffered saline solution were purchased from the Sigma-Aldrich (Sigma-Aldrich, Schnelldorf, Germany). Ultrasil P3-11 was purchased from Ecolab-Hygiene Kft (Ecolab-Hygiene Kft, Budapest, Hungary). Citric acid (99%), hydrochloric acid (≥99%), urea (≥99%), dithiothreitol (DTT) and sodium hydroxide (≥99%) were purchased from Reanal Kft (Reanal Kft, Budapest, Hungary). Ferric chloride (≥99%), sodium acetate (anhydrous, ≥99%), sodium chloride (≥99%), zinc chloride (≥99%), bacteriological agar powder, soybean casein digestive medium and ascorbic acid (99.7%) were procured from Merck (Merck, Darmstadt, Germany). Sodium-dodecyl sulphate (≥99%), acrylamide (≥99%), ammonium persulfate (≥99%), bis-acrylamide (≥99%), tetramethylethylenediamine (≥99%), tris(hydroxymethyl)aminomethane hydrochloride (TRIS HCl), ethyl alcohol (≥99%), glycine (≥99%), coomassie blue stain R250 (≥99%), acetic acid (≥99%), glycerol (≥99%), isopropanol (≥99%), 2β-mercaptoethanol (≥99%), and bromophenol blue (≥99%) were procured from Bio-Rad (Bio-Rad, Hercules, USA). High performance liquid chromatography mass spectrometry (HPLC-MS)-grade acetonitrile, formic acid, and trisodium citrate (≥99%) were purchased from VWR International Ltd. (VWR International Ltd., Debrecen, Hungary). Recombinant angiotensin converting enzyme was kindly provided by Division of Clinical Physiology, Institute of Cardiology, University of Debrecen (University of Debrecen, Debrecen, Hungary). Bacillus cereus and Staphylococcus aureus ATCC 6538 were collected from the Strain collection unit of Szent István University (Szent István University, Budapest, Hungary). Milli-Q ultrapure deionized water (18.2 MΩ·cm) was obtained from Milli-Q Synergy/Elix water purification system (Merck-Millipore, Molsheim, France) and used throughout the experiment.

2.2. Ultra-Heat-Treated Skimmed Cow Milk

Ultra-heat-treated skimmed cow milk was procured from local supermarkets, in and around Budapest, Hungary. Concentrations of protein, lactose, and fat in milk were in average 31 ± 0.16 g·L−1, 47 ± 0.15 g·L−1 and 1 ± 0.02 g·L−1, respectively. Average pH of milk was 6.8. Milk was stored in a refrigerator at temperature 10 °C.

2.3. Production of Hypoallergenic Liquid Milk Protein Concentrate with Functional Values

An attempt was considered to develop a process to produce allergen-free liquid milk protein concentrate with functional values, such as antioxidant capacity, angiotensin converting enzyme inhibitory activity, and antibacterial activity. Combination of different physical- and biochemical- based technologies were adopted for this purpose (Figure 1).

2.3.1. Concentrate Milk Proteins in Ultra-Heat-Treated Skimmed Milk by Membrane Technology

De-watering (remove of milk serum) of ultra-heat-treated skimmed milk was performed by a tubular nanofiltration membrane with active filtration area 5 × 10−3 m2 and pore size 5 nm (Pall Corporation, Crailsheim, Germany), placed in a stainless steel-made cross-flow membrane module (Figure 2). The active layer, support layer, length, inner diameter, and outer diameter of the membrane were titanium oxide, aluminum oxide, 250 mm, 7 mm, and 10 mm, respectively. In the membrane module, feed flow rate was controlled by a centrifugal pump (Verder Hungary Kft, Budapest, Hungary). Flow rate of fluid (milk, water) in the membrane module was also controlled by a rotameter at a retentate flow channel and a bypass channel at the inlet channel of the membrane module. TMP of the membrane module was controlled by pressure gauges, fitted at inlet and retentate flow channels of the membrane module. A mechanical agitator was fitted inside of the storage tank of the membrane module. Temperature in the storage tank of the membrane module was maintained by a temperature sensor and automated circulation of warm/cold water within the water jacket.
A mechanical device, known as static turbulence promoter, made of stainless steel (SS316), was inserted within the membrane tube. Detailed geometry of the static turbulence promoter is mentioned in an earlier publication [65]. To investigate the effects of process parameters in filtration process, different TMPs, such as 2 bar and 3 bar, and retention flow rates (RFRs), such as 100 L·h−1 and 200 L·h−1, were used with or without the static turbulence promoter. Each membrane filtration experiment started with 1 L of ultra-heat-treated skimmed milk, and volume reduction factor 2 was considered. Membrane filtration process was performed with a batch recirculation mode. Constant volume of permeate from the membrane was collected at different time fractions and permeate flux (J) was calculated with the following equation.
J = V/(A × t)
where, J = permeate flux during filtration (L m−2·h−1), V = volume of permeate (L), A = active membrane filtration area (m2) and t = filtration time (h) [66].
In the feed tank, after 50% reduction of volume (volume reduction factor 2), reduction of permeate flux (∆J) was calculated from initial permeate flux with Equation (2).
J (%) = (JinitialJfinal) × 100/Jinitial
where, ∆J = reduction of permeate flux (−), Jinitial = initial permeate flux (L m−2·h−1) and Jfinal = final permeate flux (L m−2·h−1) [65].
During filtration, pressures at inlet and retentate flow channels of the membrane module were recorded. Specific energy consumption (Es) was calculated with Equation (3).
Es = (RFR × ∆p)/(Jinitial × A)
where, Es = specific energy consumption (kWh·m−3), QR = retention flow rate (L·h−1), Δp = difference of pressure (Newton·m−2), A = active membrane filtration area (m2) and Jinitial = initial permeate flux (L m−2·h−1) [65].
After removing the milk serum, membrane cleaning was performed with 1% ultrasil and 1% citric acid in a sequential way with intermediate water cleaning. During cleaning with ultrasil and citric acid, TMP 0.8 bar and RFR 200 L h−1 were used. During cleaning with water, TMP 4 bar and RFR 200 L h−1 were used. Prior to removing the milk serum, membrane compaction was performed with de-ionized water to achieve the steady state water permeate flux. For that purpose, TMP 4 bar and RFR 200 L·h−1 were used [67].

2.3.2. Enzymatic Hydrolysis of Concentrated Proteins in Milk

Milk with concentrated proteins was collected from the storage tank of the membrane module. Prior to enzymatic reaction, milk with concentrated proteins, pH 7 was pre-incubated until the temperature reached 40 °C in a laboratory-scale well-controlled jacketed bioreactor, working volume 0.6 L and aspect ratio H/D* 2:1 (Solida Biotech, München, Germany). After pre-incubation of milk with concentrated proteins, it was treated with different concentrations of trypsin, such as 0.008 g·L−1, 0.016 g·L−1, 0.032 g·L−1, and 0.064 g·L−1. Individual batch-mode experiments were performed for protein hydrolysis process. For that purpose, 450 µL, 900 µL, 1.8 mL, and 3.6 mL of trypsin solution from stock solution (concentration of trypsin 0.009 g·mL−1) were injected through 0.22 μm of polytetrafluoroethylene (PTFE) syringe filter (VWR International, Pennsylvania, USA) to 500 mL of milk with concentrated proteins in bioreactor [67]. Enzymatic reaction was performed at a temperature of 40 °C for 10 min [68,69]. During enzymatic reaction, agitation speed in the bioreactor was maintained, 175 rpm, and the pH of milk in the bioreactor was controlled, 6.8, by automated addition of 2.0 N of sodium hydroxide or hydrochloric acid [67]. After 10 min of enzymatic reaction, 20 mL of sample was collected by a syringe from the bioreactor and kept in a sample tube. Activity of trypsin in enzymatic reaction was stopped by heat treatment. For that purpose, sample tubes were immediately placed in a water bath at temperature 70 °C for 30 min because denaturation temperatures of bovine α-lactalbumin and β-lactoglobulin are ~75 °C [70,71,72]. Two control samples (without enzyme treatment) were considered in the experiment: control 1: ultra-heat-treated skimmed milk was heated at temperature 40 °C for 10 min and subsequently placed at temperature 70 °C for 30 min; control 2: milk with concentrated protein was heated at temperature 40 °C for 10 min and subsequently placed at temperature 70 °C for 30 min. After inactivation of trypsin, the temperature of samples (reaction mixture) was reduced to ambient temperature (~25 °C) and freshly prepared samples were used for all kinds of biochemical assay, described in Section 2.4.

2.4. Analytical Method

2.4.1. Understanding of Molecular Weight of Proteins in Concentrated Milk

Molecular weight of proteins in concentrated milk was determined by liquid chromatography-electrospray ionization time-of-flight mass spectrometry (LC-ESI-TOF-MS) (Agilent Technologies, Santa Clara, CA, USA). Sample preparation was performed according to the protocol, mentioned by Rauh et al., 2015 [73]. Briefly, 200 µL of concentrated milk was treated with 20 µL of 0.5 M DTT, 1 mL of 100 mM trisodium citrate, and 6 M urea at temperature 30 °C for 1 h in a thermostat. Subsequently, a sample was centrifuged at 9500 g for 20 min at temperature 4 °C by a laboratory centrifuge (HERMLE Labortechnik, Wehingen, Germany). Aliquot of the clear phase was collected aseptically and used for LC-MS analysis. Chromatographic separation was achieved by an XBridge BEH300 C4 column with particle size: 3.5 µm, and inner diameter x length: 2.1 mm × 150 mm (Waters, Milford, USA), placed in an Agilent 1200 HPLC system (Agilent Technologies, Santa Clara, CA, USA). The column temperature was 30 °C during chromatographic separation. The binary mobile phase consisted of Milli-Q ultrapure deionized water with 0.1% formic acid (eluent A), and acetonitrile (eluent B) was used for that purpose. The flow rate was set to 0.5 mL·min−1. Gradient separation started at 3% B and linearly increased to reach 90% in 9 min. The eluent was kept constant at 90% B until 11 min and then the column was re-equilibrated at the initial conditions for 8 min. A UV signal was recorded at 280 nm using the diode-array detector (DAD) in the LC system and the effluent was connected to an Agilent 6530 high-resolution, accurate-mass, quadrupole time-of-flight mass spectrometry system equipped with a dual sprayer electrospray ion source. The mass spectrometry was run with full scan, MS-only mode (2 GHz, extended dynamic range setup) scanning in the range of 50–3200 m/z in positive ionization mode. A continuous reference mass correction was applied using purine and HP-921 (Hexakis(1H,1H,3H-perfluoropropoxy)phosphazene) as reference substances. The ion source temperature was maintained at 325 °C, and capillary and fragmentor voltages were set to −4000 V and 140 V, respectively. The Mass Hunter (MH) Workstation software package (version B02.01) and MH BioConfirm (version B 09.00) (Agilent Technologies, Santa Clara, CA, USA) were used for data acquisition and data evaluation, respectively. For raw mass spectrum deconvolution, the maximum entropy algorithm was used with automatic mass range detection (for intact protein), and for multiply charged ions, 500–3000 m/z limited range was considered.

2.4.2. Understanding of Hydrolysis of Liquid Milk Protein Concentrate

Molecular weight of proteins in ultra-heat-treated skimmed milk, milk with concentrated proteins, and enzyme-treated milks was determined by the sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) method. For this purpose, a vertical electrophoresis system (Bio-Rad Mini Protean Tetra system) and standard protein marker (precision plus protein standards) from Bio-Rad (Bio-Rad, Hercules, CA, USA) were used. In the SDS-PAGE method, concentration of stacking gel and running gel were 6% and 15%, respectively. The Laemmli sample buffer with 2-mercaptoethanol was used for dilution of samples and 10 µL of appropriate diluted sample was loaded into the respective wells. 0.2% Coomassie Brillant Blue R250 was used for gel staining. After 30 min of gel staining, de-staining of gel was performed with 50% (volume basis) of methanol-water and 10% (volume basis) of acetic acid. Gel image was captured using a Gel Doc XR+ System (Bio-Rad, Hercules, USA) and the molecular weight of bands were determined using Quantity One software program (version 4.6) (Bio-Rad, Hercules, CA, USA) [74].

2.4.3. Immunoblotting of Concentrated Milk Proteins

Proteins from SDS-PAGE gel were transferred onto a 0.45 μm of polyvinylidene difluoride (PVDF) membrane (Merck-Millipore, Molsheim, France) by a trans blot semi-dry transfer cell (Bio-Rad, Hercules, CA, USA). It was operated with 0.25 V and 0.08 mA/cm2 for 60 min. Immune-reactive proteins were identified with clinically proved milk positive pooled human serum and peroxidase-labelled anti-human Immunoglobulin E. The binding patterns were visualized using a substrate solution containing 4-chloronaphtol, hydrogen peroxide, and ethanol in phosphate buffered saline solution. Image analysis of blots was carried out with Gel Doc 2000 system (Bio-Rad, Hercules, CA, USA) [75].

2.4.4. Determination of Antioxidant Capacity

Antioxidant capacity of ultra-heat-treated skimmed milk, milk with concentrated proteins and enzyme-treated milks was measured using the Ferric reducing ability of plasma method with respect of ascorbic acid [76]. Appropriate diluted 100 µL of all kinds of milk samples with 2.9 mL of reagent (20 mM of ferric chloride: 10 mM of 2,4,6-Tris(2-pyridyl)-s-triazine with 40 mM of hydrochloric acid: 300 mM of acetate buffer, pH 3.6 = 1:1:10 (volume basis)) were incubated at temperature ~35 °C for 30 min in a water bath. Colorimetric determination was performed in room temperature (~25 °C) with a UV-Vis spectrophotometer (Thermo ScientificTM, Waltham MA, USA). Spectrophotometric measurement was performed with wavelength 593 nm.

2.4.5. Estimation of Angiotensin-Converting-Enzyme Inhibitory Activity

Enzymatic reaction mixture (final volume 200 µL in each well), consisted of 50 mM of sodium chloride, 100 mM of TRIS HCl (pH 7), 10 µM of zinc chloride, 15 µM of substrate Abz-FRK(Dnp)-P, recombinant angiotensin converting enzyme, and milk samples (in a dilution range of 10-fold to 106-fold) was used in investigation. The amount of the recombinant angiotensin converting enzyme was chosen to result in about 10-fold activity than that in the human serum (dilution was 200 to 400-fold from the stock). Reaction was initiated by the addition of substrate. Changes in fluorescent intensities were recorded in each 2–3 min and then changes were plotted as the function of time. These plots were fitted by a linear fit, and the slope was used to estimate enzyme activity (slope represents the change in fluorescent intensity in one minute). Activities in the absence of milk samples (uninhibited samples) were used as controls. The level of inhibition was calculated as % of uninhibited activity in each plate. Measurements were performed in a fluorescent plate reader (BMG Labtech, Ortenberg, Germany) at temperature 37 °C in Corning 96 wells black and flat bottom plates (Corning, New York, USA). Changes in optical density were measured with wavelength 340 nm for at least 90 min with 5 min intervals [77].

2.4.6. Determination of Protein Concentration

Concentration of protein in ultra-heat-treated skimmed milk, milk with concentrated proteins, and enzyme-treated milks were determined by the Bradford assay. Appropriate dilution of 100 µL of all kinds of milk samples with 3 mL of Bradford reagent were incubated at room temperature (~25 °C) for 30 min in a water bath. Colorimetric determination was performed with wavelength 280 nm in a UV-Vis spectrophotometer (Thermo ScientificTM, Waltham, MA, USA). Assay was performed in room temperature (~25 °C) and bovine serum albumin as a standard was used in assay [78].

2.4.7. Microbiological Assay

Antibacterial activity of ultra-heat-treated skimmed milk, milk with concentrated proteins, and enzyme-treated milks against Bacillus cereus and Staphylococcus aureus ATCC 6538 were investigated. Antibacterial activity was measured by agar well diffusion method. Sterile soybean casein digestive agar medium was used in the investigation. Freshly prepared (overnight grown culture) each culture was diluted with maximum recovery diluent (MRD) solution (8.5 g sodium chloride + 1 g peptone in 1 L of de-ionized water) to reach the bacterial concentration 106 colony-forming units mL−1 in respective agar plate [79]. Bacterial culture was spread on solidified agar in respective petri plates (pour plated) and agar wells with diameter 5 mm were filled with 100 μL of control milk and enzyme-treated milk samples. Petri plates were incubated at temperature 37 °C for 48 h in a biological incubator (HACH, Düsseldorf, Germany) [65,67]. The diameter of zone of inhibitions in microbial plates were measured by excluding the diameter of wells (5 mm) using a digital Vernier caliper (UEMATSU SHOKAI CO., LTD., Sendai, Japan).

2.5. Statistical Analysis

All experiments were performed at three times (technical triplicate). The mean value and standard deviation were calculated by Microsoft Excel (version 2013) (Microsoft Corporation, Washington, DC, USA). Subsequently, one-way analysis of variance method followed by the Tukey’s post hoc test were performed to understand the significant difference (P < 0.05) between different groups. SPSS 15.0 (version 25.0) (IBM, Armonk, NY, USA) was used for statistical analysis.

3. Results and Discussion

3.1. Concentrate Milk Proteins in Skimmed Milk by Membrane Filtration

A ceramic tubular membrane with active filtration area 5 × 10−3 m2 and pore size 5 nm was used to concentrate milk proteins in ultra-heat-treated skimmed milk by removing milk serum as a permeate. At room temperature and pH ~7, casein micelle may have a mean radius of 50 nm, whereas, the radius of whey proteins, such as α-lactalbumin, β-lactoglobulin, bovine serum albumin, and tetrad immunoglobulin are ~1.8 nm, ~1.8 nm, ~4 nm, and ~6 nm, respectively [80]. Typically, ultra-heat-treated milk is prepared with temperature 135–145 °C and treatment exposure time 1–8 s [81]. Due to heat treatment with high temperature, beside the Maillard reaction, sizes of proteins in milk are changed compared to their conventional sizes. When milk is heated at a temperature above 80 °C, the tertiary structure of whey protein turns to unfold [82]. It has been reported that at a temperature higher than 80 °C, denaturation rate of α-lactalbumin is faster than β-lactoglobulin’s and denaturation of α-lactalbumin is faster when β-lactoglobulin is present [83,84]. Subsequently, intramolecular highly reactive thiol groups, broken hydrophobic, and disulphide bonds may bind with covalent and hydrophobic bonds among themselves or with casein molecules, especially with κ-casein, present in periphery of casein micelle [85,86,87]. Furthermore, some whey proteins with sulfur containing thiol group (R-SH) can bind with other proteins by covalent bonds. Bovine serum albumin and β-lactoglobulin [88,89], and κ-casein and β-lactoglobulin [90,91,92] may bind together due to heat treatment. However, α-lactalbumin does not contain -SH group, it may conjugate with caseins in presence of β-lactoglobulin [93]. In addition, heat treatment may promote the formation of isopeptide bond between lysine and glutamine (N-ε-(γ-glutamyl)-lysine) or asparagine (N-ε-(β-aspartyl)-lysine) among different proteins, present in liquid milk protein concentrate [94,95,96]. Due to faster thermodenaturation of α-lactalbumin in presence of β-lactoglobulin, it may completely conjugate with casein micelle [80]. Therefore, it might expect that most of whey proteins have chance to conjugate with casein and the size of casein micelle might increase. On the other hand, due to intermolecular conjugation of whey proteins, the size of whey proteins might increase. Because of it, most of the proteins might reject by the nanofiltration membrane and residual (unbounded) whey proteins and lactose might permeate with milk serum through membrane pores during nanofiltration. As nanofiltration is a pressure-driven membrane separation process, a gel layer is developed on the membrane surface during separation process. A detailed investigation was performed to reduce the development of gel layer on the membrane surface by changing TMP and RFR. In Table 1, initial permeate flux and percentage change of permeate flux for different TMPs and RFRs are reported.
It is observed that permeate flux of serum was increased with the increase of TMP, because TMP provided driving force on the membrane surface. At higher TMP, the formation of gel layer on membrane surface was reduced and convective flux of serum increased due to the driving force on the membrane surface. For a similar reason, the percentage change of permeate flux decreased with the increase of TMP. As an outcome, concentration of protein in retentate side of the membrane, increased. As an example, after volume reduction 2, concentrations of protein in storage tank of the membrane module were 59.2 g·L−1 and 42 g·L−1, when filtration process was performed with TMP 3 bar, RFR 100 L·h−1 and 2 bar, RFR 100 L·h−1, respectively. At constant TMP, permeate flux was increased at higher RFR; however, results were not statistically significant. The tubular membrane had the lower surface area to volume ratio and, therefore, high feed flow rate promoted permeation. In cross-flow module, fluid on the membrane surface flowed with horizontal direction on the membrane surface with higher velocity and created more turbulence at higher RFR. Due to the sweeping action of fluid on the membrane surface, the deposition of solute molecules on the membrane surface reduced. Lower deposition of solute molecules on the membrane surface reduced the formation of concentration polarization and gel layer resistance, accompanied by the increase rate of permeation. Moreover, it was found that rate of flux declination was lower when the static turbulence promoter was used in the filtration process. The static turbulence promoter offered tangential velocity of fluid across the membrane surface, which created turbulence and vorticity of fluid on the membrane surface. Furthermore, the static turbulence promoter provided centrifugal force on the fluid, which contributed driving force on the membrane surface. All these factors reduced the deposition of solute molecules on the membrane surface and membrane gel layer resistance, which offered higher permeate flux in filtration process. As permeate flux was significantly higher in the static turbulence promoter-implemented filtration process, specific energy consumption was studied with different TMPs and RFRs in static turbulence promoter-implemented filtration process (Figure 3).
At constant TMP, values of specific energy consumption in filtration process were significantly low at lower RFR. When RFR was 200 L·h−1, permeate flux was not significantly increased compared to 100 L·h−1 because RFR could not provide driving force on the membrane surface. Therefore, permeate flux was not significantly increased compared to pressure drop at two opposite ends of the membrane. Filtration process with static turbulence promoter, higher TMP and lower RFR, tangential velocity of fluid across the membrane surface, driving, and centrifugal force on the fluid were generated. As an outcome, permeate flux was increased and pressure drop was reduced. Protein concentration in the retentate side of the membrane is also represented in Figure 3. It is noted that concentration of protein in the retentate side was higher with TMP 3 bar compared to TMP 2 bar. Protein concentration increased due to higher permeation of serum through membrane pores at higher TMP. Concentration of protein in retentate was not significantly increased at RFR 200 L·h−1 compared to 100 L·h−1, because RFR could not generate the driving force on the membrane surface and osmotic pressure. In Figure 4, time histories of the permeate flux, without and with the static turbulence promoter, are presented.
It was noted that when the static turbulence promoter was used in membrane filtration process, rate of flux declination and filtration time were reduced because the formation of gel layer was reduced in the presence of static turbulence promoter inside of tubular membrane.

3.2. Molecular Weight of Different Proteins in Concentrated Milk and Their Enzymatic Hydrolysis

Analysis of molecular weight of different proteins in concentrated milk was performed using UV chromatogram, total ion chromatogram (TIC), and deconvoluted mass spectrum of observed protein spectra (Figure 5).
Major proteins provided pronounced UV signal at 280 nm. In that UV wavelength, three peaks at 5.4 min, 5.6 min, and 5.8 min retention times were detected. Corresponding MS (TIC) signals are fully matched with UV peaks and deconvolution of each chromatographic peak spectrum was performed. Deconvoluted mass spectrum of proteins appeared in the retention time 5.4 min is provided in Figure 5(C1). In this figure, it is noted that there are two major deconvoluted masses, such as 19006 Da and 19038 Da. According to the previously published results, they might represent κ-casein. Deconvoluted mass spectrum of proteins appeared in retention time 5.6 min is provided in Figure 5(C2). In this figure, two major deconvoluted masses, such as 23617 Da and 23697 Da are observed. Comparing with the previously published results, they might represent α-casein [97]. Protein with molecular mass 23617 Da might be dephosphorylated form of α-casein (-80 Da mass shift from 23697 Da). Different types of caseins have a high degree of phosphorylation, which is generally affected by high temperature treatment during milk processing [98]. Interestingly, two protein with molecular mass shift +324 Da were observed in Figure 5(C2). These proteins with molecular mass 23941 Da and 24022 Da might be the lactosylated form of their original protein. Lactosylation of protein took place due to heat treatment during ultra-heat-treated milk production and, subsequently, their storage. The Amadori product ε-lactulosyllysine is produced by free ε-amino group of lysine in protein chain and milk sugar lactose [99]. It has been reported that protein become more hydrophilic due to addition of lactose in its structure, which results a shift to lower retention time [100,101]. Deconvoluted mass spectrum of proteins appeared in retention time 5.8 min is provided in Figure 5(C3). In this figure, two major deconvoluted masses, such as 18282 Da and 18368 Da, along with their lactosylated form with molecular mass 18606 Da and 18692 Da are observed. According to the already published results, the original protein might represent β-lactoglobulin [97]. According to the electrophoretic pattern, represented in Figure 6, ultra-heat-treated skimmed milk and milk with concentrated proteins may have had immunoglobulin, lactoferrin, lactoperoxidase, bovine serum albumin, α-casein, conjugated β-lactoglobulin, and α-lactalbumin or dimer β-lactoglobulin, β-casein, γ-casein, κ-casein, β-lactoglobulin, and α-lactalbumin with molecular weight ~150 kDa, ~80 kDa, ~78 kDa, ~66 kDa, ~35 kDa, ~34 kDa, ~25 kDa, ~22 kDa, ~20 kDa, ~18 kDa, and ~14 kDa, respectively. Some other investigators also published similar results [83,102,103,104].
In the PAGE image, hydrophobic protein conjugate with molecular weight ~34 kDa is clearly visualized. It was reported that due to heat treatment of skimmed milk, sometimes β-lactoglobulin and α-lactalbumin might participate in intermolecular thiol-disulphide bond interchange to produce covalently bonded hydrophobic aggregates [105]. Another group of investigators reported that at temperature more than 90 °C, β-lactoglobulin might present with disulphide-bonded dimer with molecular weight ~34 kDa and monomer [106]. However, some researchers reported about the formation of dimer α-lactalbumin with molecular weight ~28 kDa [107], but it was not found in our investigation. From the above discussion, it may say that the molecular weight of casein in concentrated milk, determined by SDS-PAGE and mass-spectroscopy is not directly comparable. It may explain by the fact that the electrophoretic mobility of caseins in electrophoresis gel is lower than expected from their molar mass [108]. It may be justified by the fact that phosphorylation [109] and lactosylation of caseins [73] change the migration of casein molecules in electrophoresis gel. However, in SDS-PAGE, several protein aggregates were present, they were absent in mass-spectrum. The possible reason is that dissociation of protein molecules and disruption of any type of protein aggregate might done by reducing agents DTT and chaotropic agent urea in sample preparation for mass-spectroscopy [110].
Without any contradiction, it was found that the numbers of peptide bands were increased due to tryptic digestion of milk proteins (lane 6–9). The hydrolysis of concentrated milk proteins was dose-dependent because it was noted that band numbers with lower molecular weight were increased gradually with increase of enzyme concentration in hydrolysis reaction. Immunoglobulin were hydrolyzed at more than 99% when concentration of trypsin was increased from 0.016 g·L−1 to 0.032 g·L−1. Lactoferrin, lactoperoxidase, and bovine serum albumin were hydrolyzed at more than 99% due to treatment with 0.008 g·L−1 of trypsin. Furthermore, κ-casein and β-casein were hydrolyzed at more than 99% when concentration of trypsin was increased from 0.008 g·L−1 to 0.016 g·L−1, whereas α-casein was retained. α-casein was hydrolyzed at more than 99% with 0.064 g·L−1 of trypsin. This can be justified by the fact that α-casein might has less chance to participate in enzymatic reaction because in the interior part of casein micelle, calcium phosphate clusters bind with the phosphoseryl residues of αs-casein and β-casein, whereas κ-casein was present in the periphery of casein micelle and received chance to participate in enzymatic reaction [111]. Due to partial hydrolysis of β-casein with 0.008 g·L−1 of trypsin, some peptone and γ-casein with molecular weight ~22 kDa might produce and they were hydrolyzed when milk with concentrated proteins was treated with 0.032 g·L−1 of trypsin. Dimer β-lactoglobulin with molecular weight ~32 kDa was hydrolyzed when trypsin was increased from 0.016 g·L−1 to 0.032 g·L−1.

3.3. Antioxidant Capacity

Antioxidant capacity of milk with concentrated proteins was 167.35 ± 9.8 mg equivalent ascorbic acid L−1 and it was increased after enzyme treatment. In Figure 7, it is noted that change of antioxidant capacity in enzyme-treated milks was dose-dependent. Similar types of findings were also published by other researchers [62,63].
Trypsin is a serine endopeptidase, consisting of three amino acids, such as His 57, Ser 195, and Asp 102 in catalytic triad. In amino acid sequence, trypsin cleaves between the carboxyl group of basic amino acid lysine or arginine in N terminal position and the amino group of the adjacent amino acid with hydrophobic side chain in C terminal position. This cleavage does not occur when lysine or arginine is followed by proline. Adjacent hydrophobic amino acid, such as alanine, isoleucine, leucine, methionine, phenylalanine, valine, proline, and glycine in peptide chain, derived from milk proteins by tryptic hydrolysis, offered reducing activity towards ferric ion [64,112,113].

3.4. Angiotensin Converting Enzyme-Inhibitory Activity

Angiotensin converting enzyme inhibitory activity of ultra-heat-treated skimmed milk and milk with concentrated proteins was negligible. Inhibitions of angiotensin converting enzyme were ~15% and ~6% for ultra-heat-treated skimmed milk and milk with concentrated proteins, respectively (Figure 8A). Our finding was similar in accordance with other researchers [114,115].
It can be justified by the fact that interaction between active side of angiotensin converting enzyme and native milk proteins might not be facilitated, because steric hindrance might present [116]. Angiotensin converting enzyme in concentrated milk proteins significantly increased after trypsin treatment. Similar types of findings were reported by other investigators [117,118]. Changes of IC50 value in liquid milk protein concentrate due to enzyme treatment were dose-dependent (Figure 8B). Because of tryptic hydrolysis of milk proteins, active sides in low molecular weight peptides were exposed and interaction with angiotensin converting enzyme was facilitated. It was reported that peptides with hydrophobic amino acids, such as proline, tryptophan, tyrosine, and phenylalanine at C-terminal position, are able to bind with angiotensin converting enzyme [119,120]. In our investigation, more than 95% inhibition was not achieved. This can be justified by the fact that angiotensin converting enzyme inhibitory peptides, produced by tryptic hydrolysis of milk proteins might change the structural configuration of angiotensin converting enzyme, which might not favorable for interaction between substrate and angiotensin converting enzyme [121].

3.5. Antibacterial Activity

Antibacterial activity (represented in zone of inhibition) of enzyme-treated liquid milk protein concentrate towards Bacillus cereus and Staphylococcus aureus was proven. No zone of inhibition was found when milk with concentrated proteins was tested. In Figure 9, radius of zone of inhibition represented the antibacterial activity of milk with concentrated proteins after enzyme treatment is mentioned.
Zone of inhibition (radius of inhibition zone) was significantly increased when concentration of trypsin was increased from 0.008 g·L−1 to 0.032 g·L−1 during hydrolysis of liquid milk protein concentrate. It was found that for Bacillus cereus, values of zone of inhibition (radius of inhibition zone) were 2.5 ± 0.5 mm and 5.2 ± 0.3 mm, when concentrated milk protein was treated with 0.008 g·L−1 and 0.064 g·L−1 of trypsin, respectively. For Staphylococcus aureus, values of zone of inhibition (radius of inhibition zone) were 2.3 ± 0.03 mm and 4.5 ± 0.5 mm when concentrated milk protein was treated with 0.008 g·L−1 and 0.064 g·L−1 of trypsin, respectively. Several biochemical mechanisms about antibacterial activity of enzyme-treated liquid milk protein concentrate were reported. Trypsin cleaves the peptide bond at the C-terminus of lysin and arginine, when the N terminus is not a proline. Several peptides with hydrophobic, hydrophilic or amphipathic amino acids, produced due to tryptic hydrolysis of milk proteins. These peptides may interact with peptidoglycan in bacterial cell membrane, create a complex with bacterial cell wall components, and, subsequently, create pores in bacterial cell membrane. These pores might expedite the permeabilization of cellular contents to the abiotic environment and subsequently, destruction of cell. It has been also reported that interaction between bacterial cell membrane with antibacterial peptides frequently leads to lipid segregation in the cell membrane. It leads to delocalization of essential membrane proteins, increase membrane permeability, inhibit cell division, followed by cellular death [122,123].

3.6. Allergenicity

Immunoblotting, a combination of gel electrophoresis and antigen-antibody reaction was performed with positive pooled human sera to understand the allergenic potentiality of proteins, present in liquid milk protein concentrate. It was reported that major cow milk allergens belong to the casein fraction (αS1-, αS2-, β-, and κ-casein), and whey proteins α-lactalbumin and β-lactoglobulin; however, lactoferrin, bovine serum albumin and immunoglobulins, which are present with lower quantities in cow milk, have importance in allergenic reaction [22]. In present investigation, it was noted that however, immunoglobulin, lactoferrin, lactoperoxidase, bovine serum albumin, and casein had strong interaction with antibody, monomeric β-lactoglobulin had weak interaction, and monomeric α-lactalbumin had no detectable interaction (Figure 10).
This result can be explained by the following justifications. As the experiment was performed with ultra-heat-treated skimmed milk, due to heat treatment most of α-lactalbumin and β-lactoglobulin were unfolded, and allergenic epitopes in α-lactalbumin and β-lactoglobulin were destroyed [104,124]. Allergenic epitopes in higher molecular weight proteins, such as bovine serum albumin, lactoperoxidase, conjugated α-lactalbumin, and β-lactoglobulin or α-casein were not fully affected during heat treatment because their denaturation temperatures were quite high [104,125]. In an investigation with 20 children (median age 4 months), it was also found that, however, αS1-casein, αS2-casein, β-casein, κ-casein, bovine serum albumin, Immunoglobulin-G heavy chain, and lactoferrin had allergenic cross-linking, α-lactalbumin did not have any allergenic activity [126]. However, all proteins in concentrated milk, except α-lactalbumin, were immunoreactive; they lost allergenic activity due to enzymatic hydrolysis. Residual immunoreactivity of caseins and dimeric β-lactoglobulin or conjugated α-lactalbumin-β-lactoglobulin were still present at 0.016 g·L−1 of trypsin treatment. They lost allergenic potentiality when 0.032 g·L−1 of trypsin was used in enzymatic reaction.

3.7. Superiority of the Process

After skillful experiment to prepare liquid skimmed milk protein concentrate, the superior operation strategy was trans-membrane pressure 3 bar, retention flow rate 100 L·h−1, and implementation of a static turbulence promoter within a tubular ceramic membrane with pore size 5 nm and filtration area 5 × 10−3 m2. In the present investigation, a cross-flow membrane module was adopted and batch-mode filtration was performed with volume reduction factor 2. To prepare milk protein concentrate, polymeric spiral wound [19,127,128], single flat sheet [13,129,130], and tubular [131] membranes were used. Furthermore, ceramic tubular membrane was used to prepare milk protein concentrate by several investigators [132,133,134,135,136]. Single stage membrane filtration [19,127,129,131,136] and ultrafiltration with diafiltration were used by several investigators to prepare milk protein concentrate [137,138]. Both continuous and discontinuous diafiltration were adopted to reduce major whey proteins, such as α-lactalbumin and β-lactoglobulin from casein fraction by ultrafiltration or nanofiltration membrane. In continuous diafiltration process, sterilize water was added to the feed tank to maintain the feed volume, whereas discontinuous diafiltration was introduced in the process when the feed concentration reached to a certain level to overcome the high viscosity of feed and protect the membrane by fouling. However, milk protein concentrate was prepared by both polymeric and ceramic membrane; flux declination is a considerable drawback. To overcome this issue, stepwise increase of TMP was adopted in sometimes [131,134] and it may feel that in respect of energy consumption, this approach is not appreciable. In our investigation, due to application of the static turbulence promoter reduction of permeate flux was remarkably low. The experiment was started with 1 L of milk in the feed tank and after volume reduction factor 2 without diafiltration, initial permeate flux 34 L·m−2·h−1 reduced to 26 L·m−2·h−1, i.e., only 24% reduction in permeate flux. During filtration, consumption of mechanical energy, contributed by fluid flow rate and TMP was 4.9 kWh m−3. However, in SDS-PAGE image all proteins in concentrated milk were clearly visualized, but in immunoblot it was found that monomeric α-lactalbumin and β-lactoglobulin did not offer allergenicity, due to change of their structural configuration during heat treatment. Hence, diafiltration was not required to fulfil the objective of present investigation. Filtration process without diafiltration might be reduced the water consumption and process time. According to our experimental finding, application of 0.032 g·L−1 of trypsin to liquid milk protein concentrate at temperature 40 °C for 10 min can delete allergenic epitopes and increase the antioxidant capacity, anti-angiotensin enzyme activity, and antibacterial activity compared to native milk protein concentrate. According to the literature review, it may be considered that it is the first approach in the context of development of dairy-based hypoallergenic functional food by membrane- and enzyme- based technologies.

4. Conclusions

In the present investigation, liquid milk protein concentrate with antioxidant capacity, anti-angiotensin activity, antibacterial activity, and allergen-free was produced from ultra-heat-treated skimmed milk by combination of membrane- and enzyme- based technologies. A ceramic-made tubular nanofiltration membrane with pore size 5 nm, placed in a cross-flow membrane house, was used for the production of liquid milk protein concentrate by reducing milk serum as a permeate. As the mean radius of casein micelle and their conjugated form with whey proteins in ultra-heat-treated skimmed milk were quite high compared to the pore size of the nanofiltration membrane, they were almost rejected by the membrane. Membrane filtration process alone cannot change the biological activities of milk proteins. Biological activities of protein derivatives, i.e., peptides offer antioxidant capacity, angiotensin converting enzyme inhibitory activity, and antibacterial activity. Therefore, milk with concentrated proteins from the retentate side of the membrane was treated with the different concentrations of trypsin, such as 0.008 g·L−1, 0.016 g·L−1, 0.032 g·L−1, and 0.064 g·L−1 in individual batch-mode experiments. Hydrolysis of milk protein was enzyme dose dependent because trypsin cleaves the peptide bond at the C-terminus of lysin and arginine in specific way. Antioxidant capacity, angiotensin converting enzyme inhibitory activity, and antibacterial activity of liquid milk protein concentrate were increased depending on the enzymatic hydrolysis of milk proteins. Trypsin concentration 0.032 g·L−1 was able to reduce allergenic epitopes at more than 99.9% in liquid milk protein concentrate.
After summarizing all experimental results, one may believe that the proposed technology may reduce the limitations of milk protein concentrate production and increase the consumption of dairy products. To the best of our knowledge, this is the first attempt to produce liquid milk protein concentrate with antioxidant capacity, angiotensin converting enzyme inhibitory activity, antibacterial activity, and hypoallergenic property by membrane filtration and enzymatic modification of proteins. In general, small-scale and medium-scale dairy plants prefer to use liquid milk protein concentrate for preparing cultured-dairy products instead of using dried milk protein concentrate for economical issue. Therefore, to implement this process in industrial scale, further systematic investigation with techno-economical viewpoint is a prerequisite. However, in the present investigation molecular weight of peptides derived from concentrated milk proteins were determined by SDS-PAGE; more accurate results about molecular mass distribution of peptides and their sequences may be determined by a LC-ESI-Q-TOF-MS-based bottom-up sequencing in future research. In the groundbreaking research area of food biotechnology, one may expect that the proposed research may receive attention from academic and industrial sectors.

Author Contributions

A.N. was involved in performing the experiment, evaluating the results, and writing the whole manuscript. B.A.E. and A.C. were involved in assisting with the experiments, preparing graphs, and tables. G.K. was involved in the microbiological assay. L.A., A.T., E.S., and A.K. were involved in performing the biochemical assay. Z.K. was involved in statistical analysis. K.P.-H. and G.V. were involved in cross checking results and correcting the whole manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Authors acknowledge the support from the European Union project (grant agreement no. EFOP-3.6.3-VEKOP-16-2017-00005). A. Csighy acknowledges Doctoral School of Food Science, Szent István University, Hungary. Z. Kovacs acknowledges the support of the New National Excellence Program of the Ministry for Innovation and Technology (ÚNKP-19-4-SZIE-27) and the Bolyai János Scholarship from the Hungarian Academy of Sciences. L. Abrankó is grateful for the János Bolyai Research Scholarship from the Hungarian Academy of Sciences. This research was supported by the Higher Education Institutional Excellence Program (FEKUTSTRAT) awarded by the Ministry of Human Capacities within the framework of plant breeding and plant protection researches of Szent István University, Budapest, Hungary.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. St-Onge, M.-P.; Farnworth, E.R.; Jones, P.J. Consumption of fermented and nonfermented dairy products: Effects on cholesterol concentrations and metabolism. Am. J. Clin. Nutr. 2000, 71, 674–681. [Google Scholar] [CrossRef] [PubMed]
  2. Guinee, T.P.; O’Kennedy, B.T.; Kelly, P.M. Effect of Milk Protein Standardization Using Different Methods on the Composition and Yields of Cheddar Cheese. J. Dairy Sci. 2006, 89, 468–482. [Google Scholar] [CrossRef]
  3. Amatayakul, T.; Sherkat, F.; Shah, N.P. Syneresis in set yogurt as affected by EPS starter cultures and levels of solids. Int. J. Dairy Technol. 2006, 59, 216–221. [Google Scholar] [CrossRef]
  4. Schuck, P.; Jeantet, R.; Bhandari, B.; Chen, X.D.; Perrone, Í.T.; De Carvalho, A.F.; Fenelon, M.; Kelly, P. Recent advances in spray drying relevant to the dairy industry: A comprehensive critical review. Dry. Technol. 2016, 34, 1773–1790. [Google Scholar] [CrossRef]
  5. Ur Rehman, S.; Farkye, N.Y.; Considine, T.; Schaffner, A.; Drake, M.A. Effects of Standardization of Whole Milk with Dry Milk Protein Concentrate on the Yield and Ripening of Reduced-Fat Cheddar Cheese. J. Dairy Sci. 2003, 86, 1608–1615. [Google Scholar] [CrossRef]
  6. Meena, G.S.; Singh, A.K.; Panjagari, N.R.; Arora, S. Milk protein concentrates: Opportunities and challenges. J. Food Sci. Technol. 2017, 54, 3010–3024. [Google Scholar] [CrossRef]
  7. Amelia, I.; Barbano, D.M. Production of an 18% protein liquid micellar casein concentrate with a long refrigerated shelf life. J. Dairy Sci. 2013, 96, 3340–3349. [Google Scholar] [CrossRef] [Green Version]
  8. Henriques, M.H.F.; Gomes, D.M.G.S.; Pereira, C.J.D.; Gil, M.H.M. Effects of Liquid Whey Protein Concentrate on Functional and Sensorial Properties of Set Yogurts and Fresh Cheese. Food Bioprocess Technol. 2013, 6, 952–963. [Google Scholar] [CrossRef] [Green Version]
  9. Rosenberg, M. Current and future applications for membrane processes in the dairy industry. Trends Food Sci. Technol. 1995, 6, 12–19. [Google Scholar] [CrossRef]
  10. Pouliot, Y. Membrane processes in dairy technology—From a simple idea to worldwide panacea. Int. Dairy J. 2008, 18, 735–740. [Google Scholar] [CrossRef]
  11. Kumar, P.; Sharma, N.; Ranjan, R.; Kumar, S.; Bhat, Z.F.; Jeong, D.K. Perspective of Membrane Technology in Dairy Industry: A Review. Asian Australas. J. Anim. Sci. 2013, 26, 1347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Li, H.; Hsu, Y.-C.; Zhang, Z.; Dharsana, N.; Ye, Y.; Chen, V. The influence of milk components on the performance of ultrafiltration/diafiltration of concentrated skim milk. Sep. Sci. Technol. 2017, 52, 381–391. [Google Scholar] [CrossRef]
  13. Arunkumar, A.; MR, E. Milk Protein Concentration Using Negatively Charged Ultrafiltration Membranes. Foods 2018, 7, 134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Gavazzi-April, C.; Benoit, S.; Doyen, A.; Britten, M.; Pouliot, Y. Preparation of milk protein concentrates by ultrafiltration and continuous diafiltration: Effect of process design on overall efficiency. J. Dairy Sci. 2018, 101, 9670–9679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Ng, K.S.Y.; Haribabu, M.; Harvie, D.J.E.; Dunstan, D.E.; Martin, G.J.O. Mechanisms of flux decline in skim milk ultrafiltration: A review. J. Membr. Sci. 2017, 523, 144–162. [Google Scholar] [CrossRef]
  16. Bacchin, P.; Aimar, P.; Field, R.W. Critical and sustainable fluxes: Theory, experiments and applications. J. Membr. Sci. 2006, 281, 42–69. [Google Scholar] [CrossRef] [Green Version]
  17. Marshall, A.D.; Munro, P.A.; Trägårdh, G. The effect of protein fouling in microfiltration and ultrafiltration on permeate flux, protein retention and selectivity: A literature review. Desalination 1993, 91, 65–108. [Google Scholar] [CrossRef]
  18. Bahnasawy, A.H.; Shenana, M.E. Flux behavior and energy consumption of ultrafiltration (UF) process of milk. Aust. J. Agric. Eng. 2010, 1, 54–65. [Google Scholar]
  19. Méthot-Hains, S.; Benoit, S.; Bouchard, C.; Doyen, A.; Bazinet, L.; Pouliot, Y. Effect of transmembrane pressure control on energy efficiency during skim milk concentration by ultrafiltration at 10 and 50 °C. J. Dairy Sci. 2016, 99, 8655–8664. [Google Scholar] [CrossRef] [Green Version]
  20. Guetouache, M.; Guessas, B.; Medjekal, S. Composition and nutritional value of raw milk 2014. Bio. Sci. Pharm. Res. 2014, 2, 115–122. [Google Scholar]
  21. Walsh, J.; Meyer, R.; Shah, N.; Quekett, J.; Fox, A.T. Differentiating milk allergy (IgE and non-IgE mediated) from lactose intolerance: Understanding the underlying mechanisms and presentations. Br. J. Gen. Pract. 2016, 66, e609–e611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Lifschitz, C.; Szajewska, H. Cow’s milk allergy: Evidence-based diagnosis and management for the practitioner. Eur. J. Pediatrics 2015, 174, 141–150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Sicherer, S.H.; Sampson, H.A. Food allergy: Epidemiology, pathogenesis, diagnosis, and treatment. J. Allergy Clin. Immunol. 2014, 133, 291–307. [Google Scholar] [CrossRef]
  24. US Food and Drug Administration. Food Allergen Labeling and Consumer Protection Act of 2004 (FALCPA). J. Acad. Nutr. Diet. 2004, 106, 1742–1744. [Google Scholar]
  25. Sackesen, C.; Altintas, D.U.; Bingol, A.; Bingol, G.; Buyuktiryaki, B.; Demir, E.; Kansu, A.; Kuloglu, Z.; Tamay, Z.; Sekerel, B.E.; et al. Current Trends in Tolerance Induction in Cow’s Milk Allergy: From Passive to Proactive Strategies. Front. Pediatrics 2019, 7, 372. [Google Scholar] [CrossRef]
  26. Forouhi, N.G.; Krauss, R.M.; Taubes, G.; Willett, W. Dietary fat and cardiometabolic health: Evidence, controversies, and consensus for guidance. BMJ 2018, 361. [Google Scholar] [CrossRef] [Green Version]
  27. Lordan, R.; Tsoupras, A.; Mitra, B.; Zabetakis, I. Dairy Fats and Cardiovascular Disease: Do We Really Need to be Concerned? Foods 2018, 7, 29. [Google Scholar] [CrossRef] [Green Version]
  28. Ferreira, L.S.S.; Fernandes, C.S.; Vieira, M.N.N.; De Felice, F.G. Insulin Resistance in Alzheimer’s Disease. Front. Neurosci. 2018, 12. [Google Scholar] [CrossRef] [Green Version]
  29. Lee, H.J.; Seo, H.I.; Cha, H.Y.; Yang, Y.J.; Kwon, S.H.; Yang, S.J. Diabetes and Alzheimer’s Disease: Mechanisms and Nutritional Aspects. Clin. Nutr. Res. 2018, 7, 229–240. [Google Scholar] [CrossRef] [Green Version]
  30. Xu, Q.; Shi, J.; Yao, M.; Jiang, M.; Luo, Y. Effects of heat treatment on the antigenicity of four milk proteins in milk protein concentrates. Food Agric. Immunol. 2016, 27, 401–413. [Google Scholar] [CrossRef]
  31. Villa, C.; Costa, J.; Mafra, I. Detection and Quantification of Milk Ingredients as Hidden Allergens in Meat Products by a Novel Specific Real-Time PCR Method. Biomolecules 2019, 9, 804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Villa, C.; Costa, J.; Oliveira, M.B.P.P.; Mafra, I. Bovine Milk Allergens: A Comprehensive Review. Compr. Rev. Food Sci. Food Saf. 2018, 17, 137–164. [Google Scholar] [CrossRef] [Green Version]
  33. Bu, G.; Luo, Y.; Chen, F.; Liu, K.; Zhu, T. Milk processing as a tool to reduce cow’s milk allergenicity: A mini-review. Dairy Sci. Technol. 2013, 93, 211–223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Kleber, N.; Maier, S.; Hinrichs, J. Antigenic response of bovine β-lactoglobulin influenced by ultra-high pressure treatment and temperature. Innov. Food Sci. Emerg. Technol. 2007, 8, 39–45. [Google Scholar] [CrossRef]
  35. Chicón, R.; Belloque, J.; Alonso, E.; López-Fandiño, R. Immunoreactivity and digestibility of high-pressure-treated whey proteins. Int. Dairy J. 2008, 18, 367–376. [Google Scholar] [CrossRef]
  36. Kleber, N.; Hinrichs, J. Antigenic response of β-lactoglobulin in thermally treated bovine skim milk and sweet whey. Milchwissenschaft 2007, 62, 121–124. [Google Scholar]
  37. Fiocchi, A.; Restani, P.; Riva, E.; Mirri, G.P.; Santini, I.; Bernardo, L.; Galli, C.L. Heat treatment modifies the allergenicity of beef and bovine serum albumin. Allergy 2008, 53, 798–802. [Google Scholar] [CrossRef]
  38. Bu, G.; Luo, Y.; Zheng, Z.; Zheng, H. Effect of heat treatment on the antigenicity of bovine α-lactalbumin and β-lactoglobulin in whey protein isolate. Food Agric. Immunol. 2009, 20, 195–206. [Google Scholar] [CrossRef]
  39. Ismahan, B.D.; Hanane, K.; Omar, K.; Djamel, S. Microwave irradiation of bovine milk reduces allergic response in mouse model of food allergy. Front. Immunol. 2014, 10. [Google Scholar] [CrossRef]
  40. Cheison, S.C.; Wang, Z.; Xu, S.-Y. Preparation of Whey Protein Hydrolysates Using a Single- and Two-Stage Enzymatic Membrane Reactor and Their Immunological and Antioxidant Properties: Characterization by Multivariate Data Analysis. J. Agric. Food Chem. 2007, 55, 3896–3904. [Google Scholar] [CrossRef]
  41. Guadix, A.; Camacho, F.; Guadix, E.M. Production of whey protein hydrolysates with reduced allergenicity in a stable membrane reactor. J. Food Eng. 2006, 72, 398–405. [Google Scholar] [CrossRef]
  42. Kobayashi, K.; Hirano, A.; Ohta, A.; Yoshida, T.; Takahashi, K.; Hattori, M. Reduced Immunogenicity of β-Lactoglobulin by Conjugation with Carboxymethyl Dextran Differing in Molecular Weight. J. Agric. Food Chem. 2001, 49, 823–831. [Google Scholar] [CrossRef]
  43. Bu, G.; Lu, J.; Zheng, Z.; Luo, Y. Influence of Maillard reaction conditions on the antigenicity of bovine α-lactalbumin using response surface methodology. J. Sci. Food Agric. 2009, 89, 2428–2434. [Google Scholar] [CrossRef]
  44. Bu, G.; Luo, Y.; Lu, J.; Zhang, Y. Reduced antigenicity of β-lactoglobulin by conjugation with glucose through controlled Maillard reaction conditions. Food Agric. Immunol. 2010, 21, 143–156. [Google Scholar] [CrossRef]
  45. Pahud, J.J.; Monti, J.C.; Jost, R. Allergenicity of whey protein: Its modification by tryptic in vitro hydrolysis of the protein. J. Pediatric Gastroenterol. Nutr. 1985, 4, 408–413. [Google Scholar] [CrossRef]
  46. Nakamura, T.; Sado, H.; Syukunobe, Y.; Hirota, T. Antigenicity of whey protein hydrolysates prepared by combination of two proteases. Milchwissenschaft 1993, 48, 667–670. [Google Scholar]
  47. Wroblewska, B.; Karamac, M.; Amarowicz, R.; Szymkiewicz, A.; Troszynska, A.; Kubicka, E. Immunoreactive properties of peptide fractions of cow whey milk proteins after enzymatic hydrolysis. Int. J. Food Sci. Technol. 2004, 39, 839–850. [Google Scholar] [CrossRef]
  48. Ena, J.M.; Beresteijn, E.C.H.; Robben, A.J.P.M.; Schmidt, D.G. Whey Protein Antigenicity Reduction by Fungal Proteinases and a Pepsin/Pancreatin Combination. J. Food Sci. 1995, 60, 104–110. [Google Scholar] [CrossRef]
  49. Chicón, R.; Belloque, J.; Recio, I.; López-Fandiño, R. Influence of high hydrostatic pressure on the proteolysis of β-lactoglobulin A by trypsin. J. Dairy Res. 2006, 73, 121–128. [Google Scholar] [CrossRef]
  50. Bonomi, F.; Fiocchi, A.; Frøkiær, H.; Gaiaschi, A.; Iametti, S.; Poiesi, C.; Rasmussen, P.; Restani, P.; Rovere, P. Reduction of immunoreactivity of bovine β-lactoglobulin upon combined physical and proteolytic treatment. J. Dairy Res. 2003, 70, 51–59. [Google Scholar] [CrossRef]
  51. Beran, M.; Klubal, R.; Molik, P.; Strohalm, J.; Urban, M.; Klaudyova, A.A.; Prajzlerova, K. Influence of high-hydrostatic pressure on tryptic and chymotryptic hydrolysis of milk proteins. High Press. Res. 2009, 29, 23–27. [Google Scholar] [CrossRef]
  52. El Mecherfi, K.E.; Saidi, D.; Kheroua, O.; Boudraa, G.; Touhami, M.; Rouaud, O.; Curet, S.; Choiset, Y.; Rabesona, H.; Chobert, J.-M.; et al. Combined microwave and enzymatic treatments for β-lactoglobulin and bovine whey proteins and their effect on the IgE immunoreactivity. Eur. Food Res. Technol. 2011, 233, 859–867. [Google Scholar] [CrossRef]
  53. Izquierdo, F.J.; Alli, I.; Yaylayan, V.; Gomez, R. Microwave-assisted digestion of β-lactoglobulin by pronase, α-chymotrypsin and pepsin. Int. Dairy J. 2007, 17, 465–470. [Google Scholar] [CrossRef] [Green Version]
  54. Izquierdo, F.J.; Peñas, E.; Baeza, M.L.; Gomez, R. Effects of combined microwave and enzymatic treatments on the hydrolysis and immunoreactivity of dairy whey proteins. Int. Dairy J. 2008, 18, 918–922. [Google Scholar] [CrossRef] [Green Version]
  55. Yang, W.; Tu, Z.; Wang, H.; Zhang, L.; Kaltashov, I.A.; Zhao, Y.; Niu, C.; Yao, H.; Ye, W. The mechanism of reduced IgG/IgE-binding of β-lactoglobulin by pulsed electric field pretreatment combined with glycation revealed by ECD/FTICR-MS. Food Funct. 2018, 9, 417–425. [Google Scholar] [CrossRef]
  56. Quintieri, L.; Monaci, L.; Baruzzi, F.; Giuffrida, M.G.; De Candia, S.; Caputo, L. Reduction of whey protein concentrate antigenicity by using a combined enzymatic digestion and ultrafiltration approach. J. Food Sci. Technol. 2017, 54, 1910–1916. [Google Scholar] [CrossRef]
  57. El Mecherfi, K.E.; Curet, S.; Lupi, R.; Larré, C.; Rouaud, O.; Choiset, Y.; Rabesona, H.; Haertlé, T. Combined microwave processing and enzymatic proteolysis of bovine whey proteins: The impact on bovine β-lactoglobulin allergenicity. J. Food Sci. Technol. 2019, 56, 177–186. [Google Scholar] [CrossRef]
  58. Spies, J.B. Milk Allergy. Milk Food Technol. 1973, 36, 225–231. [Google Scholar] [CrossRef]
  59. El-Agamy, E.I. The challenge of cow milk protein allergy. Small Rumin. Res. 2007, 68, 64–72. [Google Scholar] [CrossRef]
  60. Kim, S.B.; Ki, K.S.; Khan, M.A.; Lee, W.S.; Lee, H.J.; Ahn, B.S.; Kim, H.S. Peptic and Tryptic Hydrolysis of Native and Heated Whey Protein to Reduce Its Antigenicity. J. Dairy Sci. 2007, 90, 4043–4050. [Google Scholar] [CrossRef] [Green Version]
  61. Ahmad, N.; Imran, M.; Khan, M.K.; Nisa, M.U. Degree of hydrolysis and antigenicity of buffalo alpha S1 casein and its hydrolysates in children with cow milk allergy. Food Agric. Immunol. 2016, 27, 87–98. [Google Scholar] [CrossRef]
  62. Abd El-Fattah, A.M.; Sakr, S.S.; El-Dieb, S.M.; Elkashef, H.A.S. Bioactive peptides with ACE-I and antioxidant activity produced from milk proteolysis. Int. J. Food Prop. 2017, 20, 3033–3042. [Google Scholar] [CrossRef] [Green Version]
  63. Luo, Y.; Pan, K.; Zhong, Q. Physical, chemical and biochemical properties of casein hydrolyzed by three proteases: Partial characterizations. Food Chem. 2014, 155, 146–154. [Google Scholar] [CrossRef] [PubMed]
  64. Bamdad, F.; Shin, S.H.; Suh, J.-W.; Nimalaratne, C.; Sunwoo, H. Anti-Inflammatory and Antioxidant Properties of Casein Hydrolysate Produced Using High Hydrostatic Pressure Combined with Proteolytic Enzymes. Molecules 2017, 22, 609. [Google Scholar] [CrossRef] [PubMed]
  65. Nath, A.; Szécsi, G.; Csehi, B.; Mednyánszky, Z.; Kiskó, G.; Bányai, É.; Dernovics, M.; Koris, A. Production of Hypoallergenic Antibacterial Peptides from Defatted Soybean Meal by a Membrane Associated Bioreactor: A Bioprocess Engineering Study with Comprehensive Product Characterization. Food Technol. Biotechnol. 2017, 55, 308–324. [Google Scholar] [CrossRef] [PubMed]
  66. Nath, A.; Chakraborty, S.; Bhattacharjee, C.; Chowdhury, R. Studies on the separation of proteins and lactose from casein whey by cross-flow ultrafiltration. Desalin. Water Treat. 2015, 54, 481–501. [Google Scholar] [CrossRef]
  67. Nath, A.; Kailo, G.G.; Mednyánszky, Z.; Kiskó, G.; Csehi, B.; Pásztorné-Huszár, K.; Gerencsér-Berta, R.; Galambos, I.; Pozsgai, E.; Bánvölgyi, S.; et al. Antioxidant and Antibacterial Peptides from Soybean Milk through Enzymatic- and Membrane-Based Technologies. Bioengineering 2019, 7, 5. [Google Scholar] [CrossRef] [Green Version]
  68. Nongonierma, A.B.; Paolella, S.; Mudgil, P.; Maqsood, S.; FitzGerald, R.J. Dipeptidyl peptidase IV (DPP-IV) inhibitory properties of camel milk protein hydrolysates generated with trypsin. J. Funct. Foods 2017, 34, 49–58. [Google Scholar] [CrossRef] [Green Version]
  69. Cheison, S.C.; Schmitt, M.; Leeb, E.; Letzel, T.; Kulozik, U. Influence of temperature and degree of hydrolysis on the peptide composition of trypsin hydrolysates of β-lactoglobulin: Analysis by LC–ESI-TOF/MS. Food Chem. 2010, 121, 457–467. [Google Scholar] [CrossRef]
  70. Elfagm, A.A.; Wheelock, J.V. Interaction of Bovine α-Lactalbumin and β-Lactoglobulin during Heating. J. Dairy Sci. 1978, 61, 28–32. [Google Scholar] [CrossRef]
  71. De Wit, J.; Klarenbeek, B. Effects of Various Heat Treatments on Structure and Solubility of Whey Proteins. J. Dairy Sci. 1984, 67, 2701–2710. [Google Scholar] [CrossRef]
  72. Cassiani, D.M.; Yamul, D.K.; Conforti, P.A.; Pérez, V.A.; Lupano, C.E. Structure and Functionality of Whey Protein Concentrate-Based Products with Different Water Contents. Food Bioprocess Technol. 2013, 6, 217–227. [Google Scholar] [CrossRef]
  73. Rauh, V.M.; Johansen, L.B.; Bakman, M.; Ipsen, R.; Paulsson, M.; Larsen, L.B.; Hammershøj, M. Protein lactosylation in UHT milk during storage measured by Liquid Chromatography-Mass Spectrometry and quantification of furosine. Int. J. Dairy Technol. 2015, 68, 486–494. [Google Scholar] [CrossRef]
  74. Csehi, B.; Szerdahelyi, E.; Pásztor-Huszár, K.; Salamon, B.; Tóth, A.; Zeke, I.; Jónás, G.; Friedrich, L. Changes of protein profiles in pork and beef meat caused by high hydrostatic pressure treatment. Acta Aliment. 2016, 45, 565–571. [Google Scholar] [CrossRef] [Green Version]
  75. Hajós, G.; Polgár, M.; Farkas, J. High-pressure effects on IgE immunoreactivity of proteins in a sausage batter. Innov. Food Sci. Emerg. Technol. 2004, 5, 443–449. [Google Scholar] [CrossRef]
  76. Benzie, I.F.F.; Strain, J.J. The Ferric Reducing Ability of Plasma (FRAP) as a Measure of “Antioxidant Power”: The FRAP Assay. Anal. Biochem. 1996, 239, 70–76. [Google Scholar] [CrossRef] [Green Version]
  77. Fagyas, M.; Úri, K.; Siket, I.M.; Daragó, A.; Boczán, J.; Bányai, E.; Édes, I.; Papp, Z.; Tóth, A. New Perspectives in the Renin-Angiotensin-Aldosterone System (RAAS) I: Endogenous Angiotensin Converting Enzyme (ACE) Inhibition. PLoS ONE 2014, 9, e87843. [Google Scholar] [CrossRef] [Green Version]
  78. Bradford, M.M. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 1976, 72, 248–254. [Google Scholar] [CrossRef]
  79. Hussein, K.; Friedrich, L.; Kisko, G.; Ayari, E.; Nemeth, C.; Dalmadi, I. Use of allyl-isothiocyanate and carvacrol to preserve fresh chicken meat during chilling storage. Czech J. Food Sci. 2019, 37, 417–424. [Google Scholar] [CrossRef]
  80. De Wit, J.N. Nutritional and Functional Characteristics of Whey Proteins in Food Products. J. Dairy Sci. 1998, 81, 597–608. [Google Scholar] [CrossRef]
  81. Penfield, M.P.; Campbell, A.M.; Penfield, M.P.; Campbell, A.M. Chapter 8—Milk and Milk Products. In Food Science and Technology, 3rd ed.; Penfield, M.P., Campbell, A.M.B.T.-E.F.S., Eds.; Academic Press: San Diego, CA, USA, 1990; pp. 162–183. ISBN 978-0-12-157920-3. [Google Scholar]
  82. Relkin, P.; Mulvihill, D.M. Thermal unfolding of β-lactoglobulin, α-lactalbumin, and bovine serum albumin. A thermodynamic approach. Crit. Rev. Food Sci. Nutr. 1996, 36, 565–601. [Google Scholar] [CrossRef] [PubMed]
  83. Lee, Y.-R.; Hong, Y.-H. Electrophoretic Behaviors of α-Lactalbumin and β-Lactoglobulin Mixtures Caused by Heat Treatment. Asian Australas. J. Anim. Sci. 2003, 16, 1041–1045. [Google Scholar] [CrossRef]
  84. Manderson, G.A.; Hardman, M.J.; Creamer, L.K. Effect of Heat Treatment on the Conformation and Aggregation of β-Lactoglobulin A, B, and C. J. Agric. Food Chem. 1998, 46, 5052–5061. [Google Scholar] [CrossRef]
  85. Law, A.J.R.; Horne, D.S.; Banks, J.M.; Leaver, J. Heat-induced changes in the whey proteins and caseins. Milchwissenschaft 1994, 49, 125–129. [Google Scholar]
  86. Dannenberg, F.; Kessler, H.-G. Reaction Kinetics of the Denaturation of Whey Proteins in Milk. J. Food Sci. 1988, 53, 258–263. [Google Scholar] [CrossRef]
  87. Dannenberg, F.; Kessler, H.G. Thermodynamic approach to kinetics of β-lactoglobulin denaturation in heated skim milk and sweet whey. Milchwissenschaft 1988, 43, 139–142. [Google Scholar]
  88. Dalgleish, D.G. Denaturation and aggregation of serum proteins and caseins in heated milk. J. Agric. Food Chem. 1990, 38, 1995–1999. [Google Scholar] [CrossRef]
  89. Gezimati, J.; Singh, H.; Creamer, L.K. Heat-Induced Interactions and Gelation of Mixtures of Bovine β-Lactoglobulin and Serum Albumin. J. Agric. Food Chem. 1996, 44, 804–810. [Google Scholar] [CrossRef]
  90. Sawyer, W.H. Complex Between β-Lactoglobulin and ĸ-Casein. A Review. J. Dairy Sci. 1969, 52, 1347–1355. [Google Scholar] [CrossRef]
  91. Sawyer, W.H.; Coulter, S.T.; Jenness, R. Role of Sulfhydryl Groups in the Interaction of κ-Casein and β-Lactoglobulin. J. Dairy Sci. 1963, 46, 564–565. [Google Scholar] [CrossRef]
  92. Jang, H.; Swaisgood, H. Disulfide Bond Formation Between Thermally Denatured β-Lactoglobulin and κ-Casein in Casein Micelles. J. Dairy Sci. 1990, 73, 900–904. [Google Scholar] [CrossRef]
  93. Morr, C.V.; Ha, E.Y.W. Whey protein concentrates and isolates: Processing and functional properties. Crit. Rev. Food Sci. Nutr. 1993, 33, 431–476. [Google Scholar] [CrossRef] [PubMed]
  94. Buchert, J.; Ercili Cura, D.; Ma, H.; Gasparetti, C.; Monogioudi, E.; Faccio, G.; Mattinen, M.; Boer, H.; Partanen, R.; Selinheimo, E.; et al. Crosslinking Food Proteins for Improved Functionality. Annu. Rev. Food Sci. Technol. 2010, 1, 113–138. [Google Scholar] [CrossRef]
  95. Gerrard, J.A. Protein–protein crosslinking in food: Methods, consequences, applications. Trends Food Sci. Technol. 2002, 13, 391–399. [Google Scholar] [CrossRef]
  96. Singh, H. Modification of food proteins by covalent crosslinking. Trends Food Sci. Technol. 1991, 2, 196–200. [Google Scholar] [CrossRef]
  97. Miranda, G.; Bianchi, L.; Krupova, Z.; Trossat, P.; Martin, P. An improved LC–MS method to profile molecular diversity and quantify the six main bovine milk proteins, including genetic and splicing variants as well as post-translationally modified isoforms. Food Chem. X 2020, 5, 100080. [Google Scholar] [CrossRef]
  98. Singh, H. Heat stability of milk. Int. J. Dairy Technol. 2004, 57, 111–119. [Google Scholar] [CrossRef]
  99. Kastrup Dalsgaard, T.; Holm Nielsen, J.; Bach Larsen, L. Proteolysis of milk proteins lactosylated in model systems. Mol. Nutr. Food Res. 2007, 51, 404–414. [Google Scholar] [CrossRef]
  100. Czerwenka, C.; Maier, I.; Pittner, F.; Lindner, W. Investigation of the lactosylation of whey proteins by liquid chromatography—Mass spectrometry. J. Agric. Food Chem. 2006, 54, 8874–8882. [Google Scholar] [CrossRef]
  101. Losito, I.; Carbonara, T.; Monaci, L.; Palmisano, F. Evaluation of the thermal history of bovine milk from the lactosylation of whey proteins: An investigation by liquid chromatography—Electrospray ionization mass spectrometry. Anal. Bioanal. Chem. 2007, 389, 2065–2074. [Google Scholar] [CrossRef]
  102. Almaas, H.; Cases, A.-L.; Devold, T.; Holm, H.; Langsrud, T.; Aabakken, L.; Aadnoey, T.; Vegarud, G. In vitro digestion of bovine and caprine milk by human gastric and duodenal enzymes. Int. Dairy J. 2006, 16, 961–968. [Google Scholar] [CrossRef]
  103. Costa, F.F.; Vasconcelos Paiva Brito, M.A.; Moreira Furtado, M.A.; Martins, M.F.; Leal de Oliveira, M.A.; Mendonça De Castro Barra, P.; Amigo Garrido, L.; Siqueira De Oliveira Dos Santos, A. Microfluidic chip electrophoresis investigation of major milk proteins: Study of buffer effects and quantitative approaching. Anal. Methods 2014, 6, 1666–1673. [Google Scholar] [CrossRef] [Green Version]
  104. Wróblewska, B.; Kaliszewska, A. Cow’s Milk Proteins Immunoreactivity and Allergenicity in Processed Food. Czech J. Food Sci. 2012, 30, 211–219. [Google Scholar] [CrossRef] [Green Version]
  105. Hong, Y.-H.; Creamer, L.K. Changed protein structures of bovine β-lactoglobulin B and α-lactalbumin as a consequence of heat treatment. Int. Dairy J. 2002, 12, 345–359. [Google Scholar] [CrossRef]
  106. Dunnill, P.; Green, D.W. Sulphydryl groups and the N⇄R conformational change in β-lactoglobulin. J. Mol. Biol. 1966, 15, 147–151. [Google Scholar] [CrossRef]
  107. Havea, P.; Singh, H.; Creamer, L.K. Characterization of heat-induced aggregates of β-lactoglobulin, α-lactalbumin and bovine serum albumin in a whey protein concentrate environment. J. Dairy Res. 2001, 68, 483–497. [Google Scholar] [CrossRef] [Green Version]
  108. Raak, N.; Abbate, R.; Lederer, A.; Rohm, H.; Jaros, D. Size Separation Techniques for the Characterisation of Cross-Linked Casein: A Review of Methods and Their Applications. Separations 2018, 5, 14. [Google Scholar] [CrossRef] [Green Version]
  109. Wu, Z.; Tiambeng, T.N.; Cai, W.; Chen, B.; Lin, Z.; Gregorich, Z.R.; Ge, Y. Impact of Phosphorylation on the Mass Spectrometry Quantification of Intact Phosphoproteins. Anal. Chem. 2018, 90, 4935–4939. [Google Scholar] [CrossRef] [Green Version]
  110. Bobe, G.; Beitz, D.C.; Freeman, A.E.; Lindberg, G.L. Sample Preparation Affects Separation of Whey Proteins by Reversed-Phase High-Performance Liquid Chromatography. J. Agric. Food Chem. 1998, 46, 1321–1325. [Google Scholar] [CrossRef]
  111. Cheema, M.; Mohan, M.S.; Campagna, S.R.; Jurat-Fuentes, J.L.; Harte, F.M. The association of low-molecular-weight hydrophobic compounds with native casein micelles in bovine milk. J. Dairy Sci. 2015, 98, 5155–5163. [Google Scholar] [CrossRef] [Green Version]
  112. Assem, F.M.; Abd El-Gawad, M.A.M.; Kassem, J.M.; Abd El-Salam, M.H. Proteolysis and antioxidant activity of peptic, tryptic and chymotryptic hydrolysates of cow, buffalo, goat and camel caseins. Int. J. Dairy Technol. 2018, 71, 236–242. [Google Scholar] [CrossRef]
  113. Power, O.; Jakeman, P.; FitzGerald, R.J. Antioxidative peptides: Enzymatic production, in vitro and in vivo antioxidant activity and potential applications of milk-derived antioxidative peptides. Amino Acids 2013, 44, 797–820. [Google Scholar] [CrossRef] [PubMed]
  114. Manso, M.A.; López-Fandiño, R. Angiotensin I converting enzyme-inhibitory activity of bovine, ovine, and caprine kappa-casein macropeptides and their tryptic hydrolysates. J. Food Prot. 2003, 66, 1686–1692. [Google Scholar] [CrossRef]
  115. Vermeirssen, V.; Van Camp, J.; Decroos, K.; Van Wijmelbeke, L.; Verstraete, W. The Impact of Fermentation and In Vitro Digestion on the Formation of Angiotensin-I-Converting Enzyme Inhibitory Activity from Pea and Whey Protein. J. Dairy Sci. 2003, 86, 429–438. [Google Scholar] [CrossRef]
  116. Wang, C.; Tu, M.; Wu, D.; Chen, H.; Chen, C.; Wang, Z.; Jiang, L. Identification of an ACE-Inhibitory Peptide from Walnut Protein and Its Evaluation of the Inhibitory Mechanism. Int. J. Mol. Sci. 2018, 19, 1156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Pihlanto-Leppälä, A.; Rokka, T.; Korhonen, H. Angiotensin I converting enzyme inhibitory peptides derived from bovine milk proteins. Int. Dairy J. 1998, 8, 325–331. [Google Scholar] [CrossRef]
  118. Abubakar, A.; Saito, T.; Kitazawa, H.; Kawai, Y.; Itoh, T. Structural Analysis of New Antihypertensive Peptides Derived from Cheese Whey Protein by Proteinase K Digestion. J. Dairy Sci. 1998, 81, 3131–3138. [Google Scholar] [CrossRef]
  119. López-Fandiño, R.; Otte, J.; Van Camp, J. Physiological, chemical and technological aspects of milk-protein-derived peptides with antihypertensive and ACE-inhibitory activity. Int. Dairy J. 2006, 16, 1277–1293. [Google Scholar] [CrossRef]
  120. Cheung, H.S.; Wang, F.L.; Ondetti, M.A.; Sabo, E.F.; Cushman, D.W. Binding of peptide substrates and inhibitors of angiotensin-converting enzyme. Importance of the COOH-terminal dipeptide sequence. J. Biol. Chem. 1980, 255, 401–407. [Google Scholar]
  121. Medeiros, V.; Rainha, N.; Paiva, L.; Lima, E.; Baptista, J. Bovine Milk Formula Based on Partial Hydrolysis of Caseins by Bromelain Enzyme: Better Digestibility and Angiotensin-Converting Enzyme-Inhibitory Properties. Int. J. Food Prop. 2014, 17, 806–817. [Google Scholar] [CrossRef]
  122. Malanovic, N.; Lohner, K. Antimicrobial Peptides Targeting Gram-Positive Bacteria. Pharmaceuticals 2016, 9, 59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Travkova, O.G.; Moehwald, H.; Brezesinski, G. The interaction of antimicrobial peptides with membranes. Adv. Colloid Interface Sci. 2017, 247, 521–532. [Google Scholar] [CrossRef] [PubMed]
  124. Qian, F.; Sun, J.; Cao, D.; Tuo, Y.; Jiang, S.; Mu, G. Experimental and Modelling Study of the Denaturation of Milk Protein by Heat Treatment. Korean J. Food Sci. Anim. Resour. 2017, 37, 44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Abe, S.; Kabashima, K.; Moriyama, T.; Tokura, Y. Food-dependent anaphylaxis with serum IgE immunoreactive to dairy products containing high-molecular-weight proteins. J. Dermatol. Sci. 2010, 57, 137–140. [Google Scholar] [CrossRef]
  126. Natale, M.; Bisson, C.; Monti, G.; Peltran, A.; Perono Garoffo, L.; Valentini, S.; Fabris, C.; Bertino, E.; Coscia, A.; Conti, A.; et al. Cow’s milk allergens identification by two-dimensional immunoblotting and mass spectrometry. Mol. Nutr. Food Res. 2004, 48, 363–369. [Google Scholar] [CrossRef]
  127. Meyer, P.; Petermeier, J.; Hartinger, M.; Kulozik, U. Concentration of Skim Milk by a Cascade Comprised of Ultrafiltration and Nanofiltration: Investigation of the Nanofiltration of Skim Milk Ultrafiltration Permeate. Food Bioprocess Technol. 2017, 10, 469–478. [Google Scholar] [CrossRef]
  128. Mistry, V.V.; Hassan, H.N. Delactosed, High Milk Protein Powder. 1. Manufacture and Composition. J. Dairy Sci. 1991, 74, 1163–1169. [Google Scholar] [CrossRef]
  129. Luo, X.; Ramchandran, L.; Vasiljevic, T. Lower ultrafiltration temperature improves membrane performance and emulsifying properties of milk protein concentrates. Dairy Sci. Technol. 2015, 95, 15–31. [Google Scholar] [CrossRef] [Green Version]
  130. Rinaldoni, A.N.; Tarazaga, C.C.; Campderrós, M.E.; Padilla, A.P. Assessing performance of skim milk ultrafiltration by using technical parameters. J. Food Eng. 2009, 92, 226–232. [Google Scholar] [CrossRef]
  131. Grandison, A.S.; Youravong, W.; Lewis, M.J. Hydrodynamic factors affecting flux and fouling during ultrafiltration of skimmed milk. Le Lait 2000, 80, 165–174. [Google Scholar] [CrossRef]
  132. Meena, G.S.; Singh, A.K.; Arora, S.; Borad, S.; Sharma, R.; Gupta, V.K. Physico-chemical, functional and rheological properties of milk protein concentrate 60 as affected by disodium phosphate addition, diafiltration and homogenization. J. Food Sci. Technol. 2017, 54, 1678–1688. [Google Scholar] [CrossRef] [Green Version]
  133. Meena, G.S.; Singh, A.K.; Gupta, V.K.; Borad, S.; Parmar, P.T. Effect of change in pH of skim milk and ultrafiltered/diafiltered retentates on milk protein concentrate (MPC70) powder properties. J. Food Sci. Technol. 2018, 55, 3526–3537. [Google Scholar] [CrossRef] [PubMed]
  134. Meyer, P.; Mayer, A.; Kulozik, U. High concentration of skim milk proteins by ultrafiltration: Characterisation of a dynamic membrane system with a rotating membrane in comparison with a spiral wound membrane. Int. Dairy J. 2015, 51, 75–83. [Google Scholar] [CrossRef]
  135. Samuelsson, G.; Dejmek, P.; Trägårdh, G.; Paulsson, M. Minimizing whey protein retention in cross-flow microfiltration of skim milk. Int. Dairy J. 1997, 7, 237–242. [Google Scholar] [CrossRef]
  136. Pompei, C.; Resmini, P.; Peri, C. Skim Milk Protein Recovery and Purification by Ultrrafiltration Influence of Temperature on Permeation Rate and Retention. J. Food Sci. 1973, 38, 867–870. [Google Scholar] [CrossRef]
  137. Meena, G.S.; Singh, A.K.; Gupta, V.K.; Borad, S.; Arora, S.; Tomar, S.K. Effect of pH adjustment, homogenization and diafiltration on physicochemical, reconstitution, functional and rheological properties of medium protein milk protein concentrates (MPC70). J. Food Sci. Technol. 2018, 55, 1376–1386. [Google Scholar] [CrossRef]
  138. Meena, G.S.; Singh, A.K.; Gupta, V.K.; Borad, S.G.; Arora, S.; Tomar, S.K. Alteration in physicochemical, functional, rheological and reconstitution properties of milk protein concentrate powder by pH, homogenization and diafiltration. J. Food Sci. Technol. 2019, 56, 1622–1630. [Google Scholar] [CrossRef]
Figure 1. Experimental steps for preparing hypoallergenic liquid milk protein concentrate with functional values (antioxidant capacity, angiotensin converting enzyme inhibitory activity, and antibacterial activity).
Figure 1. Experimental steps for preparing hypoallergenic liquid milk protein concentrate with functional values (antioxidant capacity, angiotensin converting enzyme inhibitory activity, and antibacterial activity).
Processes 08 00871 g001
Figure 2. Schematic diagram of the cross-flow membrane module.
Figure 2. Schematic diagram of the cross-flow membrane module.
Processes 08 00871 g002
Figure 3. Specific energy consumption and concentration of protein in the retentate side of membrane for different trans-membrane pressures (TMPs) and retention flow rates (RFRs) in static turbulence promoter-implemented filtration process. Results are represented by mean value with standard deviation (±values). In superscript, dissimilar alphabet represents the significant difference between results, evaluated by the Tukey’s post hoc method.
Figure 3. Specific energy consumption and concentration of protein in the retentate side of membrane for different trans-membrane pressures (TMPs) and retention flow rates (RFRs) in static turbulence promoter-implemented filtration process. Results are represented by mean value with standard deviation (±values). In superscript, dissimilar alphabet represents the significant difference between results, evaluated by the Tukey’s post hoc method.
Processes 08 00871 g003
Figure 4. Time history of permeate flux without the static turbulence promoter (A) and with the static turbulence promoter (B).
Figure 4. Time history of permeate flux without the static turbulence promoter (A) and with the static turbulence promoter (B).
Processes 08 00871 g004
Figure 5. Results of liquid chromatography-mass spectrometry; (A) UV chromatogram of proteins in concentrated milk, (B) total ion chromatogram of proteins in concentrated milk, (C) deconvoluted mass spectra of different proteins, peaks appear at retention time 5.4 min (C1), 5.6 min (C2), and 5.8 min (C3).
Figure 5. Results of liquid chromatography-mass spectrometry; (A) UV chromatogram of proteins in concentrated milk, (B) total ion chromatogram of proteins in concentrated milk, (C) deconvoluted mass spectra of different proteins, peaks appear at retention time 5.4 min (C1), 5.6 min (C2), and 5.8 min (C3).
Processes 08 00871 g005
Figure 6. Sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) image of ultra-heat-treated skimmed milk, milk with concentrated proteins and milk with concentrated proteins after enzyme treatment; lane 1: marker protein, lane 2: standard casein, lane 3: standard α-lactalbumin and β-lactoglobulin, lane 4: ultra-heat-treated skimmed milk, lane 5: concentrated ultra-heat-treated skimmed milk, lane 6: concentrated liquid milk protein treated with 0.008 g L−1 of trypsin, lane 7: concentrated liquid milk protein treated with 0.016 g L−1 of trypsin, lane 8:concentrated liquid milk protein treated with 0.032 g L−1 of trypsin, lane 9: concentrated liquid milk protein treated with 0.064 g L−1 of trypsin.
Figure 6. Sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) image of ultra-heat-treated skimmed milk, milk with concentrated proteins and milk with concentrated proteins after enzyme treatment; lane 1: marker protein, lane 2: standard casein, lane 3: standard α-lactalbumin and β-lactoglobulin, lane 4: ultra-heat-treated skimmed milk, lane 5: concentrated ultra-heat-treated skimmed milk, lane 6: concentrated liquid milk protein treated with 0.008 g L−1 of trypsin, lane 7: concentrated liquid milk protein treated with 0.016 g L−1 of trypsin, lane 8:concentrated liquid milk protein treated with 0.032 g L−1 of trypsin, lane 9: concentrated liquid milk protein treated with 0.064 g L−1 of trypsin.
Processes 08 00871 g006
Figure 7. Antioxidant capacity of ultra-heat-treated skimmed milk, milk with concentrated proteins, and milk with concentrated proteins after enzyme treatment. Results are represented by mean value with standard deviation (±values). In superscript, dissimilar alphabet represents the significant difference between results, evaluated by the Tukey’s post hoc method.
Figure 7. Antioxidant capacity of ultra-heat-treated skimmed milk, milk with concentrated proteins, and milk with concentrated proteins after enzyme treatment. Results are represented by mean value with standard deviation (±values). In superscript, dissimilar alphabet represents the significant difference between results, evaluated by the Tukey’s post hoc method.
Processes 08 00871 g007
Figure 8. Angiotensin converting enzyme (ACE)-inhibitory activity of ultra-heat-treated skimmed milk and milk with concentrated proteins (A), and values of IC50 in milk with concentrated proteins after enzyme treatment (B). Results are represented by mean value with standard deviation (±values). In superscript, dissimilar alphabet represents the significant difference between results, evaluated by the Tukey’s post hoc method.
Figure 8. Angiotensin converting enzyme (ACE)-inhibitory activity of ultra-heat-treated skimmed milk and milk with concentrated proteins (A), and values of IC50 in milk with concentrated proteins after enzyme treatment (B). Results are represented by mean value with standard deviation (±values). In superscript, dissimilar alphabet represents the significant difference between results, evaluated by the Tukey’s post hoc method.
Processes 08 00871 g008
Figure 9. Antibacterial activity of milk with concentrated proteins after enzyme treatment. Results are represented by mean value with standard deviation (±values). Superscript dissimilar alphabet represents the significant difference between results, evaluated by the Tukey’s post hoc method.
Figure 9. Antibacterial activity of milk with concentrated proteins after enzyme treatment. Results are represented by mean value with standard deviation (±values). Superscript dissimilar alphabet represents the significant difference between results, evaluated by the Tukey’s post hoc method.
Processes 08 00871 g009
Figure 10. Immunoblot of ultra-heat-treated skimmed milk, milk with concentrated proteins and milk with concentrated proteins after enzyme treatment; lane 1: marker protein, lane 2: standard casein, lane 3: standard α-lactalbumin and β-lactoglobulin, lane 4: ultra-heat-treated skimmed milk, lane 5: concentrated ultra-heat-treated skimmed milk, lane 6: concentrated liquid milk protein treated with 0.008 g·L−1 of trypsin, lane 7: concentrated liquid milk protein treated with 0.016 g·L−1 of trypsin, lane 8: concentrated liquid milk protein treated with 0.032 g·L−1 of trypsin, lane 9: concentrated liquid milk protein treated with 0.064 g·L−1 of trypsin.
Figure 10. Immunoblot of ultra-heat-treated skimmed milk, milk with concentrated proteins and milk with concentrated proteins after enzyme treatment; lane 1: marker protein, lane 2: standard casein, lane 3: standard α-lactalbumin and β-lactoglobulin, lane 4: ultra-heat-treated skimmed milk, lane 5: concentrated ultra-heat-treated skimmed milk, lane 6: concentrated liquid milk protein treated with 0.008 g·L−1 of trypsin, lane 7: concentrated liquid milk protein treated with 0.016 g·L−1 of trypsin, lane 8: concentrated liquid milk protein treated with 0.032 g·L−1 of trypsin, lane 9: concentrated liquid milk protein treated with 0.064 g·L−1 of trypsin.
Processes 08 00871 g010
Table 1. Difference of pressure, initial permeate flux and percentage change of permeate flux for different trans-membrane pressures (TMPs) and retention flow rates (RFRs) in absence and presence of static turbulence promoter. Results are represented by mean value with standard deviation (±values). In superscript, dissimilar alphabet represents the significant difference between results, evaluated by the Tukey’s post hoc method.
Table 1. Difference of pressure, initial permeate flux and percentage change of permeate flux for different trans-membrane pressures (TMPs) and retention flow rates (RFRs) in absence and presence of static turbulence promoter. Results are represented by mean value with standard deviation (±values). In superscript, dissimilar alphabet represents the significant difference between results, evaluated by the Tukey’s post hoc method.
TMP (Bar)RFR (L·h−1)Without StaticWith Static
p (Bar)Jinitial (L·h−1·m−2)J (%)p (Bar)Jinitial (L·h−1·m−2)J (%)
21000.18.06 ± 1 a41.69 ± 1.27 a0.315.58 ± 1.1 a32.33 ± 1.25 a
22000.18.2 ± 1.2 a37.39 ± 2.35 a b0.715.88 ± 1 a31.70 ± 2.5 a
31000.113.45 ± 1.1 b36.95 ± 1.55 a, b0.334.22 ± 1.08 b24.01 ± 1.19 b
32000.118 ± 3.9 b33.33±2.79 b0.734.55 ± 1.02 b23.61 ± 2.31 b

Share and Cite

MDPI and ACS Style

Nath, A.; Eren, B.A.; Csighy, A.; Pásztorné-Huszár, K.; Kiskó, G.; Abrankó, L.; Tóth, A.; Szerdahelyi, E.; Kovács, Z.; Koris, A.; et al. Production of Liquid Milk Protein Concentrate with Antioxidant Capacity, Angiotensin Converting Enzyme Inhibitory Activity, Antibacterial Activity, and Hypoallergenic Property by Membrane Filtration and Enzymatic Modification of Proteins. Processes 2020, 8, 871. https://doi.org/10.3390/pr8070871

AMA Style

Nath A, Eren BA, Csighy A, Pásztorné-Huszár K, Kiskó G, Abrankó L, Tóth A, Szerdahelyi E, Kovács Z, Koris A, et al. Production of Liquid Milk Protein Concentrate with Antioxidant Capacity, Angiotensin Converting Enzyme Inhibitory Activity, Antibacterial Activity, and Hypoallergenic Property by Membrane Filtration and Enzymatic Modification of Proteins. Processes. 2020; 8(7):871. https://doi.org/10.3390/pr8070871

Chicago/Turabian Style

Nath, Arijit, Burak Atilla Eren, Attila Csighy, Klára Pásztorné-Huszár, Gabriella Kiskó, László Abrankó, Attila Tóth, Emőke Szerdahelyi, Zoltán Kovács, András Koris, and et al. 2020. "Production of Liquid Milk Protein Concentrate with Antioxidant Capacity, Angiotensin Converting Enzyme Inhibitory Activity, Antibacterial Activity, and Hypoallergenic Property by Membrane Filtration and Enzymatic Modification of Proteins" Processes 8, no. 7: 871. https://doi.org/10.3390/pr8070871

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop