Next Article in Journal
Implementation of an Analytical Method for Spectrophotometric Evaluation of Total Phenolic Content in Essential Oils
Previous Article in Journal
Flexible Liquid Crystal Polymer Technologies from Microwave to Terahertz Frequencies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antioxidant and Antimicrobial Peptides Derived from Food Proteins

by
Guadalupe López-García
1,
Octavio Dublan-García
1,*,
Daniel Arizmendi-Cotero
2 and
Leobardo Manuel Gómez Oliván
1
1
Food and Environmental Toxicology Laboratory, Chemistry Faculty, Universidad Autónoma del Estado de México, Paseo Colón Intersección Paseo Tollocan s/n. Col. Residencial Colón, Toluca 50120, Mexico
2
Department of Industrial Engineering, Engineering Faculty, Campus Toluca, Universidad Tecnológica de México (UNITEC), Estado de México, Toluca 50160, Mexico
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(4), 1343; https://doi.org/10.3390/molecules27041343
Submission received: 30 December 2021 / Revised: 11 February 2022 / Accepted: 13 February 2022 / Published: 16 February 2022

Abstract

:
Recently, the demand for food proteins in the market has increased due to a rise in degenerative illnesses that are associated with the excessive production of free radicals and the unwanted side effects of various drugs, for which researchers have suggested diets rich in bioactive compounds. Some of the functional compounds present in foods are antioxidant and antimicrobial peptides, which are used to produce foods that promote health and to reduce the consumption of antibiotics. These peptides have been obtained from various sources of proteins, such as foods and agri-food by-products, via enzymatic hydrolysis and microbial fermentation. Peptides with antioxidant properties exert effective metal ion (Fe2+/Cu2+) chelating activity and lipid peroxidation inhibition, which may lead to notably beneficial effects in promoting human health and food processing. Antimicrobial peptides are small oligo-peptides generally containing from 10 to 100 amino acids, with a net positive charge and an amphipathic structure; they are the most important components of the antibacterial defense of organisms at almost all levels of life—bacteria, fungi, plants, amphibians, insects, birds and mammals—and have been suggested as natural compounds that neutralize the toxicity of reactive oxygen species generated by antibiotics and the stress generated by various exogenous sources. This review discusses what antioxidant and antimicrobial peptides are, their source, production, some bioinformatics tools used for their obtainment, emerging technologies, and health benefits.

1. Introduction

Human beings require oxygen (O2) to produce energy via the oxidative metabolism of the mitochondria; they produce ATP, which is the molecule from which the energy necessary for various vital processes is released. However, an excess of O2 in cells is harmful due to the formation of reactive species generated during oxidation. The mitochondria are considered to be the main endogenous source of free radicals (FRs) since, in order to produce ATP, the oxidative metabolism reduces 95–98% of the oxygen in a tetra-electronic manner; the rest is reduced in a monoelectronic manner, forming intermediaries known as reactive oxygen species (ROS) or reactive nitrogen species (RNS). These are normally produced within the body and act as a defense against infections by bacteria and viruses, participate in the maturation of the reticulocytes, and participate in the degradation of proteins. Each mitochondria produces approximately 7–10 M FR/day and is capable of producing other reactive species of biological relevance that are formed because of a cascade reaction, such as nitric oxide (NO), peroxynitrite (ONOO-), and hydrogen peroxide (H2O2) [1].
Oxidative stress is responsible for cellular degeneration because FRs can chemically react with proteins, lipids, and DNA, thereby producing various modifications, cell damage, and even cell death. It has been reported that FRs oxidize the amino acids of proteins, leading to the formation of carbonyl groups, the association of protein fragments via the cross-linking of disulfide bonds, the breaking of peptide bonds, the loss of affinity for metals, and an increase in hydrophobicity, resulting in proteins with oxidative damage that present deteriorations in hormonal and enzymatic activity, as well as ion transport, and a greater susceptibility to proteolytic degradation [2]. Polyunsaturated fatty acids (PUFAs), once oxidized by FRs, lead to lipoperoxidation, a reaction in which PUFAs give their electrons to FRs and generate changes in the molecular structure of the membrane, as well as the formation of disulfide bonds between membrane proteins, that leads to a loss in permeability and the stability of the membrane, resulting in cell death [3].
The free radicals also attack DNA, damaging the genes that codify proteins, which leads to a lack of execution in cell functions. It has been reported that OH· is responsible for various modifications in deoxyribose resulting in the release of nitrogenated bases (which are linked together by the aforementioned sugar) and the breaking of one or both chains, resulting in deletions, mutations, chromosomic rearrangements, and the activation or inactivation of genes, all of which affect the biosynthesis of DNA chains; likewise, reactive oxygen species produce errors during the transcription and translation of RNA [4,5].
To counter the harmful effects of O2 and its derivatives, the cell possesses enzymes, electron scavengers and nutrients that act as an antioxidant system in charge of maintaining the equilibrium of oxidation–reduction reactions and cell survival. However, FRs are also generated by exogenous sources and factors that favor their formation, such as exposure to X-rays, ozone, tobacco, air contaminants, and industrial products; certain medications and inadequate eating habits; the consumption of foods with low nutritional quality and antioxidant capacity, fast food with high fat content, junk food, canned goods that contain additives, drinks with high sugar content; and the low consumption of natural foods, which provokes an increase in the concentration of FRs and a loss of equilibrium between the speed of formation and their neutralization by the body’s endogenous antioxidant system. This leads to oxidative stress, which has been linked to obesity, chronic degenerative illnesses, neurodegenerative illnesses, cardiovascular illnesses, diabetes mellitus and cancer, as well as severe cell damage [6].
The harmful effects of oxidative stress on human health can be reduced via the ingestion of dietary antioxidants present in various foods. A balance must be achieved between the abundance of ROS and antioxidants, which are compounds capable of donating electrons, to stabilize the free radicals and neutralize their harmful effects. These can be of endogenous origin (synthesized within the body), and superoxide dismutase and glutathione peroxidase are among the endogenous antioxidants [7]).
Exogenous antioxidants (from external sources), such as coenzyme Q and bioactive peptides (antioxidants and antimicrobials), that humans can obtain from their diet can act in two ways: (1) preventing the excessive production of free radicals, thus avoiding cell damage due to the effects of oxidative stress; or (2) controlling the levels of free radicals after damage has been produced, thus preventing further damage and thereby preventing some symptoms of illnesses produced by the effects of oxidative stress [7,8].
Many studies have suggested that some side effects produced by various antibiotics such as amoxicillin, dapsone, vancomycin, ampicillin, doxorubicin, isoniazid, rifampin, ciprofloxacin, enrofloxacin, chloramphenicol, and gentamicin can increase oxidative stress in human cells [9,10,11,12,13,14,15,16]. This has motivated the search for new molecules with antimicrobial activity that are effective with fewer side effects. Antimicrobial peptides are characterized by being positively charged molecules and presenting hydrophobic residues, which can act against a wide spectrum of microorganisms including Gram-positive bacteria and Gram-negative bacteria, viruses and fungi; accordingly, these peptides can be used to contribute adjuvants to medication and reduce the side effects of these drugs [7].
The present review focuses on what antioxidant and antimicrobial peptides are, their source, production, some bioinformatics tools for their obtainment, emerging technologies and health benefits.

2. Bioactive Peptides

Bioactive peptides are short sequences of 2–40 units of amino acids that are inactive within the precursor protein and, once released via chemical or enzymatic hydrolysis, in vivo digestion, or food processing can carry out a variety of biological activities [17,18]. It has been reported that bioactive peptides exhibit hormonal activity or activity like that of drugs and can be classified according to their mechanism of action as antioxidants, antimicrobial agents, antithrombotic agents, antihypertensive agents, opioids, immunomodulators or metal chelators [19,20,21,22,23]. Several studies have indicated that these activities are determined by the degree of hydrolysis (number of peptide bonds broken in relation to the original protein), substrate concentration, enzyme/substrate ratio, incubation time, microorganism employed, and physicochemical conditions such as pH and temperature [24].

2.1. Sources of Bioactive Peptides

Peptides have been identified and isolated from animal and vegetable sources, and they are abundantly found in protein hydrolysates [23] and fermented dairy products, as well as some fungi such as Ganoderma lucidum [25]. They can be obtained via protein digestion by employing endogenous or exogenous enzymes, microbial fermentation, processing, or gastrointestinal digestion. Due to the relevance of these emerging molecules and with the aim of reducing costs, other protein sources have been sought; among these are food industry by-products, such as the heads and viscera of fish and bird feathers [26,27,28]. Some studies have reported that, depending on the protein source, the enzyme used, and the processing conditions, obtained biological activity and the peptides vary [21,29,30], as shown in Table 1.

2.1.1. Peptides of Animal Origin

Red meats, birds, fish and eggs are valuable sources of protein that are not present in other foods. Bioactive substances, such as fatty acids (conjugated linoleic acid), carnosine, L-carnitine, glutathione, taurine, coenzyme Q10, and creatin, have also been identified in proteins of animal origin [31], as have peptides that have been reported to be physiologically functional.
Livestock, sheep, goats, pigs and poultry are the most common sources of meat for human consumption and proteins. From pork, for example, RPR, KAPVA and PTPVP—peptides with antihypertensive activity—have been identified [32]. Additionally, the most-studied peptides are those derived from milk caseins, such as lactotransferrin, which possesses antimicrobial activity towards a wide range of microorganisms including Staphylococcus spp. and Streptococcus pyogenes [33].
Another source of bioactive peptides is seafood, such as algae, fish, mollusks and crustaceans, from which some endogenous bioactive peptides, antioxidants and angiotensin-converting enzyme inhibitors have been identified, through the fermentation of salmon, sardine, anchovy, carp and mackerel [27,30,34,35,36]. Table 1 shows some sources of bioactive peptides of animal origin.

2.1.2. Peptides of Vegetable Origin

The procurement of peptides from proteins of vegetable origin through enzymes of animal, vegetable, or microbial origin has also been reported. Some protein sources of vegetable origin from which bioactive peptides have been isolated are soy milk, soy seed, wheat, corn, rice, barley, and sunflower; it has been found that these present antihypertensive, hypercholesteremic, anti-obesity and anticarcinogenic effects [37]. Likewise, it has been reported that 51% of the peptides present in plants exhibit antimicrobial activity and a structure similar to those of animals and insects [38,39]. In Table 1, some sources of bioactive peptides of vegetable origin can be seen.

2.1.3. Fungi-Derived Peptides

Fungi are one of the most important natural sources of bioactive compounds due to the presence of numerous products with therapeutic properties. Commonly used in traditional Chinese medicine for centuries, fungi are rich in natural antioxidants, antimicrobials, antitumor agents, antiviral agents, and immunomodulatory agents, with medicinal effects proven by researchers [1,2,3]. It has been shown that fungi can produce many bioactive compounds such as polysaccharides, phenols, sterols, proteins. and peptides. Which are responsible for the therapeutic effects attributed to this species [52,53,54,55,56].
Fungi such as Ganoderma lucidum (Curtis: Fr.) Karsten (Ling-chih or reishi), Ganoderma tsugae, G. lucidum, and Marasmius oreades were analyzed by Sun et al. (2004) [25], who isolated a fraction of bioactive peptides called GLP (peptide from G. lucidum) [25]. This peptide fraction has a molecular mass <10,000 Da, while the N-terminal sequence is unknown. The GLP fraction showed a high antioxidant activity comparable to that of the synthetic antioxidant butylhydroxytoluene (BHT) in soybean oil and lard systems. The authors showed that the GLP fraction at 0.1% (w:w) after 13 days inhibited the oxidation of soybean oil and lard at rates of 60 and 30%, respectively. The GLP fraction blocked the oxidation of polyunsaturated fatty acids responsible for the formation of lipid hydroperoxides by inhibiting 90% of lipoxygenase activity in vitro at 0.3 mg/mL. Furthermore, using mouse liver homogenates, 0.7 mg/mL of GLP fraction inhibited 72% of H2O2-induced malondialdehyde formation resulting from the lipid peroxidation of polyunsaturated fatty acids.
Finally, although the antioxidant properties of G. lucidum are mainly attributed to the fraction of water-soluble polysaccharides, the authors showed that aqueous extracts of low molecular weight, mostly made up of peptides, showed greater antioxidant activity than fractions of G. lucidum, which has high molecular weight [25]. A bioactive peptide called plectasin extracted from P. nigrella mRNA (FGCNGPWDE DDMQCHNHCK SIKGYKGGYC AKGGFVCKCY) was identified [57,58]. With a molecular mass of 4398.80 Da, plectasin is considered a bactericide, with antimicrobial activity against Gram-positive bacteria tested at physiological ionic strength in contrast to most vertebrate defensins, which require very low ionic strength. On the other hand, this peptide is selective for bacterial cells over mammalian cells in vitro, as it has not shown cytotoxicity against murine L929 fibroblasts or normal human epidermal cells. In an in vivo study using mouse models of pneumococcal peritoneal infection, the anti-infective action of plectasin was highlighted [58].
Due to its therapeutic potential, plectasin is considered to be a novel antimicrobial peptide with excellent penetration into cerebrospinal fluid, suggesting its possible efficacy in the treatment of central nervous system infections caused by Gram-positive pathogens, such as pneumococcal meningitis.

2.1.4. Peptides from Agri-Food By-Products

Recent innovations for the generation of peptides with biological activity include the use of proteins from food by-products such as substrates for enzymatic and microbial action to take advantage of their technological properties, improve nutritional properties, reduce production costs. The by-products used as protein sources include bones from the meat industry, from which 10–13% of collagen and gelatin, which are proteins that contains biologically active peptides within their sequences, can be obtained [59]. The bones of hydrolyzed marine by-products, such as tuna bones, have been found to contain the VKAGFAWTANQQLS peptide, which has been identified as presenting antioxidant activity [60]. Hydrolyzed chicken skin, from which four inhibitor peptides of the angiotensin-converting enzyme can be extracted, has also been identified as a source [51]. Blood from bovine and porcine sources has also been evaluated; it has been reported that hemoglobin and plasma hydrolysates exhibit antihypertensive, antioxidant, antimicrobial and opioid activity [61]. Table 1 shows some bioactive peptides obtained from food by-products and their corresponding activity.
For the choice of a suitable protein source, the use that hydrolysate is going to have must be considered, as is the added value of the final product with respect to the initial substrate. For example, to obtain hydrolysates with gelling and emulsifying properties, collagen and gelatin are often used due to their ability to form transparent gels [55,62], but proteins from eggs, meat, blood, viscera, and even cereals have also been used. As a source of fermentation for the growth of microorganisms, yeast or casein hydrolysates have been used. They are also sources for hydrolysates that are used in cosmetics. When a hydrolysate is intended to be used as a nitrogen source, fish proteins and microbial proteins are used in animal feed and soy and dairy proteins (especially whey proteins, the ideal raw material for the preparation of infant foods and enteral diets) are used in human nutrition [63,64,65].

3. Production of Peptides

Bioactive peptides can be found in most food products rich in proteins, and the consumption of foods with higher contents of proteins increases the possibility of acquiring them; nonetheless, their release is achieved through various mechanisms: (1) heat, (2) acid–base conditions in which the protein is hydrolyzed, (3) enzymatic hydrolysis, and (4) microbial activity (mainly in fermented foods) [66]. Once the structure of the bioactive peptides is known, it is also possible to synthesize them via chemical synthesis, recombinant DNA technology, and enzymatic synthesis [67]. Figure 1 shows the general process used for the production and identification of antioxidant peptides.

3.1. Enzymatic Hydrolysis

The enzymatic hydrolysis of proteins is the most common way to produce bioactive peptides. Once the protein source is selected, the hydrolysis is carried out using one or multiple specific proteases for the release of the peptides of interest. The factors that influence the release of the peptides are the duration of hydrolysis, the degree of protein hydrolysis, the enzyme–substrate ratio, and the treatment of the proteins (prior to hydrolysis) [71]. A great number of known bioactive peptides have been produced using a digestive enzymatic process and different combinations of proteinases such as pepsin, trypsin, alcalase, chymotrypsin, pancreatin and thermolysin.
Some studies have shown that biologically active peptides can be produced via the hydrolysis of milk proteins by digestive enzymes [72,73]. Pepsin, trypsin, and chymotrypsin are important enzymes that have been demonstrated to release antihypertensive peptides, phosphopeptides bonded to calcium, antibacterial peptides, immunomodulators, opioid peptides, and serum proteins (α-lactalbumin, ß-lactoglobulin and glycomacropeptide [19,73,74,75,76]. Likewise, angiotensin-converting enzyme inhibitor peptides (and algae hydrolysates, which have also been reported as a source of bioactive peptides [21]) have been obtained from the enzymatic digestion of proteins from marine products such as tuna, sardine muscle, and mackerel [28,36,77,78]. Alcalase, thermolysin and subtilysin are examples of other proteolytic enzymes that have been employed to release various bioactive peptides [73,79], such as inhibitors of the angiotensin-converting enzyme [73,80,81,82,83], antibacterial agents [84,85], antioxidants [86], immunomodulators [87] and opioids [73,88].
The advantages of enzymatic hydrolysis are that it is more easily reproducible because it allows for full process control [89], it is easy to scale up and it generally has a shorter reaction time than microbial fermentation [90], it can be optimized through the control of physical–chemical parameters such as pH or temperature to enable the ideal conditions for proteases to act [91], and it can be carried out by proteases of animal, vegetable, or food-grade microbials. Additionally, other proteases, such as microbial proteases that have different specificities for substrates and can release different bioactive peptides, can be sequentially used. An example is the Alcalase® enzyme from Bacillus licheniformis, which has been used to obtain peptides with a good antioxidant, hypocholesterolemic and hypoglycemic profiles [92,93]. The proteases also have advantages in the generation of well-defined peptide profiles that have a high productivity rate and do not generate molecules that can harm health, as happens with chemical hydrolysis, which makes them viable for applications in the formulation of functional foods or nutraceuticals [94].
However, it is necessary to optimize the temperature and pH for each of the proteases [90]. Furthermore, both the choice of the protease used and the enzymatic hydrolysis time are important in deciding the type of peptides generated. Likewise, the stability of bioactive peptides produced during the phases of absorption, distribution, and metabolism must be evaluated to guarantee their stability in vivo [95]. To produce bioactive peptides, an international consensus methodology that simulates the physiological conditions of human gastrointestinal digestion in its different phases (oral, stomach and intestinal) has been established to guarantee the stability of bioactive peptides and facilitate their arrival to the systemic pathway where they can exert their therapeutic effects [96].

3.2. Microbial Fermentation

In this process, starter cultures are employed for the inoculation of microbial strains within a reactor that contains homogenous mixtures of concentrated proteins, water, and microbial nutrients such as simple sugars, which can be used for the growth and development of the strains [97] The protein breaks mainly due to the activity of the microbes and the peptidases secreted by them during fermentation [98]. Lactic acid bacteria have been mainly used for this purpose due to their high proteolytic activity. Some lactic acid bacteria employed as starter and non-starter cultures belong to strains of Lactococcus lactis, Lactobacillus helveticus and Lactobacillus delbrueckii ssp. bulgaricus; in their characterization processes, the participation of proteinases, coupled to the bacterial cell wall, has been observed, as has the presence of intracellular peptidases, including endopeptidases, aminopeptidases, tripeptidases and dipeptidases [99]. The type, quantity and activity of the peptides produced depend on the specific cultures employed. Peptides obtained via lactic acid bacteria from fermented goat milk have been reported to reduce oxidative stress and teratogenicity in human beings [100]. It has been reported that lactic acid bacteria used in yogurt fermentation produce peptides that can stabilize reactive oxygen species and inhibit lipoperoxidation [101]. Peptides from Leuconostoc mesenteroides ssp.; strains of Cremoris, Jensenii lactobacilli (ATCC25258), and Lactobacillus acidophilus (ATCC4356); and strains isolated from a whey fermentation system have also exhibited the inhibition of free radicals and lipoperoxidation [102]. In the same way, the heptapeptide HFGDPFH has been obtained from fermented mussel meats and exhibited high free radical scavenging activity [103].
Additionally, co-cultures with different combinations of bacteria, yeasts and fungi can be used to modulate hydrolysis processes, but the degree of hydrolysis depends on the fermentation type, time, microbial strain, and the protein source [104]; nevertheless, soybean proteins fermented by Bacillus subtilis MTCC5480 have been shown to produce a higher degree of hydrolysis than B. subtilis MTCC1747 [105].
Furthermore, this process is strongly influenced by environmental conditions such as humidity, pH, water oxide index, temperature, the availability of the substrate, the availability of free oxygen, and the size of the particles. To study the process on a pilot scale, researchers have developed bioreactors where heat and mass transfer are monitored, but this option is extremely expensive for commercial use. Likewise, the high density of the substrate makes it difficult to separate the biomass and therefore increases the energy required by the enzymes to produce metabolites [106].

3.3. Emerging Technologies for the Development of Antioxidant Peptides

Biochemical methods are the most common strategies to produce bioactive peptides; however, at an industrial level, a higher yield and lower cost is required; therefore, methods such as hydrolysis and technologies such as hydrostatic high-pressure, microwaves, and pulsated electric fields have been combined [107,108,109].
High pressure is a non-thermal treatment that applies pressures between 100 and 1000 MPa over a protein solution to modify the conformation and extension of the molecular chain of the proteins, aid the proteolysis reaction, and allow the enzymes to cleave peptide bonds in the new sites and produces bioactive peptides [110]. It has been observed that the high-pressure treatment improved the degree of hydrolysis, reducing power, and DPPH radical scavenging capacity of the peanut protein hydrolysate, with the effect of sequence on hydrolysis of 300 > 100 > 500 MPa (when all the treatments were 20 min) revealing that those subjected to 100 and 300 MPa were significantly higher (p < 0.05) than those treated with 500 MPa [109]. Likewise, it has been reported that the enzymatic hydrolysis of protein casein, under 100 MPa of pressure, improved the degree of hydrolysis and antioxidant properties compared to 200 MPa and atmospheric pressure with similar enzymatic hydrolysis [111]. These studies show that the treatment of proteins with high pressure improves the enzymatic hydrolysis and yield of antioxidant peptides, as well as bioactivity. Processing assisted by microwaves utilizes electromagnetic radiation, in a frequency range of 300 MHz–300 GHz, to heat a solvent in a sample more rapidly than conventional heating [57] due to inter- and intramolecular friction in conjunction with the movement and collision of ions, which provoke the rapid heating of the reaction and the rupture of the walls and cell membranes [112]. Microwave heating has been applied as a pretreatment followed by enzymatic hydrolysis, and it has been found that the use of this method increases the degree of hydrolysis, protein solubility, and radical scavenging, thus improving the accessibility and susceptibility of bonds to enzymes [110,113,114]. Studies have shown that peptides with antimicrobial activity have been obtained with pepsin from different variants of casein and from whey proteins [76]. Processing assisted by ultrasound uses the acoustic cavitation induced by ultrasound, which creates mechanical and thermal effects and modifies the physicochemical properties of proteins [115,116]. Some studies have provided evidence on the use of ultrasonic processing as a pre-treatment for the improvement of the protein hydrolysis procedure and to increase the antioxidant activity of the peptide hydrolysates. Wen et al. (2020) [108] employed type-S ultrasound microfrequencies as a pretreatment for the arrowhead protein for its posterior hydrolysis, and they showed that the hydrolysates obtained with double-frequency ultrasound (20/40 kHz) had an effect over the antioxidant activity of the hydrolysates without ultrasonic treatment. There was a 63.61% inhibition of DPPH• and a 65.11% inhibition of ABTS•+. The authors of [117] employed a 750 W ultrasound as pre-treatment, thermal treatment at 50 °C, and hydrolysis with alcalase for the procurement of collagen from salmon scales, and they reported a higher degree of hydrolysis and the higher inhibition of ABTS, DPPH, and ferric-reducing antioxidant power in comparison to other pre-treatments.
Another treatment consists of a pulsated electric field, consisting of a 10–50 kV/cm electric field that is executed as multiple short pulses (typically 1–5 μs) in frequencies of 0.2–0.4 MHz. The treated proteins expose hydrolysis sites that were previously inaccessible to peptidases [118], and the processing—carried out with subcritical water (it is an ecological technology)—employs the generated pressure to heat the water at a temperature between its atmospheric boiling point (100 °C and 0.1 MPa) and critical point (374 °C and 22.1 MPa); the yield and bioactivity of the hydrolysate of the treated proteins depend on the hydrolysis temperature and pressure [119]
Some authors such as Chian et al. (2019), Alvarez et al. (2016), Mikhaylin et al. (2017), and Ahmed et al. (2018) [118,120,121,122] have applied pre-treatments to obtain peptides, and these treatments have led to improved antioxidant and functional properties. For example, the use of ultrasound as a pre-treatment [123] and its simultaneous application during enzymatic hydrolysis [124] could modify protein structure and conformation, affecting the hydrogen bonds or hydrophilic interactions and the tertiary and quaternary structure thereof. In addition to the physical and mechanical effects, ultrasound can expose hydrolysis sites to a greater extent, resulting in an increase in susceptibility to the action of proteinases and causing an increase in the degree of hydrolysis. The adverse effects that may occur include damage to the conformation of the enzyme itself and therefore to its hydrolytic activity [115]. The advantages of microwave-assisted peptide synthesis over other conventional heating processes include its selectivity, allowing for digestion to be carried out in a short period of time [125,126] due to the selective heating of the intermolecular friction product of the alignment of ions and dipoles to the oscillating electric field of the microwaves and a subsequent dissipation of the energy via heating in the nucleus. This is also a clean process that leads to energy savings of between 25 and 75%. It allows for the control of parameters such as pressure, time and working temperature, and enables the obtainment of optimal results in shorter times. Experimental results have indicated that digestion temperature, reaction time, enzyme-to-substrate ratio, and digestion buffer are the primary factors required to optimize the method for successful hydrolysis [127].
Supercritical fluid extraction is a very useful technique as a method for the separation and pre-concentration of bioactive peptides. One advantage of using supercritical fluids is that they can penetrate a porous solid more easily, resulting in faster extraction. The dissolution of the supercritical fluid is controlled as a function of pressure and temperature. In supercritical extraction, fresh fluid is continuously passed through a sample. Supercritical fluids are easily recovered from the extract by depressurization. Supercritical extraction uses environmentally friendly solvents, and the supercritical fluid can be recycled or reused, minimizing the generation of waste.
However, disadvantages include that a supercritical fluid must have a series of properties in order to be used as a solvent in industry and especially in the food industry, e.g., high solvent capacity; selective, zero or low flammability; easy obtainment in high purity; low price; low or null toxicity; environment friendliness; non-corrosiveness; moderate critical conditions; and gaseous under ambient temperature and pressure conditions. In addition to obtaining food ingredients, ethanol, water and CO2, which is cheap and commercially available (even with high purity and declared Generally Recognized as Safe (GRAS) designation), must be used as modifiers. Due to the nature of the supercritical fluid extraction process, high pressures are required to carry out extraction. The method has a high cost due to investment for equipment and maintenance. Solvent compression requires elaborate recirculation measures to reduce energy costs. The use of modifiers can alter the polarity of the CO2, but they can remain in the extract, thus requiring a subsequent separation process [128,129]. Regarding concern for the environment, supercritical fluid extraction has shown excellent opportunities to achieve the key objectives of “green chemistry”, a term that refers to the design of chemical products and processes that reduce or eliminate the use or production of environmentally hazardous substances. This methodology meets several objectives such as the reductions in the use of solvents and lowering their potential harm as much as possible. The coupling of the aforementioned emerging technologies, together with enzymatic hydrolysis, have therefore proven to be a feasible method to produce bioactive peptides since they can be obtained in less time and at lower costs compared to the usual methods.
The effect of two pretreatments on the antioxidant activity in the antioxidant protein hydrolysates of quinoa has been evaluated using extraction with supercritical CO2 and ethanol as co-solvent; this type of pretreatment was compared to a conventional method of extraction with the ether of quinoa oil without the recovery of bioactive compounds [130]. Significant effects were found in the degree of hydrolysis (23.93%) and antioxidant activity (1181.64 μmol TE/g of protein) compared to the conventional method (24.33% and 1448.84 μmol TE/g protein, respectively). An economic evaluation was carried out using the SuperPro Designer 9.0® simulator to estimate the manufacturing cost (MC) considering laboratory scale lots (1.5 kg/lot) and industrial scale (2500 kg/lot). A total investment in the range of 400,000–500,000 dollars was calculated for the industrial scale for the conventional extraction method and 900,000–1,000,000 dollars when using the supercritical fluid extraction method. This finding is very interesting because it provides a global overview of these processes, with which it will be possible to propose changes to make the processes more efficient and economical.

3.4. Bioinformatic Tools in Obtaining Bioactive Peptides

For the identification and characterization of bioactive peptides, researchers have used bioinformatics techniques, e.g., computer-based (in silico) simulations have been applied to discover encrypted bioactive peptides in food proteins [17,131,132,133].
The in silico approach uses information accumulated in databases (Table 2) in an analysis prior to the experimental study as a rapid prediction tool to discover structure–function relationships, predict peptide sequences likely to exhibit specific activities, locate encrypted peptides in protein sources, anticipate the release of these fragments by the action of specific enzymes, and simulate or optimize processes of protein hydrolysis in order to predict the amino acids or peptides released when hydrolyzing the sequence of a protein with one or more enzymes and to identify peptides with biological activity and predict bioactive properties of peptides [17,134,135]. This approach also allows for the simultaneous evaluation of multiple food proteins and proteolytic enzymes to determine the frequency of occurrence of cryptic bioactive peptides in the primary structure of food proteins.
Protein sequences can be obtained from databases. In particular, the SwissProt and TrEMBL databases of the universal protein knowledge base (UniProtKB) (Table 2) can be used to calculate the release frequency of bioactive peptide fragments via a/N, where “a” is the number of bioactive peptides with specific activity released by the action of a certain enzyme and “N” is the total number of amino acid residues in the protein, and to simulate the proteolytic specificities of enzymes for generate in silico peptide profiles because in silico proteolysis such as BIOPEP, ExPASy PeptideCutter (http://web.expasy.org/peptide_cutter, accessed on 28 January 2022) and PoPS (http://pops.csse.monash.edu.au, accessed on 28 January 2022) programs include “enzyme action” tools. These tools include databases of peptide sequences with various bioactivities and programs for the simulation and prediction of peptide bioactivities [135].
Table 2 shows several peptide databases, some of which are dedicated to food peptides. These bases provide the collection of peptides with diverse bioactivities and sequences with different biofunctions.
The peptides resulting from in silico proteolysis can then be matched with bioactive peptides in databases for predetermined bioactivities. This approach has recently been used to study the distribution of antimicrobial peptides, antioxidants, inhibitors of disease-related enzymes within the primary structure of typical food proteins [132,158], and precursors of bioactive peptides [133], as well as to obtain sustainable proteins.
Barrero et al. (2021) [159] carried out a quantitative comparison of the bioactive peptides obtained from the caseins and seroproteins present in the milk of bovines (Bos taurus), sheep (Ovis aries), goats (Capra hircus) and buffalo (Bubalus bubalis) from in silico digestion processes catalyzed by proteases present in the human digestive system: pepsin (EC 3.4.23.1), trypsin (EC 3.4.21.4) and chymotrypsin (EC 3.4.21.1). The characterization of bioactive peptides and in silico digestion was performed using BIOPEP-UMW and quantitative evaluation from the calculation of release frequencies. The results showed 9 stimulant peptides for the four species and a slightly higher mean frequency for the buffalo with a value of 0.096, 7 renin inhibitor peptides with a mean frequency of 0.0074 for cattle and buffalo, 68 DPP4 inhibitor peptides with a frequency mean of 0.0716 for goats and sheep, 10 antioxidant peptides with a mean frequency of 0.0105 for cattle, 41 ACE inhibitor peptides with a mean frequency of 0.0432 for goats and sheep, and 8 DPP3 inhibitory peptides for the cattle, goats, and buffalo—the last with a slightly higher average frequency with a value of 0.0085. One peptide with hypocholesterolemic action and one peptide inhibitor of CaMPDE were obtained from the 𝛽-lactoglobulin of the four species.
Panjaitan et al. (2018) [160] attempted to match sequences identified in a tryptic hydrolysate of giant grouper (Epinephelus lancetolatus) roe proteins with bioactive peptide sequences listed in the BIOPEP-UWM database. Using this research strategy, it was reported that peptides found in giant grouper roe protein hydrolysates mainly showed ACE inhibitors, DPP-IV, and antioxidant bioactivities. Once in silico studies have been carried out, it is expected that the best combinations of protein proteases will be followed up and selected in order to experimentally validate the production and bioactivity of the peptides. In this study, the results of BIOPEP-UWM’s in silico and in vitro hydrolysis showed similar sequences to vitellogenin.
In silico studies applied to peptides can reduce the time required to detect bioactive peptides present in various protein sources using various proteases and can lead to the discovery of new and sustainable precursors of known bioactive peptides. However, it is not guaranteed that peptides in silico can be experimentally reproduced due to the complex nature of protease–protein interactions. Another disadvantage is that the currently used bioinformatic approach does not consider other peptide sequences within “in silico hydrolysates” that do not currently exist in bioactive peptide databases. Furthermore, the information that is available in databases often involves well-characterized and-purified proteolytic enzymes compared to enzymes used commercially for food processes that are less substrate-specific and of variable purity [161]. Furthermore, currently existing databases are often not regularly updated and can become unreliable, especially with information on new bioactive peptides constantly accumulating.
With the vast amount of information on the preclinical bioactivities of food-derived peptides, future efforts should consider exploring opportunities to promote the use of these peptides as functional ingredients to develop products with targeted health benefits, which could emphasize improving the taste, gastric stability, and bioavailability of peptides; providing clinical evidence to support their health effects; optimizing their inclusion in food products to limit undesirable reactions; and sourcing sustainable protein raw materials for their production in an attempt to reduce the heavy dependence on primary human food.

4. Health Benefits

Some studies have shown that bioactive peptides have beneficial effects on health, although the relationship between chemical structure and biological activity has not been predicted to date since the activity of a peptide depends on its structure, the composition of its amino acids, the type of terminal amino acid, the length of the peptide chain, the charge of the amino acids that make up the peptide, and the hydrophobic and hydrophilic characteristics of the amino acid chain, among others [21,133].

4.1. Antioxidant Activity

The biological activity of peptides has been widely studied, since, in addition to their antioxidant activity, they have nutritional and functional characteristics.
Antioxidant peptides are purified and linked amino acids from hydrolysates. They contain 5–16 amino acid residues, and those that come from food are considered safer because they are healthy compounds with low molecular weight, low cost, high activity, and easy absorption. They have advantages over enzymatic antioxidants because they have a simpler structure, greater stability, and do not generate dangerous immunological reactions [162,163,164].
Antioxidant peptides can be used as hydrolysates of the precursor proteins or as antioxidant peptides. A hydrolysate is a mixture mainly composed of peptides and amino acids that are produced through the hydrolysis of proteins by treatment with enzymes, acids, alkalis or fermentation. There is a classification by degree of hydrolysis, and this determines applications. For instance, hydrolysates with a low degree of hydrolysis and better functional characteristics [165] and hydrolysates with various degrees of hydrolysis can be used either as emulsifiers (such as in mayonnaise, minced meat, sausage, or ice cream) [166,167,168], as flavorings and protein supplements or in medical diets [169,170]. Furthermore, hydrolysates with a high degree of hydrolysis are used primarily as nutritional supplements and in special medical diets.
The 20 amino acids present in proteins can react with free radicals; the most reactive are those that include the sulfur amino acids M and C; the aromatics W, Y, and F; and an imidazole ring such as H. According to Girgih et al. (2015) [171], the last group comprises amino acids that have a high antioxidant capacities due to the imidazole rings that they use as a proton donors. Histidine and the indole ring of tryptophan, as well as lysine and valine, can scavenge hydroxyl radicals by acting as electron donors that generate a hydrophobic microenvironment in the molecule, which favors the antioxidant capacity of the peptide [27,123].
Various studies have been carried out to investigate the antioxidant properties of hydrolysates or bioactive peptides from plants and animals. Table 3 presents a list of various peptide fractions with antioxidant activity.

Mechanisms of Antioxidant Peptides

Several studies have shown that antioxidant peptides can act as (1) chelators of a metal, (2) radical inhibitors, or (3) physical shielding [103,191,192,193]. In Figure 2, these mechanisms are shown.
Mechanism 1, as chelators of a metal: These peptides inhibit the production of free radicals by stabilizing metallic pro-oxidants through sequestering agents. The carboxyl and amino groups of the side chains have a chelating function of metal ions due to their ability to dissociate and be proton donors. It has also been reported that chelating peptides are rich in histidine and prevent oxidative activity by chelating metal ions. Copper-chelating peptides are rich in histidine and prevent copper’s oxidative activity by chelating metal ions. The imidazole ring of this residue is directly involved in binding to copper. These peptides have also been found to be rich in arginine. Although this amino acid lacks chelating properties, it may favor the binding of the peptide to metal ions. Therefore, copper chelating peptides may be useful by preventing not only copper oxidative activity that can damage cells in the luminal space of the stomach but also the copper-induced oxidation of LDL if it reaches the bloodstream. They can also be useful in organs such as the brain, where the oxidative process is involved in the development of certain diseases [103,194,195]. R, a basic amino acid, confers peptide charge and hydrogen-bonding interactions, which are essential properties for combination of with the abundant anionic component of the bacterial membrane, and it appears that W residues play the role of natural aromatic activators of R-rich antimicrobial peptides via π-ion pair interactions, thus promoting enhanced peptide–membrane interactions [196].
Mechanism 2, as radical inhibitors: Peptides stabilize radicals by donating electrons, maintaining their own stability via the resonance of their structure. Some reported peptides are those that contain aromatic amino acids such as Y, H, W, and F and tripeptides with C that contain SH groups and are radical scavengers. In an oxidizing environment where unsaturated lipids are present, active peptides are more inclined to donate electrons than lipids and form more stable radicals or non-reactive polymers through radical–radical condensation [194,197,198].
Mechanism 3, physical shielding: Peptides inhibit lipid peroxidation by acting as a physical barrier or membrane. They are surfactant components and can split at the oil–water interface, forming a thick membrane or coating to avoid the direct contact of lipids and radicals with other oxidizing compounds [199].
The amino acid sequence and the presence of certain amino acid residues, such as histidine, tyrosine, tryptophan, methionine, cysteine, and proline, have also been shown to be determining factors in the antioxidant capacity of peptides [199]. The relationship between the presence of certain amino acids and their respective antioxidant activity is shown in Table 4.

4.2. Antimicrobial Activity

Antimicrobial peptides (AMPs) are molecules comprising 10–55 amino acids; they can be neutral, aniconic, or (mainly) cationic, and they have important innate antibacterial and antifungal properties. Almost 50% are hydrophobic and have a molecular weight of less than 10 kDa; they tend to be heat-stable due to the presence of fewer amino acids and can be generated in vitro via enzymatic hydrolysis [199,200,201,202,203].
It has been reported that the main mechanism of action of AMPs against microorganisms is the rupture of membranes, is which an electrostatic union with the membrane that carries negative charge is initiated through the positive charge, therefore leading to the leakage of metabolites, membrane depolarization, interruption of respiratory processes and finally, and cell death [204].
The mechanism of action of AMPs is divided into three stages:
  • Stage 1. Attraction for the bacterial cell wall
Antimicrobial peptides are attracted to bacterial surfaces thanks to their amphipathic structure, the positive charge at physiological pH through electrostatic bonding between anionic or cationic peptides, and structures on the bacterial surface. It is likely that cationic AMPs are first attracted to the net negative charges that exist on the outer envelope of Gram-negative bacteria (e.g., anionic phospholipids and lipopolysaccharides (LPS)) and to teichoic acids on the surface of Gram-positive bacteria, acting as detergents that break down the bacterial membrane [205,206].
  • Stage 2. Union to the cell membrane
Once on the microbial surface, peptides must pass through capsular polysaccharides before they can interact with the outer membrane, which contains LPS in Gram-negative bacteria and transverse capsular polysaccharides, teichoic acids, and lipoteichoic acids [207].
  • Stage 3. Peptide insertion and cell membrane permeabilization
Peptides interact with the cytoplasmic membrane in lipid Gram-positive bacteria and then gain access to the cytoplasmic membrane and can interact with lipid bilayers [207]. This interaction allows for the accumulation of intramembrane peptides until reaching a threshold at which collapse occurs and the membrane ruptures.
The potential mechanisms of action of antimicrobial peptides include:
  • Bacterial membrane permeabilization: The mixed hydrophobic and cationic composition of antimicrobial peptides makes them highly suitable for interacting with and disrupting microbial cytoplasmic membranes that present anionic, lipid-rich surfaces, such as phosphatidylglycerol and cardiolipin. The fact that all Gram-negative and Gram-positive bacteria show this type of negatively charged lipids explains the lack of specificity of most antimicrobial peptides, promotes the attraction between antimicrobial peptides and bacterial membranes, and prevents their binding to most host cell membranes. The selective toxicity of antimicrobial peptides is based on differences in the membrane potential of mammalian microbes and cells. Microbes tend to have a significantly large charge difference across their membranes compared to mammalian cells, favoring cationic defensins to selectively attack microbes [208,209,210,211].
  • Enzymatic attack on bacterial wall structures: Several epithelial antimicrobial peptides kill bacteria through enzymatic attacks on key cell wall structures. Lysosomes are effective against Gram-positive bacteria, where peptidoglycan is more accessible, than against Gram-negative organisms, where peptidoglycan is protected by the outer membrane [207]. In addition to directly killing bacteria, the enzymatic activity of lysozymes can regulate the innate immune responses to certain microorganisms. Secretory phospholipase A2 (sPLA2) is an example of an antimicrobial peptide that kills bacteria through an enzymatic mechanism. Bacterial membranes, rich in phosphatidylglycerol and phosphatidylethanolamine, are the key targets of sPLA2, but the enzyme can break down other phosphotriglyceride substrates. The sPLA2 peptide penetrates the bacterial cell wall to access the membrane, where it hydrolyzes phospholipids and thus compromises the integrity of the bacterial membrane. This enzyme is bactericidal, with preferential activity against Gram-positive bacteria [212].
  • Interference at the intracellular level: It seems likely that many AMPs can translocate across microbial membranes at concentrations that do not induce permeabilization. Once in the cytoplasm, they can attack DNA and chaperonins, alter the formation of the cytoplasmic membrane septum, inhibit cell wall synthesis, reduce nucleic acid synthesis, suppress protein synthesis, or inhibit enzyme activity [213,214].
  • Antimicrobial peptides that exploit multiple antimicrobial strategies: The peptide can affect microbes in ways other than cell destruction such as the destabilization of the membrane, the filamentation of bacterial cells due to the insertion of peptides in the membrane, the blocking of DNA replication, and the inhibition of membrane proteins involved in septum formation [214], lectin-like behavior, and antitoxin activity.
In addition to their antimicrobial capacity, peptides are also effective against fungal infections since they suppress the reproduction or growth of fungi. Antifungal peptides are differentiated by their mechanisms of action: the first group are amphipathic peptides—lytic peptides that break down membranes, are naturally abundant, and have hydrophilic and hydrophobic elements—that have two surfaces; one is positively charged and the other uncharged (neutral). Some penetrate cells and can alter the structure of a membrane without crossing the membrane simply by attaching to its surface. The second group of peptides can obstruct the synthesis of the cell wall or the biosynthesis of glucan or chitin, essential cellular components [215,216].
In medicinal applications, antimicrobial peptides are preferred over conventional bactericidal antibiotics because they kill bacteria faster and are not affected by antibiotic-resistance mechanisms [217,218,219]. Various studies have reported the presence of antimicrobial peptides in milk and breast secretions, as well as from other food sources such as derivatives of rice protein, soy protein, buckwheat, and pepper [87,220,221]. Additionally, peptides derived from casein have shown antimicrobial activity in in vitro models in the presence of various microorganisms such as Staphylococcus ssp., Streptococcus pyogenes, Diplococcus pneumoniae and Bacilus subtilis. The presence of casecidin, obtained from α-s1 casein, after being hydrolyzed by chymosin, inhibits the growth of these microorganisms at high concentrations [84].
Recent research [222,223] has shown that are peptides that exist in different organisms such as amphibians and marine animals could be a source of peptides with antimicrobial activity, e.g., the amphibian Rana cancrivora is the main source of carcinine [224]. On the other hand, some marine organisms present peptides with both antimicrobial and antibacterial activity, among which are AMPs that play a key role in the strategy against infections [225]. Bacillus subtilis-fermented fish meat protein hydrolysates prepared from sardinella, zebra blenny, goby, and ray fish showed antibacterial activity, and sardinella hydrolysates were the most effective against Gram-positive bacteria [226]. AMPs are small oligopeptides generally containing 10–100 amino acids, with a net positive charge and an amphipathic structure; they are the most important components of the antibacterial defense of organisms at almost all levels of life: bacteria, fungi, plants, amphibians, insects, and mammals. Examples from various sources are shown in Table 5.

5. Bioavailability of Bioactive Peptides

A protein that is ingested through the diet is digested and absorbed in the gastrointestinal tract; it is denatured by gastric acid and hydrolyzed into small peptides and amino acids by gastric and pancreatic proteases. Some amino acids are used in the cell by the enterocytes themselves as a source of energy, and others undergo metabolic transformations before passing into the blood during cell turnover (Figure 3) [243]. The final step in the digestion of dietary proteins occurs on the surface of enterocytes via the action of intestinal brush border cell proteases [244]. Once the amino acids and small peptides that have managed to cross the intestinal barrier, enter the blood, and reach the liver through the portal circulation, they are captured and used by the liver; then, some enter the systemic circulation and are used by the tissue’s 10 peripherals. The metabolic life of amino acids is complex and starts with their use as energy or gluconeogenic substrates and ends with the synthesis of new proteins and peptides [243].
The potential beneficial effects of biopeptides depend on their ability to reach the organs where they are going to perform their function intact. Therefore, it is important to consider the in vivo differences of bioactive peptides, since peptides must be able to overcome barriers and actively reach their target because they are subject to degradation and modification in the intestine, vascular system, and hepatic system. Although peptides are rapidly metabolized to their constituent amino acids, some studies have shown that several peptides are resistant to these physiological processes and can reach the circulation [246].
Gardner (1998) [247] indicated that a small portion of bioactive peptides can pass the gut barrier, and, though too small to be considered nutritionally important, this proportion can have biological effects at the tissue level. However, peptides and proteins can escape digestion and be absorbed intact. One reported mechanism for intact peptide uptake is the use of paracellular water-soluble peptides. Hydrophobic peptides are absorbed by diffusion through tight junctions between cells by a passive diffusion process independent of energy [248]. Some small peptides, resistant to hydrolysis, leave the enterocyte and enter the circulatory system through a peptide transporter located in the intestinal basolateral membrane; in endocytosis [248], large polar peptides bind to cells and are absorbed into the cell through vesiculation and the lymphatic system [249]. Highly lipophilic peptides, too large to be absorbed into circulation, are absorbed from the interstitial space into the intestinal lymphatic system.
Molecular size, structural properties, and hydrophobicity also affect the major peptide transport pathway [250]. Research indicates that peptides with 2–6 amino acids are more easily absorbed than proteins and free amino acids [251].
Roberts et al. (1999) [252] reported that small (di- and tripeptides) and large (10–51 amino acids) peptide chains can cross the intestinal barrier intact and exhibit their biological functions at the tissue level. However, as the molecular weight of peptides increases, their ability to pass through the intestinal barrier decreases.
Therefore, it can be deduced that due to incomplete bioavailability of a peptide after oral ingestion, a peptide with pronounced bioactivity in vitro may exert little or no activity in vivo. However, other routes of derivation can increase the possibility of absorption of peptides, diminishing the problem [56].
Vilcacundo et al. (2017) [253] extracted a protein concentrate from quinoa that was digested under in vitro gastrointestinal conditions. Quinoa proteins were almost completely hydrolyzed by pepsin at pH 1.2, 2.0 and 3.2. At high pH, only partial hydrolysis was observed. During the duodenal phase, no intact proteins were seen, indicating their susceptibility to simulated digestive conditions in vitro. The zebrafish larval model was used to assess the in vivo the ability of gastrointestinal digests to inhibit lipid peroxidation. Gastric digestion at pH 1.2 showed the highest percentage of the inhibition of lipid peroxidation (75.15%). Lipid peroxidation increased after the duodenal phase. These peptides could be further hydrolyzed by pancreatin, releasing new fragments with greater antioxidant activity. Thus, the hydrolysate obtained at the end of the digestive process showed a higher inhibitory capacity with an inhibition percentage of 82.10%, similar to that of the positive control butylated hydroxytoluene used in the test (87.13%). The results of this study demonstrated, for the first time, that quinoa protein hydrolysates obtained during in vitro simulated gastrointestinal digestion were capable of inhibiting lipid peroxidation in the in vivo zebrafish model.
Carrillo et al. (2016) [254] identified five peptides (VAWRNRCKGTD, IRGCRL, WIRGCRL, AWIRGCRL, and WRNRCKGTD) from hen egg-white lysozyme with in vitro antioxidant activity (1970, 3123, 2743, 2393 and 0.013 µmol Trolox/µmol peptide, respectively), being the last one with less activity. They determined the inhibition of lipid peroxidation in an in vivo model of zebrafish larvae through cell damage. This assay confirmed that antioxidant peptides of hen egg-white lysozyme were not toxic for the zebrafish larvae, which were normal after 24 h of assay. When the zebrafish larvae were examined, no morphological abnormalities, such as crooked bodies, spinal deformities, or any significant effects in the growth of the body, were observed. The lysozyme peptides were shown to efficiently inhibit the lipid peroxidation in zebrafish larvae. The AWIRGCRL peptide demonstrated 63.2% thiobarbituric acid reactive substance (TBARS) inhibition. A higher antioxidant activity was shown the of WIRGCRL, AWIRGCRL and WRNRCKGTD peptides. In the case of the TBARS inhibition, peptide size probably contribues to the increased values of activity in the IRGCRL, WIRGCRL, AWIRGCRL, and WRNRCKGTD peptides.
In vivo, the bioavailability of a bioactive peptide depends on its resistance to hydrolysis by peptidases present in both the intestinal tract and the serum of the bloodstream, as well as the ability of the generated peptides to be absorbed through barriers such as the intestinal epithelium. The in vitro method tends not to consider some aspects such as the absorption, distribution properties, metabolism, and excretion of peptides, which can result in low activity and bioavailability.
The biological effects of peptides depend on their bioavailability, which is determined by their resistance to gastrointestinal digestion [255]. However, to ensure the biological effects of peptides, it is important to carry out in vitro and in vivo studies that confirm their stability, absorption capacity, and mechanism of action, which increase and diversify their nutraceutical and pharmaceutical applications.
According to what is stated in the present review, bioactive peptides present in the diet, could improve the body’s ability to regulate oxidative stress, which cannot be modified by the intervention of endogenous antioxidant defenses and thereby help to reduce the incidence of some degenerative diseases.

6. Conclusions

The objective of this review was to provide an overview of foods and by-products as sources rich in bioactive peptides such as antioxidants and antimicrobials, as well as their potential in reducing oxidative stress through exogenous factors or the consumption of antibiotics for the improvement of the health of consumers.
The production of these peptides can occur through enzymatic processes or via fermentation. New extraction technologies with supercritical fluids, microwaves, and pulse electric field are being used as pretreatments combined with conventional methods to obtain higher yields and lower costs of these bioactive compounds, with significant effects on the degree of hydrolysis, antioxidant activity, and antimicrobial activity. One of the findings observed in this review was from economic study, in which it was observed that obtaining these compounds in combination with emerging technologies enabled reductions in costs at the industrial level compared to conventional methods. Furthermore, the peptides resulting from in silico proteolysis can then be matched with bioactive peptides in databases for predetermined bioactivities.
In silico studies of peptides have been able to reduce the time required to detect bioactive peptides present in various protein sources using various proteases and to lead to the discovery of new and sustainable precursors of known bioactive peptides.
With the vast amount of available information on the preclinical bioactivities of food-derived peptides, future studies should consider exploring opportunities to promote the use of these peptides as functional ingredients to develop products with targeted health benefits, which could emphasize improving the taste, gastric stability, and bioavailability of peptides; providing clinical evidence to support their health effects; optimizing their inclusion in food products to limit undesirable reactions; and sourcing sustainable protein raw materials for their production in an attempt to reduce the heavy dependence on primary human food. However, the constant update of these approaches is necessary.
Regarding in vivo and in vitro studies, peptides obtained from some cereals and enzymes derived from by-products have shown the ability to inhibit lipoperoxidation.
There have been numerous studies on the antimicrobial peptides that can be used as protectors against oxidative stress due to the consumption of antibiotics, and deeper in vivo studies on the exact mechanism of action are recommended. Additionally, antioxidant and antimicrobial peptides could be used to produce functional foods. Therefore, it is also important to study the interaction of peptides with other food constituents, as well as the effects of incorporating these peptides and the processing conditions on the bioactivity of peptides in food matrices.

Author Contributions

Investigation and conceptualization, O.D.-G. and G.L.-G.; table and figure preparation and writing—original draft preparation, O.D.-G., D.A.-C. and G.L.-G.; review and editing, G.L.-G., D.A.-C. and L.M.G.O.; validation, O.D.-G., L.M.G.O., G.L.-G. and D.A.-C.; supervision, O.D.-G. and L.M.G.O.; project administration, O.D.-G. and G.L.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Guadalupe López-García would like to thank CONACyT for the graduate-degree scholarship granted to her.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Delgado, O.L.; Betanzos, C.G.; Sumaya, M.M.T. Importancia de los antioxidantes dietarios en la disminución del estrés oxidativo. Investig. Cienc. 2010, 8, 10–15. [Google Scholar]
  2. Estévez, M. Protein carbonyls in meat systems: A review. Meat Sci. 2011, 89, 259–279. [Google Scholar] [CrossRef] [PubMed]
  3. Huanqui, G.C. Oxidantes-antioxidantes en reumatología. Rev. Peru Reum. 1997, 3, 35–40. [Google Scholar] [CrossRef] [Green Version]
  4. Dunkan, S.; Farewell, A.; Ballesteros Radman, M.; Nystrom, T. Protein oxidation in response to increased transcriptional or translational errors. Proc. Natl. Acad. Sci. USA 2000, 97, 5746–5749. [Google Scholar] [CrossRef] [Green Version]
  5. Lieber, M.R.; Karanjawala, Z.E. Ageing, repetitive genomes and DNA damage. Nature 2004, 5, 69–75. [Google Scholar] [CrossRef]
  6. Pérez-Torres, I.; Castrejón-Téllez, V.; Soto, M.E.; Rubio-Ruiz, M.E.; Manzano-Pech, L.; Guarner-Lans, V. Oxidative Stress, Plant Natural Antioxidants, and Obesity. Int. J. Mol. Sci. 2021, 22, 1786. [Google Scholar] [CrossRef]
  7. Soledad-Bustos, P. Efecto de Flavonoides Frente al Estrés Oxidativo Inducido por Antibióticos. Ph.D. Thesis, Departamento de Farmacia–Facultad de Ciencias Químicas, Universidad Nacional de Córdoba, Córdoba, Argentina, 2017. [Google Scholar]
  8. Venereo, G.J.R. Daño Oxidativo, Radicales Libres y Antioxidantes. Rev. Cuba. Med. Mil. 2002, 31, 126–133. [Google Scholar]
  9. Monfared, S.S.M.S.; Vahidi, H.; Abdolghaffari, A.H.; Nikfar, S.; Abdollahi, M. Antioxidant therapy in the management of acute, chronic and post-ERCP pancreatitis: A systematic review. World J Gastroenterol. 2009, 15, 4481–4490. [Google Scholar] [CrossRef]
  10. Albuquerque, R.V.; Malcher, N.S.; Amado, L.L.; Coleman, M.D.; Dos Santos, D.C.; Borges, R.S.; Valente, S.A.; Valente, V.C.; Monteiro, M.C. In vitro protective effect and antioxidant mechanism of resveratrol induced by dapsone hydroxylamine in human cells. PLoS ONE 2015, 10, e0134768. [Google Scholar] [CrossRef] [Green Version]
  11. Atli, O.; Demir-Ozkay, U.; Ilgin, S.; Aydin, T.H.; Akbulut, E.N.; Sener, E. Evidence for neurotoxicity associated with amoxicillin in juvenile rats. Hum. Exp. Toxicol. 2016, 35, 866–876. [Google Scholar] [CrossRef]
  12. Correa Salde, V.A. Efecto de Fármacos a Nivel del Equilibrio de Especies Reactivas: Consecuencias Moleculares y Celulares. Ph.D. Thesis, Universidad Nacional de Córdoba, Córdoba, Argentina, 2011. [Google Scholar]
  13. Ilgin, S.; Can, O.D.; Atli, O.; Ucel, U.I.; Sener, E.; Guven, I. Ciprofloxacininduced neurotoxicity: Evaluation of possible underlying mechanisms. Toxicol. Mech. Methods 2015, 25, 374–381. [Google Scholar] [CrossRef] [PubMed]
  14. Keeney, J.T.; Miriyala, S.; Noel, T.; Moscow, J.A.; St Clair, D.K.; Butterfield, D.A. Superoxide induces protein oxidation in plasma and TNF-α elevation in macrophage culture: Insights into mechanisms of neurotoxicity following doxorubicin chemotherapy. Cancer Lett. 2015, 367, 157–161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Qin, P.; Liu, R. Oxidative stress response of two fluoroquinolones with catalase and erythrocytes: A combined molecular and cellular study. J. Hazard. Mater. 2013, 252–253, 321–329. [Google Scholar] [CrossRef] [PubMed]
  16. Sankar, M.; Rajkumar, J.; Devi, J. Hepatoprotective activity of hepatoplus on isonaizid and rifampicin induced hepatotoxicity in rats. Pak. J. Pharm. Sci. 2015, 28, 983–990. [Google Scholar] [PubMed]
  17. Carrasco-Castilla, J.; Hernández-Álvarez, A.J.; Jiménez-Martínez, C.; Gutiérrez-López, G.F.; Dávila-Ortiz, G. Use of proteomics and peptidomics methods in food bioactive peptide science and engineering. Food Eng. Rev. 2012, 4, 224–243. [Google Scholar] [CrossRef]
  18. De Castro, R.J.S.; Sato, H.H. Biologically active peptides: Processes for the generation, purification and identification and applications as natural additives in the food and pharmaceutical industries. Food Res. Int. 2015, 74, 185–198. [Google Scholar] [CrossRef]
  19. FitzGerald, R.J.; Murray, B.A.; Walsh, D.J. Hypotensive peptides from milk proteins. J. Nutr. 2004, 134, 980S–988S. [Google Scholar] [CrossRef] [Green Version]
  20. Aluko, R.E. Determination of nutritional and bioactive properties of peptides in enzymatic pea, chickpea and mung bean protein hydrolysates. J. AOAC Int. 2008, 91, 947–956. [Google Scholar] [CrossRef] [Green Version]
  21. Mulero, C.J.; Zafrilla, R.P.; Martínez-Cacha, M.A.; Leal, H.M.; Abellan, A.J. Péptidos bioactivos. Clin. Investig. Arterioscler. 2011, 23, 219–227. [Google Scholar] [CrossRef]
  22. Wu, H.; Liu, Z.; Zhao, Y.; Mingyong, Z. Enzymatic preparation and characterization of iron-chelating peptides from anchovy (EngrauLys japonicus) muscle protein. Food Res. Int. 2012, 48, 435–441. [Google Scholar] [CrossRef]
  23. Chen, D.; Liu, Z.; Huang, W.; Zhao, Y.; Dong, S.; Zeng, M. Purification and characterization of a zinc-binding peptide from oyster protein hydrolysate. J. Funct. Foods 2013, 5, 689–697. [Google Scholar] [CrossRef]
  24. Montero, B.M. Protein hydrolyzed from byproducts of the fishery industry: Obtaining and functionality. Agron. Mesoam. 2021, 32, 681–699. [Google Scholar] [CrossRef]
  25. Sun, J.; He, H.; Xie, B.J. Novel Antioxidant Peptides from Fermented Mushroom Ganoderma lucidum. J. Agric. Food Chem. 2004, 52, 6646–6652. [Google Scholar] [CrossRef] [PubMed]
  26. Deraz, S.F. Protein hydrolysate from visceral waste proteins of Bolti fish (Tilapia nilotica): Chemical and nutritional variations as affected by processing pHs and time of hydrolysis. J. Aquat. Food Prod. Technol. 2014, 24, 614–631. [Google Scholar] [CrossRef]
  27. Bougatef, A.; Nedjar-Arroume, N.; Manni, L.; Ravallec, R.; Barkia, A.; Guillochon, D.; Nasri, M. Purification and identification of novel antioxidant peptides from enzymatic hydrolysates of sardinelle (Sardinella aurita) by-products proteins. Food Chem. 2010, 118, 559–565. [Google Scholar] [CrossRef]
  28. De Macedo, M.; Segura, R.; Piñero Bonilla, J.; Coello, N. Obtención de un hidrolizado proteico por fermentación sumergida de plumas utilizando Bacillus spp. Rev. Cient. Fac. Cienc. FCV-LUZ 2002, 12, 214–220. [Google Scholar]
  29. Sun, Q.; Zhang, B.; Yan, Q.J.; Jiang, Z.Q. Comparative analysis on the distribution of protease activities among fruits and vegetable resources. Food Chem. 2016, 213, 708–713. [Google Scholar] [CrossRef]
  30. Borawska, J.; Darewicz, M.; Vegarud, G.E.; Iwaniak, A.; Minkiewicz, P. Ex vivo digestion of carp muscle tissue–ACE inhibitory and antioxidant activities of the obtained hydrolysates. Food Funct. 2015, 6, 211–221. [Google Scholar] [CrossRef]
  31. Arihara, K.; Ohata, M. Bioactive Compounds in Meat. In Meat Biotechnology; Toldrá, F., Ed.; Springer: New York, NY, USA, 2008; pp. 231–249. [Google Scholar] [CrossRef]
  32. Escudero, E.; Toldrá, F.; Sentandreu, M.; Nishimura, H.; Arihara, K. Antihypertensive activity of peptides identified in the in vitro gastrointestinal digestof pork meat. Meat Sci. 2012, 91, 382–384. [Google Scholar] [CrossRef]
  33. Marshal, K. Therapeutic applications of whey protein. Altern. Med. Rev. 2004, 9, 136–156. [Google Scholar]
  34. Jiang, H.; Tong, T.; Sun, J.; Xu, Y.; Zhao, Z.; Liao, D. Purification, and characterization of antioxidative peptides from round scad (Decapterus maruadsi) muscle protein hydrolysate. Food Chem. 2014, 154, 158–163. [Google Scholar] [CrossRef]
  35. Ko, J.Y.; Lee, J.H.; Samarakoon, K.; Kim, J.S.; Jeon, Y.J. Purification, and determination of two novel antioxidant peptides from flounder fish (Paralichthis olivaceus) using digestive proteases. Food Chem. Toxicol. 2013, 52, 113–120. [Google Scholar] [CrossRef]
  36. Fujita, H.; Yokoyama, K.; Yasumoto, R.; Yoshikawa, M. Antihypertensive Effect of thermolysin digest of dried bonito in spontaneously hypertensive rat. Clin. Exp. Pharmacol. Physiol. 1995, 22, S304–S305. [Google Scholar] [CrossRef]
  37. Capriotti, A.; Caruso, G.; Cavaliere, C.; Samperi, R.; Chiozzi, S.; Laganà, A. Identification of potential bioactive peptides generated by simulated gastrointestinal digestion of soybean seeds and soy milk proteins. J. Food Compos. Anal. 2015, 44, 205–213. [Google Scholar] [CrossRef]
  38. Salas, C.E.; Badillo-Corona, J.A.; Ramírez-Sotelo, G.; Oliver-Salvador, C. Biologically Active and Antimicrobial Peptides from Plants. Biomed Res. Int. 2015, 2015, 102129. [Google Scholar] [CrossRef] [Green Version]
  39. Hammami, R.; Hamida, B.J.; Vergoten, G.; FLyss, I. PhytAMP: A database dedicated to antimicrobial plant peptides. Nucleic Acids Res. 2009, 37, 963–968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Hartmann, R.; Meisel, H. Food-derived peptides with biological activity: From research to food aplications. Curr. Opin. Biotechnol. 2007, 18, 163–169. [Google Scholar] [CrossRef] [PubMed]
  41. Yaklich, R.W. beta-Conglycinin and glycinin in high-protein soybean sedes. J. Agric. Food Chem. 2001, 49, 729–735. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, J.; Zhang, H.; Wang, L.; Guo, X.; Wang, X.; Yao, H. Isolation and identification of antioxidative peptides from rice endosperm protein enzymatic hydrolysate by consecutive chromatography. Food Chem. 2010, 119, 226–234. [Google Scholar] [CrossRef]
  43. Przybylski, R.; Firdaous, L.; Châtaigné, G.; Dhulster, P.; Nedjar, N. Production of an antimicrobial peptide derived from slaughterhouse by-product and its potential application on meat as preservative. Food Chem. 2016, 211, 306–313. [Google Scholar] [CrossRef]
  44. Tavares, L.S.; Rettore, J.V.; Freitas, R.M.; Porto, W.F.; Do Nascimento Duque, A.P.; De Lacorte Singulani, J.; Silva, O.N.; De Lima Detoni, M.; Vasconcelos, E.G.; Dias, S.C. Antimicrobial activity of recombinant Pg-AMP1, a glycine-rich peptide from guava seeds. Peptides 2012, 37, 294–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Karkouch, I.O.; Tabbene, D.; Gharbi, M.A.; Ben Mlouka, S.; Elkahoui, C.; Rihouey, L.; Coquet, P.; Cosette Jouenne, T.; Limam, F. Antioxidant, antityrosinase and antibiofilm activities of synthesized peptides derived from vicia faba protein hydrolysate: A powerful agents in cosmetic application. Ind. Crop. Prod. 2017, 109, 310–319. [Google Scholar] [CrossRef]
  46. Pu, C.; Tang, W. The antibacterial and antibiofilm efficacies of a liposomal peptide originating from rice bran protein against: Listeria monocytogenes. Food Funct. 2017, 8, 4159–4169. [Google Scholar] [CrossRef]
  47. Chi, C.-F.; Hu, F.-Y.; Wang, B.; Ren, X.-J.; Deng, S.-G.; Wu, C.-W. Purification and characterization of three antioxidant peptides from protein hydrolyzate of croceine croaker (Pseudosciaena crocea) muscle. Food Chem. 2015, 168, 662–667. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, Q.; Li, W.; He, Y.; Ren, D.; Kow, F.; Song, L.; Yu, X. Novel antioxidative peptides from the protein hydrolysate of oysters (Crassostrea talienwhanensis). Food Chem. 2014, 14, 991–996. [Google Scholar] [CrossRef]
  49. Ahn, C.-B.; Kim, J.-G.; Je, J.-Y. Purification and antioxidant properties of octapeptide from salmon byproduct protein hydrolysate by gastrointestinal digestion. Food Chem. 2014, 147, 78–83. [Google Scholar] [CrossRef]
  50. Wenyi, W.; Gonzalez, E. A New Frontierin Soy Bioactive Peptides that May Prevent Age-related, Chronic Diseases. Compr. Rev. Food Sci. Food Saf. 2005, 4, 63–73. [Google Scholar]
  51. Saiga, A.; Iwai, K.; Hayakawa, T.; Takahata, Y.; Kitamura, S.; Nishimura, T.; Morimatsu, F. Angiotensin I-converting enzyme-inhibitory peptides obtained from chicken collagen hydrolysate. J. Agric. Food Chem. 2008, 56, 9586–9591. [Google Scholar] [CrossRef]
  52. Ferreira, I.C.; Vaz, J.A.; Vasconcelos, M.H.; Martins, A. Compounds from wild mushrooms with antitumor potential. Anti-Cancer Agents Med. Chem. 2010, 10, 424–436. [Google Scholar] [CrossRef] [Green Version]
  53. Blagodatski, A.; Yatsunskaya, M.; Mikhailova, V.; Tiasto, V.; Kagansky, A.; Katanaev, V.L. Medicinal mushrooms as an attractive new source of natural compounds for future cancer therapy. Oncotarget 2018, 9, 29259–29274. [Google Scholar] [CrossRef] [Green Version]
  54. Zhou, R.; Liu, Z.K.; Zhang, Y.N.; Wong, J.H.; Ng, T.B.; Liu, F. Research Progress of Bioactive Proteins from the Edible and Medicinal Mushrooms. Curr. Protein Pept. Sci. 2019, 20, 196–219. [Google Scholar] [CrossRef] [PubMed]
  55. Chakrabarti, S.; Guha, S.; Majumder, K. Food-Derived Bioactive Peptides in Human Health: Challenges and Opportunities. Nutrients 2018, 10, 1738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Erdmann, K.; Cheung, B.W.Y.; Schroder, H. The possible roles of food-derived bioactive peptides in reducing the risk of cardiovascular disease. J. Nutr. Biochem. 2008, 19, 643–654. [Google Scholar] [CrossRef] [PubMed]
  57. Wang, X.; Yu, H.; Xing, R.; Li, P. Characterization, preparation, and purification of marine bioactive peptides. Biomed. Res. Int. 2017, 2017, 9746720. [Google Scholar] [CrossRef] [Green Version]
  58. Mygind, P.H.; Fischer, R.L.; Schnorr, K.M.; Hansen, M.T.; Sönksen, C.P.; Ludvigsen, S.; Raventós, D.; Buskov, S.; Christensen, B.; De Maria, L. Plectasin is a peptide antibiotic with therapeutic potential from a saprophytic fungus. Nature 2005, 437, 975–980. [Google Scholar] [CrossRef]
  59. Alemán, A.; Gómez-Guillén, M.C.; Montero, P. Identification of ace-inhibitory peptides from squid skin collagen after in vitro gastrointestinal digestión. Food Res. Int. 2013, 54, 790–795. [Google Scholar] [CrossRef]
  60. Je, J.Y.; Qian, Z.J.; Byun, H.G.; Kim, S.K. Purification and characterization of an antioxidant peptide obtained from tuna backbone protein by enzymatic hydrolysis. Process Biochem. 2007, 42, 840–846. [Google Scholar] [CrossRef]
  61. Chang, C.-Y.; Wu, K.-C.; Chiang, S.H. Antioxidant properties and protein compositions of porcine haemoglobin hydrolysates. Food Chem. 2007, 100, 1537–1543. [Google Scholar] [CrossRef]
  62. Adler-Nissen, J.; Olsen, H.S. The influence of peptide chain length on taste and functional properties of enzymatically modified soy protein. In Functionality and Protein Structure; Pour-El, A., Ed.; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 1979; Volume 92, pp. 125–146. [Google Scholar] [CrossRef]
  63. Kong, X.; Zhou, H.; Qian, H. Enzymatic preparation and functional properties of wheat gluten hydrolysates. Food Chem. 2007, 101, 615–620. [Google Scholar] [CrossRef]
  64. Bhaskar, N.; Benila, T.; Radha, C.; Lalitha, R.G. Optimization of enzymatic hydrolysis of visceral waste proteins of Catla (Catla catla) for preparing protein hydrolysate using a commercial protease. Bioresour. Technol. 2008, 99, 335–343. [Google Scholar] [CrossRef]
  65. Kamnerdpetch, C.; Weiss, M.; Kasper, C.; Scheper, T. An improvement of potato pulp protein hydrolyzation process by the combination of protease enzyme systems. Enzym. Microb. Technol. 2007, 40, 508–514. [Google Scholar] [CrossRef]
  66. Pihlanto-Leppälä, A. Bioactive Peptides derived from bovine whey proteins: Opioid and ace-inhibitory peptides. Trends Food Sci. Technol. 2001, 11, 347–356. [Google Scholar] [CrossRef]
  67. Korhonen, H.; Pihlanto, A. Food-derived bioactive peptides opportunities for designing future foods. Curr. Pharm. Des. 2003, 9, 1297–1308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Martínez-Medina, G.A.; Prado-Barragán, A.; Martínez Hernández, J.L.; Ruíz, H.A.; Rodríguez, M.R.; Contreras-Esquivel, J.C.; Aguilar, C.N. Peptidos bio-funcionales: Bioactividad, producción y aplicaciones. J. Chem. Technol. Biotechnol. 2019, 13, 1–7. [Google Scholar]
  69. Mora, L.; Reig, M.; Toldrá, F. Bioactive peptides generated from meat industry by-products. Food Res. Int. 2014, 65, 344–349. [Google Scholar] [CrossRef] [Green Version]
  70. Albericio, F.; Jiménez, J.C.; Giralt, E. Péptidos y la Industria Farmacéutica. In Anales de la Real Sociedad Española de Química; Real Sociedad Española de Química: Madrid, Spain, 2004; pp. 10–16. [Google Scholar]
  71. Inouye, K.; Nakano, K.; Asaoka, K.; Yasukawa, K. Effects of thermal treatment on the coagulation of soy proteins induced by subtiLysin Carlsberg. J. Agric. Food Chem. 2009, 57, 717–723. [Google Scholar] [CrossRef]
  72. Korhonen, H.; Pihlanto, A. Bioactive peptides: Production and functionality. Int. Dairy J. 2006, 16, 945–960. [Google Scholar] [CrossRef]
  73. Korhonen, H. Milk-derived bioactive peptides: From science to applications. J. Funct. Foods 2009, 1, 177–187. [Google Scholar] [CrossRef]
  74. Meisel, H.; FitzGerald, R.J. Biofunctional peptides from milk proteins: Mineral binding and cytomodulatory effects. Curr. Pharm. Des. 2003, 9, 1289–1295. [Google Scholar] [CrossRef]
  75. Yamamoto, N.; Ejiri, M.; Mizuno, S. Biogenic peptides and their potential use. Curr. Pharm. Des. 2003, 9, 1345–1355. [Google Scholar] [CrossRef]
  76. Gobbetti, M.; Minervini, F.; Rizzello, C.G. Bioactive peptides in dairy products. In Handbook of Food Products Manufacturing; Hui, Y.H., Ed.; John Wiley & Sons, Inc.: San Francisco, CA, USA, 2007; pp. 489–517. [Google Scholar] [CrossRef]
  77. Suetsuna, K.; Osajima, K. Blood pressure reduction and vasodilatory effects in vivo of peptides originating from sardine muscle. J. Jpn. Soc. Nutr. Food Sci. 1989, 42, 47–54. [Google Scholar] [CrossRef] [Green Version]
  78. Suetsuna, K.; Chen, J.R. Identification of antihypertensive peptides from peptic digest of two microalgae Chlorella vulgaris and Spirulina platensis. Mar. Biotechnol. 2001, 3, 305–309. [Google Scholar] [CrossRef] [PubMed]
  79. McDonagh, D.; FitzGerald, R.J. Production of caseinophosphopeptides (CPPs) from sodium caseinate using a range of commercial protease preparations. Int. Dairy J. 1998, 8, 39–45. [Google Scholar] [CrossRef]
  80. Pihlanto-Leppälä, A.; Koskinen, P.; Piilola, K.; Tupasela, T.; Korhonen, H. Angiotensin I-converting enzyme inhibitory properties of whey protein digests: Concentration and characterization of active peptides. J. Dairy Res. 2000, 67, 53–64. [Google Scholar] [CrossRef]
  81. Vermeirssen, V.; Van Camp, J.; Verstraete, W. Bioavailability of angiotensin I-converting enzyme inhibitory peptides. Br. J. Nutr. 2004, 92, 357–366. [Google Scholar] [CrossRef]
  82. Roufik, S.; Gauthier, S.F.; Turgeon, S.L. In vitro digestibility of bioactive peptides derived from bovine ß-lactoglobulin. Int. Dairy J. 2006, 16, 294–302. [Google Scholar] [CrossRef]
  83. De Costa, E.L.; Rocha, J.A.; Netto, F.M. Effect of heat and enzymatic treatment on the antihypertensive activity of whey protein hydrolysates. Int. Dairy J. 2007, 17, 632–640. [Google Scholar] [CrossRef]
  84. López-Expósito, I.; Recio, I. Antibacterial activity of peptides and folding variants from milk proteins. Int. Dairy J. 2006, 16, 1294–1305. [Google Scholar] [CrossRef]
  85. López-Expósito, I.; Quiros, A.; Amigo, L.; Recio, I. Caseinhydrolysates as a source of antimicrobial, antioxydant and antihypertensive peptides. Lait 2007, 87, 241–249. [Google Scholar] [CrossRef]
  86. Pihlanto, A. Antioxidative peptides derived from milk proteins. Int. Dairy J. 2006, 16, 1306–1314. [Google Scholar] [CrossRef]
  87. Gauthier, S.F.; Pouliot, Y.; Saint-Sauveur, D. Immunomodulatory peptides obtained by the enzymatic hydrolysis of whey proteins. Int. Dairy J. 2006, 16, 1315–1323. [Google Scholar] [CrossRef]
  88. Teschemacher, H. Opioid receptor ligands derived from food proteins. Curr. Pharm. Des. 2003, 9, 1331–1344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Singh, B.P.; Vij, S.; Hati, S. Functional significance of bioactive peptides derived from soybean. Peptides 2014, 54, 171–179. [Google Scholar] [CrossRef]
  90. Mwinw-Daliri, E.B.M.; Oh, D.H.; Lee, B.H. Bioactive Peptides. Foods 2017, 6, 32. [Google Scholar] [CrossRef] [PubMed]
  91. Kim, S.K.; Wijesekara, I. Development and biological activities of marine–derived bioactive peptides: A Review. J. Funct. Foods 2010, 2, 1–9. [Google Scholar] [CrossRef]
  92. Intiquilla, A.; Jiménez-Aliaga, K.; Guzmán, F.; Álvarez, C.A.; Zavaleta, A.I.; Izaguirre, V. Novel antioxidant peptides obtained by alcalase hydrolysis of Erythrina edulis (pajuro) protein. J. Sci. Food Agric. 2019, 99, 2420–2427. [Google Scholar] [CrossRef]
  93. Maqsoudlou, A.; Mahoonak, A.S.; Mora, L.; Mohebodini, H.; Toldra, F.; Ghorbani, M. Peptide identification in alcalase hydrolysated pollen and comparison of its bioactivity with royal jelly. Food Res. Int. 2019, 116, 905–915. [Google Scholar] [CrossRef]
  94. Aluko, R.E. Bioactive Peptides. In Functional Foods and Nutraceuticals; Food Science Text Series; Springer: New York, NY, USA, 2012; pp. 37–61. [Google Scholar] [CrossRef]
  95. Hernandez-Ledesma, B.; Del Mar Contreras, M.; Recio, I. Antihypertensive peptides: Production, bioavailability and incorporation into foods. Adv. Colloid. Interface Sci. 2011, 165, 23–35. [Google Scholar] [CrossRef] [Green Version]
  96. Minekus, M.; Alminger, M.; Alvito, P.; Ballance, S.; Bohn, T.; Bourlieu, C. A standardised static in vitro digestion method suitable for food—An international consensus. Food Funct. 2014, 5, 1113–1124. [Google Scholar] [CrossRef] [Green Version]
  97. Najafian, L.; Salam, A. Fractionation and identification of novel antioxidant peptides from fermented fish (Pekasam). J. Food. Meas. Charact. 2018, 12, 2174–2183. [Google Scholar] [CrossRef]
  98. Rajendran, S.R.C.K.; Mohan, A.; Khiari, Z.; Udenigwe, C.C.; Mason, B. Yield, physicochemical, and antioxidant properties of Atlantic Salmon visceral hydrolysate: Comparison of lactic acid bacterial fermentation with Flavourzyme proteolysis and formic acid treatment. J. Food Process. Preserv. 2018, 42, e13620. [Google Scholar] [CrossRef]
  99. Christensen, J.E.; Dudley, E.G.; Pederson, J.A.; Steele, J.L. Peptidases and amino acid cataboLysm in lactic acid bacteria. Antonie Leeuwenhoek 1999, 80, 155–165. [Google Scholar]
  100. KulLysaar, T.; Songisepp, E.; Mikelsaar, M.; Zilmer, K.; Vihalemm, T.; Zilmer, M. Antioxidative probiotic fermented goats’ milk decreases oxidative stress-mediated atherogenicity in human subjects. Br. J. Nutr. 2003, 90, 449–456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Lin, M.Y.; Yen, C.L. Reactive oxygen species and lipid peroxidation product-scavenging ability of yoghurt organisms. J. Dairy Sci. 1999, 82, 1629–1634. [Google Scholar] [CrossRef]
  102. Virtanen, T.; Pihlanto, A.; Akkanen, S.; Korhonen, H. Development of antioxidant activity in milk whey during fermentation with lactic acid bacteria. J. Appl. Microbiol. 2007, 102, 106–115. [Google Scholar] [CrossRef]
  103. Rajapakse, N.; Mendis, E.; Jung, W.K.; Je, J.Y.; Kim, S.K. Purification of a radical scavenging peptide from fermented mussel sauce and its antioxidant properties. Food Res. Int. 2005, 38, 175–182. [Google Scholar] [CrossRef]
  104. Chaudhury, A.; Duvoor, C.; Reddy-Dendi, V.S.; Kraleti, S.; Chada, A.; Ravilla, R.; Marco, A.; Shekhawat, N.S.; Montales, M.T.; Kuriakose, K. Clinical review of antidiabetic drugs: Implications for type 2 diabetes mellitus management. Front. Endocrinol. 2017, 8, 6. [Google Scholar] [CrossRef] [Green Version]
  105. Sanjukta, S.; Rai, A.K.; Muhammed, A.; Jeyaram, K.; Talukdar, N.C. Enhancement of antioxidant properties of two soybean varieties of Sikkim Himalayan region by proteolytic Bacillus subtilis fermentation. J. Funct. Foods 2015, 14, 650–658. [Google Scholar] [CrossRef]
  106. Mitchell, D.A.; Berovič, M.; Krieger, N. Solid-state fermentation bioreactor fundamentals: Introduction and overview. In Solid-State Fermentation Bioreactors; Mitchell, D.A., Berovič, M., Krieger, N., Eds.; Springer: Berlin/Heidelberg, Germany, 2006; pp. 1–12. [Google Scholar] [CrossRef]
  107. Gohi, B.F.C.A.; Du, J.; Zeng, H.-Y.; Cao, X.; Zou, K. Microwave pretreatment and enzymolysis optimization of the Lotus seed protein. Bioengineering 2019, 6, 28. [Google Scholar] [CrossRef] [Green Version]
  108. Wen, C.; Zhang, J.; Zhou, J.; Cai, M.; Duan, Y.; Zhang, H.; Ma, H. Antioxidant activity of arrowhead protein hydrolysates produced by a novel multi-frequency S-type ultrasound-assisted enzymolysis. Nat. Prod. Res. 2020, 34, 3000–3003. [Google Scholar] [CrossRef]
  109. Dong, X.; Li, J.; Jiang, G.; Li, H.; Zhao, M.; Jiang, Y. Effects of combined high pressure and enzymatic treatments on physicochemical and antioxidant properties of Peanut proteins. Food Sci. Nutr. 2019, 7, 1417–1425. [Google Scholar] [CrossRef]
  110. Zhang, M.; Mu, T. Identification and characterization of antioxidant peptides from Sweet Potato protein hydrolysates by Alcalase under high hydrostatic pressure. Innov. Food Sci. Emerg. Technol. 2017, 43, 92–101. [Google Scholar] [CrossRef]
  111. Bamdad, F.; Shin, S.H.; Suh, J.; Nimalaratne, C. 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]
  112. Jin, H.; Xu, H.; Li, Y.; Zhang, Q.; Xie, H. Preparation and evaluation of peptides with potential antioxidant activity by microwave assisted enzymatic hydrolysis of collagen from sea Cucumber Acaudina Molpadioides obtained from Zhejiang province in China. Mar. Drugs 2019, 17, 169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Ketnawa, S.; Liceaga, A.M. Effect of Microwave Treatments on Antioxidant Activity and Antigenicity of Fish Frame Protein Hydrolysates. Food Bioprocess Technol. 2017, 10, 582–591. [Google Scholar] [CrossRef]
  114. Nguyen, E.; Jones, O.; Kim, Y.H.B.; San Martin-Gonzalez, F.; Liceaga, A.M. Impact of microwave-assisted enzymatic hydrolysis on functional and antioxidant properties of rainbow trout Oncorhynchus mykiss by-products. Fish. Sci. 2017, 83, 317–331. [Google Scholar] [CrossRef]
  115. Ozuna, C.; Paniagua-Martínez, I.; Castaño-Tostado, E.; Ozimek, L.; Amaya-Llano, S.L. Innovative applications of high-intensity ultrasound in the development of functional food ingredients: Production of protein hydrolysates and bioactive peptides. Food Res. Int. 2015, 77, 685–696. [Google Scholar] [CrossRef]
  116. Zou, Y.; Wang, W.; Li, Q.; Chen, Y.; Zheng, D.; Zhang, M.; Zhao, T.; Mao, G.; Feng, W.; Wu, X.; et al. Physicochemical, functional properties and antioxidant activities of Porcine Cerebral hydrolysate peptides produced by ultrasound processing. Process Biochem. 2015, 3, 441–443. [Google Scholar] [CrossRef]
  117. Sae-leaw, T.; Benjakul, S. Antioxidant activities of hydrolysed collagen from salmon scale ossein prepared with the aid of ultrasound. Int. J. Food Sci. Technol. 2018, 53, 2786–2795. [Google Scholar] [CrossRef]
  118. Chian, F.M.; Kaur, L.; Oey, I.; Astruc, T. Effect of pulsed electric fields (PEF) on the ultrastructure and in vitro protein digestibility of bovine Longissimus thoracis. LWT Food Sci. Technol. 2019, 103, 253–259. [Google Scholar] [CrossRef]
  119. Cho, Y.; Haq, M.; Park, J.; Lee, H.; Chun, B. Physicochemical and Biofunctional properties of Shrimp (Penaeus japonicus) hydrolysates obtained from hot-compressed water treatment. J. Supercrit. Fluids 2018, 147, 322–328. [Google Scholar] [CrossRef]
  120. Alvarez, C.; Tiwari, B.K.; Rendueles, M.; Díaz, M. Use of response surface methodology to describe the effect of time and temperature on the production of decoloured, antioxidant and functional peptides from Porcine haemoglobin by subcritical water hydrolysis. LWT Food Sci. Technol. 2016, 73, 280–289. [Google Scholar] [CrossRef]
  121. Mikhaylin, S.; Boussetta, N.; Vorobiev, E.; Bazinet, L. High voltage electric treatments to improve the protein susceptibility to enzymatic hydrolysis. ACS Sustain. Chem. Eng. 2017, 5, 11706–11714. [Google Scholar] [CrossRef]
  122. Ahmed, R.; Chun, B. Supercritical Fluids Subcritical water hydrolysis for the production of bioactive peptides from Tuna skin collagen. J. Supercrit. Fluids 2018, 141, 88–96. [Google Scholar] [CrossRef]
  123. Zou, T.-B.; He, T.-P.; Li, H.-B.; Tang, H.-W.; Xia, E.-Q. The Structure-Activity Relationship of the Antioxidant Peptides from Natural Proteins. Molecules 2016, 21, 72. [Google Scholar] [CrossRef]
  124. Yang, B.; Yang, H.; Li, J.; Li, Z.; Jiang, Y. Amino acid composition, molecular weight distribution and antioxidant activity of protein hydrolysates of soy sauce lees. Food Chem. 2011, 124, 551–555. [Google Scholar] [CrossRef]
  125. Pedersen Søren, L.; Tofteng, A.P.; Malik, L.; Jensen, K.J. Microwave heating in solid-phase peptide synthesis. Chem. Soc. Rev. 2012, 41, 1826–1844. [Google Scholar] [CrossRef] [Green Version]
  126. Pramanik, B.; Mirza, U.; Ing, Y.; Liu, Y.; Bartner, P.; Weber, P.; Bose, M. Microwave-enhanced enzyme reaction for protein mapping by mass spectrometry: A new approach to protein digestion in minutes. Protein Sci. 2002, 11, 2676–2687. [Google Scholar] [CrossRef]
  127. Reddy, P.; Hsu, W.; Hu, J.; Ho, Y. Digestion Completeness of Microwave-Assisted and Conventional Trypsin-Catalyzed Reactions. J. Am. Soc. Mass Spectrom. 2010, 21, 421–424. [Google Scholar] [CrossRef] [Green Version]
  128. Mendiola, J.A. Extracción de Compuestos Bioactivos de Microalgas Mediante Fluidos Supercríticos. Ph.D. Thesis, Universidad Autónoma de Madrid, Madrid, Spain, 2008. [Google Scholar]
  129. Lang, Q.; Wai, C.M. Supercritical fluid extraction in herbal and natural product studies a practical review. Talanta 2001, 53, 771–782. [Google Scholar] [CrossRef]
  130. Olivera, L. Hidrolizado Proteico de Quinua (Chenopodium quinoa Willd) con Péptidos Bioactivos y Capacidad Antioxidante: Evaluación de la Factibilidad Económica. Ph.D. Thesis, Universidad Nacional Mayor de San Marcos, Facultad de Ingeniería Industrial, Unidad de Posgrado, Lima, Perú, 2021. [Google Scholar]
  131. Holton, T.A.; Vijayakumar, V.; Khaldi, N. Bioinformatics: Current perspectives and future directions for food and nutritional research facilitated by a Food-Wiki database. Trends Food Sci. Technol. 2013, 34, 5–17. [Google Scholar] [CrossRef]
  132. Lacroix, I.M.; Li-Chan, E.C. Evaluation of the potential of dietary proteins as precursors of dipeptidyl peptidase (DPP)-IV inhibitors by an in silico approach. J. Funct. Foods 2012, 4, 403–422. [Google Scholar] [CrossRef]
  133. Undenigwe, C.C.; Aluko, R.E. Food Protein-Derived Bioactive peptides: Production, Processing, and Potential Health Benefits. J. Food Sci. 2012, 71, R11–R24. [Google Scholar] [CrossRef] [PubMed]
  134. Dziuba, B.; Dziuba, M. Milk proteins-derived bioactive peptides in dairy products: Molecular, biological and methodological aspects. Acta Sci. Pol. Technol. Aliment. 2014, 13, 5–25. [Google Scholar] [CrossRef] [PubMed]
  135. Iwaniak, A.; Minkiewicz, P.; Darewicz, M.; Protasiewicz, M.; Mogut, D. Chemometrics and cheminformatics in the analysis of biologically active peptides from food sources. J. Funct. Foods 2015, 16, 334–351. [Google Scholar] [CrossRef]
  136. Kumar, R.; Chaudhary, K.; Sharma, M.; Nagpal, G.; Chauhan Singh, J.; Singh, S.; Gautam, A.; Raghava, G.P.S. AHTPDB: A comprehensive platform for analysis and presentation of antihypertensive peptides. Nucleic Acids Res. 2015, 43, D956–D962. [Google Scholar] [CrossRef] [Green Version]
  137. Usmani, S.S.; Kumar, R.; Kumar, V.; Singh, S.; Raghava, G.P.S. AntiTbPdb: A knowledgebase of anti-tubercular peptides. Database 2018, 2018, bay025. [Google Scholar] [CrossRef]
  138. Wang, G.; Li, X.; Wang, Z. APD3: The antimicrobial peptide database as a tool for research and education. Nucleic Acids Res. 2016, 44, D1087–D1093. [Google Scholar] [CrossRef] [Green Version]
  139. Qureshi, A.; Thakur, N.; Himani, T.; Kumar, M. AVPdb: A database of experimentally validated antiviral peptides targeting medically important viruses. Nucleic Acids Res. 2013, 42, D1147–D1153. [Google Scholar] [CrossRef] [Green Version]
  140. Di Luca, M.; Maccari, G.; Maisetta, G.; Batoni, G. BaAMPs: The database of biofilm-active antimicrobial peptides. Biofouling 2015, 31, 193–199. [Google Scholar] [CrossRef]
  141. Rey, J.; Deschavanne, P.; Tuffery, P. BactPepDB: A database of predicted peptides from a exhaustive survey of complete prokaryote genomes. Database 2014, 2014, bau106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Minkiewicz, P.; Dziuba, J.; Iwaniak, A.; Dziuba, M.; Darewicz, M. BIOPEP and other programs for processing bioactive peptide sequences. J. AOAC Int. 2008, 91, 965–980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Van Dorpe, S.; Bronselaer, A.; Nielandt, J.; Stalmans, S.; Wynendaele, E.; Audenaert, K.; De Spiegeleer, B. Brainpeps: The blood-brain barrier peptide database. Brain Struct. Funct. 2012, 217, 687–718. [Google Scholar] [CrossRef]
  144. Waghu, F.H.; Barai, R.S.; Gurung, P.; Idicula-Thomas, S. CAMPR3: A database on sequences, structures, and signatures of antimicrobial peptides. Nucleic Acids Res. 2015, 44, D1094–D1097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Tyagi, A.; Tuknait, A.; Anand, P.; Gupta, S.; Sharma, M.; Mathur, D.; Raghava, G.P.S. CancerPPD: A database of anticancer peptides and proteins. Nucleic Acids Res 2015, 43, D837–D843. [Google Scholar] [CrossRef] [Green Version]
  146. Lindgren, M.; Hällbrink, M.; Prochiantz, A.; Langel, Ü. Cell-penetrating peptides. Trends Pharmacol. Sci. 2000, 21, 99–103. [Google Scholar] [CrossRef]
  147. Pirtskhalava, M.; Gabrielian, A.; Cruz, P.; Griggs, H.L.; Squires, R.B.; Hurt, D.E.; Tartakovsky, M. DBAASP vol 2: An enhanced database of structure and antimicrobial/cytotoxic activity of natural and synthetic peptides. Nucleic Acids Res. 2016, 44, D1104–D1112. [Google Scholar] [CrossRef]
  148. Zamyatnin, A.A.; Borchikov, A.S.; Vladimirov, M.G.; Voronina, O.L. The EROP-Moscow oligopeptide database. Nucleic Acids Res. 2006, 34, D261–D266. [Google Scholar] [CrossRef] [Green Version]
  149. Gautam, A.; Chaudhary, K.; Singh, S.; Joshi, A.; Anand, P.; Tuknait, A.; Raghava, G.P.S. Hemolytik: A database of experimentally determined hemolytic and nonhemolytic peptides. Nucleic Acids Res. 2014, 42, D444–D449. [Google Scholar] [CrossRef] [Green Version]
  150. Nielsen, S.D.; Beverly, R.L.; Qu, Y.; Dallas, D.C. Milk bioactive peptide database: A comprehensive database of milk protein-derived peptides and novel visualization. Food Chem. 2017, 232, 673–682. [Google Scholar] [CrossRef]
  151. Wang, Y.; Wang, M.; Yin, S.; Jang, R.; Wang, J.; Xue, Z.; Xu, T. NeuroPep: A comprehensive resource of neuropeptides. Database 2015, 2015, bav038. [Google Scholar] [CrossRef] [Green Version]
  152. Shtatland, T.; Guettler, D.; Kossodo, M.; Pivovarov, M.; Weissleder, R. PepBank–A database of peptides based on sequence text mining and public peptide data sources. BMC Bioinform. 2007, 8, e280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Wynendaele, E.; Bronselaer, A.; Nielandt, J.; D’Hondt, M.; Stalmans, S.; Bracke, N.; De Spiegeleer, B. Quorumpeps database: Chemical space, microbial origin and functionality of quorum sensing peptides. Nucleic Acids Res. 2013, 41, D655–D659. [Google Scholar] [CrossRef] [PubMed]
  154. Singh, P.; Singh, T.P.; Gandhi, N. Prevention of lipid oxidation in muscle foods by milk proteins and peptides: A review. Food Rev. Int. 2016, 34, 226–247. [Google Scholar] [CrossRef]
  155. Wang, J.; Yin, T.; Xiao, X.; He, D.; Xue, Z.; Jiang, X.; Wang, Y. StraPep: A structure database of bioactive peptides. Database 2018, 2018, bay038. [Google Scholar] [CrossRef] [PubMed]
  156. Kapoor, P.; Singh, H.; Gautam, A.; Chaudhary, K.; Kumar, R.; Raghava, G.P.S.; Xue, B. TumorHoPe: A database of tumor homing peptides. PLoS ONE 2012, 7, e35187. [Google Scholar] [CrossRef] [Green Version]
  157. Piotto, S.P.; Sessa, L.; Concilio, S.; Ianelli, P. YADAMP: Yet another database of antimicrobial peptides. Int. J. Antimicrob. 2012, 39, 346–351. [Google Scholar] [CrossRef]
  158. Gu, Y.; Majumder, K.; Wu, J. QSAR-aided in silico approach in evaluation of food proteins as precursors of ACE inhibitory peptides. Food Res. Int. 2011, 44, 2465–2474. [Google Scholar] [CrossRef]
  159. Barrero, J.A.; Cruz, C.M.; Casallas, J.; Vásquez, J.S. Evaluación in silico de péptidos bioactivos derivados de la digestión de proteínas presentes en la leche de bovino (B. taurus), oveja (O. aries), cabra (C. hircus) y búfalo (B. bubalis). TecnoLógicas 2021, 24, e1731. [Google Scholar] [CrossRef]
  160. Panjaitan, F.C.A.; Gómez, H.L.R.; Chang, Y.-W. In silico analysis of bioactive peptides released from giant grouper (Epinephelus lancetolatus) roe proteins identified by proteomic approach. Molecules 2018, 23, 2910. [Google Scholar] [CrossRef] [Green Version]
  161. Rawlings, N.D.; Barrett, A.J.; Finn, R. MEROPS: The database of proteolytic enzymes, their substrates and inhibitors. Nucleic Acids Res. 2016, 44, D343–D350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Chen, H.M.; Muramoto, K.; Yamauchi, F.; Nokihara, K. Antioxidant activity of design peptides based on the antioxidative peptide isolated from digests of a soybean protein. J. Agric. Food Chem. 1996, 44, 2619–2623. [Google Scholar] [CrossRef]
  163. Hattori, M.; Yamaji-Tsukamoto, K.A.; Kumagai, H.; Feng, Y.; Takahashi, K. Antioxidative activity of soluble elastin peptides. J. Agric. Food Chem. 1998, 46, 2167–2170. [Google Scholar] [CrossRef]
  164. Xie, Z.; Huang, J.; Xu, X.; Jin, Z. Antioxidant activity of peptides isolated from alfalfa leaf protein hydrolysate. Food Chem. 2008, 111, 370–376. [Google Scholar] [CrossRef] [PubMed]
  165. Vioque, J.; Clemente, A.; Pedroche, J.; Yust, M.M.; Millán, F. Obtención y aplicaciones de hidrolizados proteicos. Grasas Aceites 2001, 52, 132–136. [Google Scholar] [CrossRef] [Green Version]
  166. Turgeon, S.L.; Gauthier, S.F.; Paquin, P. Emulsifying property of whey peptide fractions as a function of Ph and ionic strengh. J. Food Sci. 1992, 57, 601–604, 634. [Google Scholar] [CrossRef]
  167. Kim, S.Y.; Park, P.S.W.; Rhee, K.C. Functional properties of proteolytic enzyme modified soy protein isolate. J. Agric. Food Chem. 1990, 38, 651–656. [Google Scholar] [CrossRef]
  168. Süle, E.; Tömösközi, S.; Hajós, G. Functional properties of enzymatically modified plant proteins. Food/Nahr. 1998, 42, 242–244. [Google Scholar] [CrossRef]
  169. Clemente, A.; Vioque, J.; Sánchez-Vioque, R.; Pedroche, J.; Bautista, J.; Millán, F. Protein quality of chickpea protein hydrolysates. Food Chem. 1999, 67, 269–274. [Google Scholar] [CrossRef]
  170. Vioque, J.; Sánchez-Vioque, R.; Clemente, A.; Pedroche, J.; Millán, F. Partially hydrolyzed rapeseed protein isolates with improved functional properties. JAOCS 2000, 77, 447–450. [Google Scholar] [CrossRef]
  171. Girgih, A.T.; He, R.; Hasan, F.M.; Udenigwe, C.C.; Gill, T.A.; Aluko, R.E. Evaluation of the in vitro antioxidant properties of a cod (Gadus morhua) protein hydrolysate and peptide fractions. Food Chem. 2015, 173, 652–659. [Google Scholar] [CrossRef] [PubMed]
  172. He, Y.; Pan, X.; Chi, C.; Sun, K.; Wang, B. Ten new pentapeptides from protein hydrolysate of Miiuy Croaker (Miichthis miiuy) muscle: Preparation, identification, and antioxidant activity evaluation. LWT Food Sci. Technol. 2019, 105, 1–8. [Google Scholar] [CrossRef]
  173. Agrawal, H.; Joshi, R.; Gupta, M. Purification, identification and characterization of two novel antioxidant peptides from finger millet (Eleusine coracana) protein hydrolysate. Food Res. Int. 2019, 120, 697–707. [Google Scholar] [CrossRef] [PubMed]
  174. Neves, A.C.; Harnedy, P.A.; O’Keeffe, M.B.; Alashi, M.A.; Aluko, R.E.; FitzGerald, R.J. Peptide identification in a salmon gelatin hydrolysate with antihypertensive, dipeptidyl peptidase IV inhibitory and antioxidant activities. Food Res. Int. 2017, 100, 112–120. [Google Scholar] [CrossRef]
  175. Ambigaipalan, P.; Shahidi, F. Bioactive peptides from shrimp shell processing discards: Antioxidant and biological activities. J. Funct. Foods 2017, 34, 7–17. [Google Scholar] [CrossRef]
  176. Neves, A.C.; Harnedy, P.A.; O’Keeffe, M.B.; FitzGerald, R.J. Bioactive peptides from Atlantic salmon (Salmo salar) with angiotensin converting enzyme and dipeptidyl peptidase IV inhibitory, and antioxidant activities. Food Chem. 2016, 218, 396–405. [Google Scholar] [CrossRef]
  177. Zenezini Chiozzi, R.; Capriotti, A.L.; Cavaliere, C.; La Barbera, G.; Piovesana, S.; Samperi, R.; Lagana, A. Purification and identification of endogenous antioxidant and ACE-inhibitory peptides from donkey milk by multidimensional liquid chromatography and nanoHPLC-high resolution mass spectrometry. Anal. Bioanal. Chem. 2016, 408, 5657–5666. [Google Scholar] [CrossRef]
  178. Ahmed, A.S.; El-Bassiony, T.; Elmalt, L.M.; Ibrahim, H.R. Identification of potent antioxidant bioactive peptides from goat milk proteins. Food Res. Int. 2015, 74, 80–88. [Google Scholar] [CrossRef]
  179. Chang, S.K.; Ismail, A.; Yanagita, T.; Mohd Esa, N.; Baharuldin, M.T.H. Antioxidant peptides purified and identified from the oil palm (Elaeis guineensis Jacq.) kernel protein hydrolysate. J. Funct. Foods 2015, 14, 63–75. [Google Scholar] [CrossRef]
  180. Gu, M.; Chen, H.-P.; Zhao, M.M.; Wang, X.; Yang, B.; Ren, J.Y. Identification of antioxidant peptides released from defatted walnut (Juglans Sigillata Dode) meal proteins with pancreatin. LWT Food Sci. Technol. 2015, 60, 213–220. [Google Scholar] [CrossRef]
  181. Liu, J.; Jin, Y.; Lin, S.; Jones, G.S.; Chen, F. Purification and identification of novel antioxidant peptides from egg white protein and their antioxidant activities. Food Chem. 2015, 175, 258–266. [Google Scholar] [CrossRef] [PubMed]
  182. Sudhakar, S.; Nazeer, R.A. Structural characterization of an Indian squid antioxidant peptide and its protective effect against cellular reactive oxygen species. J. Funct. Foods 2015, 14, 502–512. [Google Scholar] [CrossRef]
  183. Tenore, G.C.; Ritieni, A.; Campiglia, P.; Stiuso, P.; Di Maro, S.; Sommella, E.; Pepe, G.; D’Urso, E.; Novellino, E. Antioxidant peptides from “Mozzarella di Bufala Campana DOP” after simulated gastrointestinal digestion: In vitro intestinal protection, bioavailability, and anti-haemolytic capacity. J. Funct. Foods 2015, 15, 365–375. [Google Scholar] [CrossRef]
  184. Zeng, W.C.; Zhang, W.H.; Qiang, H.; Bi, S. Purification and characterization of a novel antioxidant peptide from bovine hair hydrolysates. Process Biochem. 2015, 50, 948–954. [Google Scholar] [CrossRef]
  185. González-García, E.; Marina, M.L.; García, M.C. Plum (Prunus Domestica L.) by-product as a new and cheap source of bioactive peptides: Extraction method and peptides characterization. J. Funct. Foods 2014, 11, 428–437. [Google Scholar] [CrossRef]
  186. Herrera-Chalé, F.G.; Ruiz-Ruiz, J.C.; Acevedo-Fernández, J.J.; Betancur-Ancona, D.A.; Segura-Campos, M.R. ACE inhibitory, hypotensive and antioxidant peptide fractions from Mucuna pruriens proteins. Process Biochem. 2014, 49, 1691–1698. [Google Scholar] [CrossRef]
  187. Ji, N.; Sun, C.; Zhao, Y.; Xiong, L.; Sun, Q. Purification, and identification of antioxidant peptides from peanut protein isolate hydrolysates using UHR-Q-TOF mass spectrometer. Food Chem. 2014, 161, 148–154. [Google Scholar] [CrossRef]
  188. Ren, Y.; Wu, H.; Li, X.; Lai, F.; Xiao, X. Purification, and characterization of high antioxidant peptides from duck egg white protein hydrolysates. Biochem. Biophys. Res. Commun. 2014, 452, 888–894. [Google Scholar] [CrossRef]
  189. Zarei, M.; Ebrahimpour, A.; Abdul-Hamid, A.; Anwar, F.; Bakar, F.A.; Philip, R.; Saari, N. Identification and characterization of papain-generated antioxidant peptides from palm kernel cake proteins. Food Res. Int. 2014, 62, 726–734. [Google Scholar] [CrossRef]
  190. Gómez, L.J.; Figueroa, O.A.; Zapata, J.E. Actividad antioxidante de hidrolizados enzimáticos de plasma bovino obtenidos por efecto de Alcalasa® 2.4 L. Inf. Tecnol. 2013, 24, 33–42. [Google Scholar] [CrossRef] [Green Version]
  191. Moure, A.; Domínguez, H.; Parajó, J. Antioxidant properties of ultrafiltration recovered soy protein fractions from industrial effluents and their hydrolysates. Process Biochem. 2006, 41, 447–456. [Google Scholar] [CrossRef]
  192. Qian, Z.J.; Jung, W.K.; Kim, S.K. Free radical scavenging activity of a novel Antioxidative peptide purified from hydrolysate of bullfrog skin, Rana catesbeiana shaw. Bioresour. Technol. Rep. 2008, 99, 1690–1698. [Google Scholar] [CrossRef] [PubMed]
  193. Wu, C.H.; Chen, H.M.; Shiau, C.Y. Free amino acids and peptides as related to antioxidant properties in protein hydrolysates of mackerel (Scomber austriasicus). Food Res. Int. 2003, 36, 949–957. [Google Scholar] [CrossRef]
  194. Xiong, Y.L. Antioxidant peptides. In Bioactive Proteins and Peptides as Functional Foods and Nutraceuticals; Wiley-Blackwell: Oxford, UK, 2010; pp. 29–42. [Google Scholar] [CrossRef]
  195. Gallegos Tintoré, S.; Chel Guerrero, L.; Corzo Ríos, L.J.; Martínez Ayala, A.L. Péptidos con actividad antioxidante de proteínas vegetales. In Bioactividad de Péptidos Derivados de Proteínas Alimentarias; Segura Campos, M.L., Guerrero, C., Betancur Ancona, D., Eds.; OmniaScience: Barcelona, Spain, 2013; pp. 111–122. [Google Scholar] [CrossRef]
  196. Walrant, A.; Bauzá, A.; Girardet, C.; Alves, I.D.; Lecomte, S.; Illien, F. Ionpair-π 1581 interactions favor cell penetration of arginine/tryptophan-rich cell-penetrating peptides. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183098. [Google Scholar] [CrossRef]
  197. Ningappa, M.B.; Srinivas, L. Purification and characterization of ~35 kDa antioxidant protein from curry leaves (Murraya koenigii L.). Toxicol. Vitr. 2008, 22, 699–709. [Google Scholar] [CrossRef] [PubMed]
  198. Cian, R.E.; Vioque, J.; Drago, S.R. Structure–mechanism relationship of antioxidant and ACE I inhibitory peptides from wheat gluten hydrolysate fractionated by pH. Food Res. Int. 2015, 69, 216–223. [Google Scholar] [CrossRef]
  199. Saito, K.; Jin, D.H.; Ogawa, T.; Muramoto, K.; Hatakeyama, E.; Yasuhara, T. Antioxidative properties of tripeptide libraries prepared by the combinatorial chemistry. J. Agric. Food Chem. 2003, 51, 3668–3674. [Google Scholar] [CrossRef]
  200. Bulet, P.; Stöcklin, R. Insect antimicrobial peptides: Structures, properties and gene regulation. Protein Pept. Lett. 2005, 12, 3–11. [Google Scholar] [CrossRef]
  201. Guerreiro, C.I.; Fontes, C.M.; Gama, M.; Domigues, L. Escherichia coli expresion and purification of our antimicrobial peptides fused to a family 3 cabohydrate-binding module (CBM) from Clostridium thermocellum. Protein Expr. Purif. 2008, 59, 161–168. [Google Scholar] [CrossRef] [Green Version]
  202. Wang, X.; Zhu, M.; Zhang, A.; Yang, F.; Chen, P. Synthesis and secretoory expression of hybrid antimicrobial peptide CecA-mag and its mutants in Pichia Pastoris. Exp. Biol. Med. 2012, 237, 312–317. [Google Scholar] [CrossRef]
  203. Shabir, U.; Ali, S.; Magray, A.R.; Ganai, B.A.; Firdous, P.; Hassan, T.; Nazir, R. Fish antimicrobial peptides (AMP’s) as essential and promising molecular therapeutic agents: A review. Microb. Pathog. 2018, 114, 50–56. [Google Scholar] [CrossRef]
  204. Barbosa-Pelegrini, P.; Del Sarto, R.P.; Silva, O.N.; Franco, O.L.; Grossi-de-Sa, M.F. Antibacterial peptides from plants: What they are and how they probably work. Biochem. Res. Int. 2011, 2011, 250349. [Google Scholar] [CrossRef] [Green Version]
  205. Lopes, J.L.; Nobre, T.M.; Siano, A.; Humpola, V.; Bossolan, N.R.; Zaniquelli, M.E. Disruption of Saccharomyces cerevisiae by Plantaricin 149 and investigation of its mechanism of action with biomembrane model systems. Biochim. Biophys. Acta 2009, 1788, 2252–2258. [Google Scholar] [CrossRef] [PubMed]
  206. Jean-Francois, F.; Elezgaray, J.; Berson, P.; Vacher, P.; Dufourc, E.J. Pore formation induced by an antimicrobial peptide: Electrostatic effects. Biophys. J. 2008, 95, 5748–5756. [Google Scholar] [CrossRef] [Green Version]
  207. Scocchi, M.; Mardirossian, M.; Runti, G.; Benincasa, M. Non-membrane permeabilizing modes of action of antimicrobial peptides on bacteria. Curr. Top. Med. Chem. 2016, 16, 76–88. [Google Scholar] [CrossRef]
  208. Kim, C.; Spano, J.; Park, E.K.; Wi, S. Evidence of pores and thinned lipid bilayers induced in oriented lipid membranes interacting with the antimicrobial peptides, magainin-2 and aurein-3.3. Biochim. Biophys. Acta 2009, 1788, 1482–1496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Mihajlovic, M.; Lazaridis, T. Antimicrobial peptides bind more strongly to membrane pores. Biochim. Biophys. Acta 2010, 1798, 1494–1502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Bozelli, J.C., Jr.; Sasahara, E.T.; Pinto, M.R.; Nakaie, C.R.; Schreier, S. Effect of head group and curvature on binding of the antimicrobial peptide tritrpticin to lipid membranes. Chem. Phys. Lipids 2012, 165, 365–373. [Google Scholar] [CrossRef]
  211. Anunthawan, T.; De la Fuente-Núñez, C.; Hancock, R.E.; Klaynongsruang, S. Cationic amphipathic peptides KT2 and RT2 are taken up into bacterial cells and kill planktonic and biofilm bacteria. Biochim. Biophys. Acta. 2015, 1848, 1352–1358. [Google Scholar] [CrossRef] [Green Version]
  212. Valdés Rodríguez, Y.C.; Bilbao Díaz, M.; León Álvarez, J.L.; Merchán González, F. Origen e importancia de la fosfolipasa A2 de secreción. Rev. Cuba. Farm. 2002, 36, 121–128. [Google Scholar]
  213. Subbalakshmi, C.; Sitaram, N. Mechanism of antimicrobial action of indolicidin. FEMS Microbiol. Lett. 1998, 160, 91–96. [Google Scholar] [CrossRef] [PubMed]
  214. Brodgen, K.A. Antimicrobial Peptides: Pore formers or metabolic inhibitors in bacteria? Nat. Rev. Microbiol. 2005, 3, 238–250. [Google Scholar] [CrossRef]
  215. De Lucca, A.J.; Bland, J.M.; Jacks, T.J.; Grimm, C.; Walsh, T.J. Fungicidal and binding properties of the natural peptides cecropin B and dermaseptin. Med. Mycol. 1998, 36, 291–298. [Google Scholar] [CrossRef] [PubMed]
  216. Jenssen, H.; Hamill, P.; Hancock, R.E.W. Peptide antimicrobial agents. Clin. Microbiol. Rev. 2006, 19, 491–511. [Google Scholar] [CrossRef] [Green Version]
  217. Wachinger, M.; Kleinschmidt, A.; Winder, D.; Von Pechmann, N.; Ludvigsen, A.; Neumann, M.; Holle, R.; Salmons, B.; Erfle, V.; Brack-Werner, R. Antimicrobial peptides melittin and cecropin inhibit replication of human immunodeficiency virus 1 by suppressing viral gene expression. J. Gen. Virol. 1998, 79, 731–740. [Google Scholar] [CrossRef]
  218. Yeaman, M.R.; Yount, N.Y. Mechanisms of antimicrobial peptide action and resistance. Pharmacol. Rev. 2001, 55, 27–55. [Google Scholar] [CrossRef] [Green Version]
  219. Pálffy, R.; Gardlík, R.; Behuliak, M.; Kadasi, L.; Turna, J.; Celec, P. On the physiology and pathophysiology of antimicrobial peptides. Mol. Med. 2009, 15, 51–59. [Google Scholar] [CrossRef]
  220. Diz, M.S.S.; Carvalho, A.O.; Rodrigues, R.; Neves-Ferreira, A.G.C.; Da Cunha, M.; Alves, E.W. Antimicrobial peptides from chili pepper seeds causes yeast plasma membrane permeabilization and inhibits the acidification of the medium by yeast cells. Biochim. Biophys. Acta 2006, 1760, 1323–1332. [Google Scholar] [CrossRef]
  221. Leung, E.H.W.; Ng, T.B. A relatively stable antifungal peptide from buckwheat seeds with antiproliferative activity toward cancer cells. J. Pept. Sci. 2007, 13, 762–767. [Google Scholar] [CrossRef]
  222. Huan, Y.; Kong, Q.; Mou, H.; Yi, H. Antimicrobial Peptides: Classification, Design, Application and Research Progress in Multiple Fields. Front. Microbiol. 2020, 11, 582779. [Google Scholar] [CrossRef]
  223. Giordano, D.; Costantini, M.; Coppola, D.; Lauritano, C.; Núñez Pons, L.; Ruocco, N.; Verde, C. Biotechnological Applications of Bioactive Peptides from Marine Sources. Adv. Microb. Physiol. 2018, 73, 171–220. [Google Scholar] [CrossRef] [PubMed]
  224. Lu, Y.; Ma, Y.F.; Wang, X.; Liang, J.G.; Zhang, C.X.; Zhang, K.Y.; Lin, G.O.; Lai, R. The first antimicrobial peptide from sea amphibian. Mol. Immunol. 2008, 45, 678–681. [Google Scholar] [CrossRef] [PubMed]
  225. Sable, R.; Parajuli, P.; Jois, S. Peptides, Peptidomimetics, and Polypeptides from Marine Sources: A Wealth of Natural Sources for Pharmaceutical Applications. Mar. Drugs 2017, 15, 124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Jemil, I.; Abdelhedi, O.; Nasri, R.; Mora, L.; Jridi, M.; Aristoy, M.-C.; Toldrá, F.; Nasri, M. Novel bioactive peptides from enzymatic hydrolysate of Sardinelle (Sardinella aurita) muscle proteins hydrolysed by Bacillus subtilis A26 proteases. Food Res. Int. 2017, 100, 121–133. [Google Scholar] [CrossRef]
  227. Anaya, K.; Podszun, M.; Franco, O.L.; De Almeida Gadelha, C.A.; Frank, J. The Coconut Water Antimicrobial Peptide CnAMP1 Is Taken up into Intestinal Cells but Does Not Alter P-Glycoprotein Expression and Activity. Plant Foods Hum. Nutr. 2020, 75, 396–403. [Google Scholar] [CrossRef]
  228. Tang, S.-S.; Prodhan, Z.H.; Biswas, S.K.; Le, C.-F.; Sekaran, S.D. Antimicrobial peptides from different plant sources: Isolation, characterisation, and purification. Phytochemistry 2018, 154, 94–105. [Google Scholar] [CrossRef]
  229. Ennaas, N.; Hammami, R.; Beaulieu, L.; Fliss, I. Purification and characterizationof four antibacterial peptides from protamex hydrolysate of Atlantic mackerel (Scomber scombrus) by-products. Biochem. Biophys. Res. Commun. 2015, 462, 195–200. [Google Scholar] [CrossRef]
  230. Tang, W.; Zhang, H.; Wang, L.; Qian, H.; Qi, X. Targeted separation of antibacterial peptide from protein hydrolysate of anchovy cooking wastewater by equilibrium dialysis. Food Chem. 2015, 168, 115–123. [Google Scholar] [CrossRef]
  231. Rozek, T.; Bowie, J.H.; Wallace, J.C.; Tyler, M.J. The antibiotic and anticancer active aurein peptides from the Australian Bell Frogs Litoria aurea and Litoria raniformis. Part 2. Sequence determination using electrospray mass spectrometry. Rapid Commun. Mass Spectrom. 2000, 14, 2002–2011. [Google Scholar] [CrossRef]
  232. Rozek, T.; Wegener, K.L.; Bowie, J.H.; Olver, I.N.; Carver, J.A.; Wallace, J.C.; Tyler, M.J. The antibiotic and anticancer active aurein peptides from the Australian Bell Frogs Litoria aurea and Litoria raniformis. Eur. J. Biochem. 2000, 267, 5330–5341. [Google Scholar] [CrossRef]
  233. Da Mata, É.C.G.; Mourão, C.B.F.; Rangel, M.; Schwartz, E.F. Antiviral activity of animal venom peptides and related compounds. J. Venom. Anim. Toxins Incl. Trop. Dis. 2017, 23, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Niidome, T.; Kobayashi, K.; Arakawa, H.; Hatakeyama, T.; Aoyagi, H. Structure–activity relationship of an antibacterial peptide, maculatin 1.1, from the skin glands of the tree frog, Litoria genimaculata. J. Pept. Sci. 2004, 10, 414–422. [Google Scholar] [CrossRef] [PubMed]
  235. Sikorska, E.; Greber, K.; Rodziewicz-Motowidło, S.; Szultka, Ł.; Łukasiak, J.; Kamysz, W. Synthesis and antimicrobial activity of truncated fragments and analogs of citropin 1.1: The solution structure of the SDS micelle-bound citropin-like peptides. J. Struct. Biol. 2009, 168, 250–258. [Google Scholar] [CrossRef] [PubMed]
  236. Kościuczuk, E.M.; Lisowski, P.; Jarczak, J.; Strzałkowska, N.; Jóźwik, A.; Horbańczuk, J.; Krzyżewski, J.; Zwierzchowski, L.; Bagnicka, E. Cathelicidins: Family of antimicrobial peptides: A review. Mol. Biol. Rep. 2012, 39, 10957–10970. [Google Scholar] [CrossRef] [Green Version]
  237. Tincu, J.A.; Taylor, S.W. Antimicrobial peptides from marine invertebrates. Antimicrob. Agents Chemother. 2004, 48, 3645–3654. [Google Scholar] [CrossRef] [Green Version]
  238. Zasloff, M. Antimicrobial peptides of multicellular organisms. Nature 2002, 415, 389–395. [Google Scholar] [CrossRef]
  239. Yang, D.; Biragyn, A.; Kwak, L.W.; Oppenheim, J.J. Mammalian defensins in immunity: More than just microbicidal. Trends Immunol. 2002, 23, 291–296. [Google Scholar] [CrossRef]
  240. Ganz, T. Defensins: Antimicrobial peptides of innate immunity. Nat. Rev. Immunol. 2003, 3, 710–720. [Google Scholar] [CrossRef]
  241. Abidi, F.; Aissaoui, N.; Gaudin, J.C.; Chobert, J.M.; Haertl, T.; Marzouki, M.N. MS analysis and molecular characterization of Botrytis cinerea protease Prot-2. Use in bioactive peptides production. Appl. Biochem. Biotechnol. 2013, 170, 231–247. [Google Scholar] [CrossRef]
  242. Cao, J.; De la Fuente-Nunez, C.; Ou, W.R.; Torres, T.M.D.; Pande, G.S.; Sinskey, J.A.; Lu, T.K. Yeast based synthetic biology platform for antimicrobial peptide production. ACS Synth. Biol. 2018, 7, 896–902. [Google Scholar] [CrossRef]
  243. Martínez, O.; Martínez, E. Proteínas y péptidos en nutrición enteral. Nutr. Hosp. 2006, 21, 1–14. [Google Scholar]
  244. Shimizu, M. Food-derived peptides and intestinal functions. Biofactors 2004, 21, 43–47. [Google Scholar] [CrossRef] [PubMed]
  245. Suárez-Santiago, E. Péptidos Antihipertensivos de Proteínas de Amaranto. Estudio de Emulsiones y Galletitas Como Sistemas Funcionales. Ph.D. Thesis, Facultad de Ciencias Exactas, Centro de Investigación y Desarrollo en Criotecnología de Alimentos, La Plata, Argentina, 2018. [Google Scholar]
  246. Hernández-Ledesma, B.; Miguel, M.; Amigo, L.; Aleixandre, M.A.; Recio, I. Effect of simulated gastrointestinal digestion on the antihypertensive properties of synthetic β-lactoglobulin peptide sequences. J. Dairy Res. 2007, 74, 336–339. [Google Scholar] [CrossRef] [PubMed]
  247. Gardner, M.L.G. Transmucosal passage on intact peptides in mamalian metabolism. In Tissue Utilization and Clinical Targeting; Grimble, G.K., Backwell, F.R.G., Eds.; Portland Press Ltd.: London, UK, 1998. [Google Scholar] [CrossRef] [Green Version]
  248. Ziv, E.; Bendayan, M. Intestinal absorption of peptides through the enterocytes. Microsc. Res. Technol. 2000, 49, 346–352. [Google Scholar] [CrossRef]
  249. Rubas, W.; Grass, G.M. Gastrointestinal lymphatic absorption of peptides and proteins. Adv. Drug Deliv. Rev. 1991, 7, 15–69. [Google Scholar] [CrossRef]
  250. Shimizu, M.; Tsunogai, M.; Arai, S. Transepithelial transport of oligopeptides in the human intestinal cell, Caco-2. Peptides 1997, 18, 681–687. [Google Scholar] [CrossRef]
  251. Grimble, G.K. The significance of peptides in clinical nutrition. Annu. Rev. Nutr. 1994, 14, 419–447. [Google Scholar] [CrossRef]
  252. Roberts, P.R.; Burney, J.D.; Black, K.W.; Zaloga, G.P. Effect of chain length on absorption of biologically active peptides from the gastrointestinal tract. Digestion 1999, 60, 332–337. [Google Scholar] [CrossRef]
  253. Vilcacundo, R.; Barrio, D.; Carpio, C.; García-Ruiz, A.; Rúales, J.; Hernández-Ledesma, B.; Carrillo, W. Digestibility of Quinoa (Chenopodium quinoa Willd.) Protein Concentrate and Its Potential to Inhibit Lipid Peroxidation in the Zebrafish Larvae Model. Plant Foods Hum. Nutr. 2017, 72, 294–300. [Google Scholar] [CrossRef]
  254. Carrillo, W.; Gómez-Ruiz, J.A.; Miralles, B. Identification of antioxidant peptides of hen egg-white lysozyme and evaluation of inhibition of lipid peroxidation and cytotoxicity in the Zebrafish model. Eur Food Res Technol. 2016, 242, 1777–1785. [Google Scholar] [CrossRef] [Green Version]
  255. Horner, K.; Drummond, E.; Brennan, L. Bioavailability of milk protein-derived bioactive peptides: A glycaemic management perspective. Nutr. Res. Rev. 2016, 29, 91–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Diagram of the bioactive peptide production process [68,69,70].
Figure 1. Diagram of the bioactive peptide production process [68,69,70].
Molecules 27 01343 g001
Figure 2. Schematic representation of the chemical and physical mechanisms of antioxidant peptides, used to inhibit oxidative processes: (1) chelators of metals, (2) radical inhibitors, and (3) physical shielding [194].
Figure 2. Schematic representation of the chemical and physical mechanisms of antioxidant peptides, used to inhibit oxidative processes: (1) chelators of metals, (2) radical inhibitors, and (3) physical shielding [194].
Molecules 27 01343 g002
Figure 3. Diagram of protein digestion and transepithelial transport of peptides until they reach circulation [245].
Figure 3. Diagram of protein digestion and transepithelial transport of peptides until they reach circulation [245].
Molecules 27 01343 g003
Table 1. Bioactive peptide sources from various foods and their activity.
Table 1. Bioactive peptide sources from various foods and their activity.
ActivitySourceProtein of OriginBioactive Peptide o Sequence
Inhibition of angiotensin-converting enzyme (ACE) and antihypertensiveSoySoy proteinNMGPLV
FishMuscle proteinLKP, IKP (derived from sardines, mackerel, tuna, squid)
MeatMuscle proteinIKW, LKP
Milkα-LA, β-LG
α-, β, ƙ-CN
Lactokinins (WLAHK, LRP, LKP)
EggOvotransferrin
Ovalbumin
Ovokinin (FRADHPPL), Ovokinin (2–7) (KVREGTTY)
WheatGliadinsCasokinins (FFVAP, FALPQY, VVP)
BroccoliEncrypted proteinIAP
Chicken skinCollagenYPK
Chicken legsCollagenGAHpGLHpGP
ImmunomodulatorRiceRice albuminOrizatensinin (GYPMYPLR)
EggOvalbuminUnspecified peptides
Milkα-β-ƙ-CN α-LAImmunopeptides (αs 1 TTMPLW)
WheatGlutenImmunopeptides
CytomodulatorMilkα-β-CNA-casomorphin (HIQKED (V)),
β- casomorphin-7 (YPFPGP)
Opioid agonistMilkα-LA, β-LG
α-, β-CN
α-lactorfin, β- lactorfin
Casomorphin
WheatGlutenGluten exorphins A4, A5 (GYYPT),
B4, B5 and C (YPISL)
Opioid antagonistMilkLactoferrin
ƙ-CN
Lactoferricin
Caxosines
AntimicrobialEggOvotransferrinOTAP-92 (f109-200)
MilkLysozyme
α-β-ƙ-CN
Unspecified peptides
Bovine cruorLactoferrinLactoferricin
Caxosines
TSKYR
Guava seedsGlycinePg-AMP1
Vicia faba seedsHydrolysate seed proteinsLSPGDVLVIPAGYPVAIK, EEYDEEKEQGEEEIR
Tomato pomaceHydrolysate proteinsUnspecified peptides
Rice branRice bran proteinsKVDHFPL
Bovine hemoglobineHemoglobine(F)VNFKLLSHSLL, (L)TSKYR, (F)KLLSHSL, (L)QADFQKVVAGVANALAHRYH, MLTAEEKAAVTAFWGKVKVDEVGGEALGRL
AntithromboticMilkƙ-CN
(glycomacropeptide)
ƙ-CN (f106-116) a, casoplatelin
Mineral-carrying
Anticarcinogenic
Milkƙ-CN
(glycomacropeptide)
ƙ-CN (f106-116) a, casoplatelin
HypocholesterolemicSoyGlycinin/conglycinin LPYPR
Milkβ-LGIIAEK
AntioxidantSardineSardine muscleMY
WheatWheat germ proteinUnspecified peptides
Milkα-LA, β-LGMHIRL, YVEEL, WYSLAMAASDI
TunaTuna bonesVKAGFAWTANQQLS
OysterOyster by-productsPVMGD
LeatherjacketLeatherjacket headsEHGV
SalmonSalmon by-productsWEGPK, GPP, GVPLT
SoyGlycinin /conglycinin LLPHH, VNHDHQN, LVNHDHQN, LLPHH
Rice endospermRice proteinFRDEHKK, KHDRGDEF
CN: casein; LA: lactoalbumin; LG: lactoglobulin; f: fragment; A: alanine; R: arginine; N: asparagine; D: aspartic acid; C: cysteine; E: glutamic acid; Q: glutamine; G: glycine; H: histidine; Hp: hydroxyproline; I: isoleucine; L: leucine; K: lysine; M: methionine; F: phenylalanine; P: proline; S: serine; T: threonine; W: tryptophan; Y: tyrosine; V: valine [30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51].
Table 2. Biopeptide databases (all web addresses were verified on 28 January 2022).
Table 2. Biopeptide databases (all web addresses were verified on 28 January 2022).
DatabaseWeb AddressContentReference
AHTPDB *http://crdd.osdd.net/raghava/ahtpdb/Antihypertensive peptides[136]
AntiTbPdbhttp://webs.iiitd.edu.in/raghava/antitbpdbAntitubercular and mycobacterial peptides[137]
APDhttp://aps.unmc.edu/AP/main.htmlAntimicrobial and anticancer peptides[138]
AVPdbhttp://crdd.osdd.net/servers/avpdb/Antiviral peptides[139]
BaAMPshttp://www.baamps.it/Antimicrobial peptides tested against microbial
films
[140]
BactPepDBhttp://bactpepdb.rpbs.univ-paris-diderot.fr/cgi-bin/home.plBacterial peptides[141]
BIOPEP-UWMTM *http://www.uwm.edu.pl/biochemiaBioactive peptides/sensory peptides and amino
acids
[142]
Brainpepshttp://brainpeps.ugent.be/Blood–brain barrier passing peptides[143]
CAMPR3http://www.camp.bicnirrh.res.in/Antimicrobial peptides[144]
CancerPPDhttp://crdd.osdd.net/raghava/cancerppd/index.phpAnticancer peptides and proteins[145]
CPPSite 2.0http://crdd.osdd.net/raghava/cppsite/Cell-penetrating peptides[146]
DBAASPhttps://dbaasp.org/Antimicrobial peptides[147]
EROP-Moscowhttp://erop.inbi.ras.ru/Bioactive peptides[148]
Hemolytikhttp://crdd.osdd.net/raghava/hemolytik/Hemolytic and non-hemolytic peptides[149]
MBPDB *http://mbpdb.nws.oregonstate.edu/Milk protein-derived bioactive peptides[150]
NeuroPephttp://isyslab.info/NeuroPep/Neuropeptides[151]
PepBankhttp://pepbank.mgh.harvard.edu/Bioactive peptides[152]
Quorumpepshttp://quorumpeps.ugent.be/Quorum sensing signaling peptides[153]
SATPdbhttp://crdd.osdd.net/raghava/satpdb/links.phpA metabase of therapeutic peptides[154]
StraPephttp://isyslab.info/StraPep/Structures of bioactive peptides[155]
THPdbhttp://crdd.osdd.net/raghava/thpdb/index.htmlFDA-approved therapeutic peptides[137]
TumorHoPehttp://crdd.osdd.net/raghava/tumorhope/Tumor homing peptides[156]
YADAMPhttp://yadamp.unisa.it/about.aspxAntimicrobial peptides[157]
* Databases containing information about peptides derived from food sources were marked with an asterisk.
Table 3. Peptide fractions with antioxidant activity.
Table 3. Peptide fractions with antioxidant activity.
SourceCharacteristicsProcurement and IdentificationActivityReference
Muscle of miiuy
corvina (Miichthis
miiuy)
YASVV, NFWWP, FWKVV, TWKVV, FMPLH, YFLWP, VIAPW, WVWWW, MWKVW and IRWWWUF/RP-HPLCRadical scavenging of DPPH and ABTS.[172]
Finger milletTSSSLNMAVRGGLTR and
STTVGLGISMRSASVR
Trypsin, pepsin
MALDI-TOF/TOF–MS/MS
Capture of hydroxyl radicals, DPPH, ABTS and chelating activity.[173]
Salmon jellyPP, GF, GPVA, GGPAGPAV, R,YAlcalase, Flavourzyme 500 L, Corolase PP,
Promod
RP-HPLC/UPLC–MS/MS
Inhibitor of dipeptidyl peptidase IV (DPP-IV) and reactive oxygen species
(ORAC).
[174]
Shells from shrimp
discard
processing
SYELPDGQVITIGNER, YPIEHGIITNWDDMEK, EEYDESGPGIVHR, EVDRLEDELVNEK, ALSNAEGEVAALNR, NLNDEIAHQDELINK, LEQTLDELEDSLERTrypsin, α-chymotrypsin, pepsin
GPC
NANO LC–LTQ MS
ABTS, DPPH and hydroxyl radical scavenging, reducing power and chelating capacity of ferrous ions, inhibition of β-carotene bleaching,
peroxidation of cholesterol and peroxyl and hydroxyl radicals.
[175]
Salmon trimmingsGPAV, VC Y FFAlcalase, Flavourzyme 500 L, Corolase PP, Promod
RP-HPLC/
UPLC–MS/MS
Inhibitor of dipeptidyl peptidase IV (DPP-IV) and reactive oxygen species
(ORAC).
[176]
Donkey milkEWFTFLKEAGQGAKDMWR, GQGAKDMWR,
REWFTFLK,
MPFLKSPIVPF
MDLC
micro-HPLC-Orbitrap–MS
Antioxidant capacity and inhibitory activity of angiotensin-converting enzyme.[177]
Goat milkSerum: 883.47–1697.82 Da
Casein: 794.44–1956.95 Da
Presence of P and H residues.
Pepsin
HPLC
RP-HPLC
DPPH and superoxide radical scavenging activity.[178]
Palm kernel<3 kDa
VVG-G-D-G-D-V
VPVTST
LTTLDSE
Pepsin, pancreatin.
UF
RP-HPLC
Radical scavenging activity
ABTS.
Iron-reducing power.
[179]
Cod muscleMW fractions < 1 kDa Pepsin, trypsin, chymotrypsin
RP-HPLC
Capture of oxygen radicals and DPPH.
Superoxide and hydroxyl scavenging activity.
Chelating activity of iron.
Inhibition of the oxidation of linoleic acid
Dose-dependent.
[171]
Juglans
sigillata
seeds
Peptides with 2–4 aa residues rich in Y, CPancreatin
GPC
RP-HPLC
Radical scavenging activity DPPH, ABTS, oxygen.
Iron chelation.
Protective activity in PC12 cells against H2O2-induced cytotoxicity.
[180]
Egg white powderDHTKE
628.64 Da
FFEFH
630.71 Da
MPDAHL
684.1 Da
Alcalase
UF
GPC
628.64 Da.
Better oxygen radical absorption capacity.
630.71 and 684.1 Da.
DPPH radical scavenging activity.
Acidic, hydrophobic and low-molecular weight peptides showed higher antioxidant activity.
[181]
SquidWCTSVS
682.5 Da
α-chymotrypsin
IC
GPC
Free radical scavenging activity (DPPH, hydroxyl, superoxide). Chelation of
metals (Fe). Avoids DNA damage. Inhibits lipid peroxidation (linoleic acid).
[182]
Campana buffalo
mozzarella cheese
CKYVCTCKMS
1326.5 Da
Gastrointestinal digestion in vitro
GPC
HPLC
Gut protection against induced oxidative stress (CaCo2 cells).[183]
Bovine hairCERPTCCEHS
1325.4 Da
Alcalase
GPC
SEC
ABTS and hydroxyl radical scavenging activity.
Inhibition of erythrocyte hemolysis and lipid peroxidation.
Protection of DNA and PC12 cells against hydrogen peroxide-induced oxidative damage.
[184]
Plum (Prunus domestica L.)MLPSLPK, HLPLL, NLPLL, HNLPLL, KGVL, HLPLLR, HGVLQ, GLYSPH, LVRVQ, YLSF, DQVPR, LPLLR, VKPVAPF.High-intensity focused ultrasound.
Alcalase
RP-HPLC-ESI-Q-TOF
Antioxidant and antihypertensive activity.[185]
Velvet bean (Mucuna
pruriens)
<1 kDa, 1–3 kDaSequential hydrolysis
Alcalase–Flavourzyme, pepsin–pancreatin
UF
Radical scavenging activity DPPH.
Iron-reducing power.
[186]
Peanut seedTPA (286 kDa)
I/LPS (315 kDa)
SP (202 kDa)
Alcalase
UF
GPC
HPLC
Peptides with MW < 3 kDa showed greater reducing power than those with PM > 3 kDa.[187]
Duck egg (egg white)202.1
294.1
382.1
426.3
514.4 Da
Sequential hydrolysis (alcalase and specific hydrolase for egg protein (SEEP))
GPC
IC
Oxygen and hydroxyl radical absorption capacity.[188]
Palm oil extraction residueAWFS 509.56 Da
WAF 422.48 Da
LPWRPATNVF
1200.41 Da
Papain Fractionation based on isoelectric point
RP-HPLC
Radical scavenging activity of DPPH.
Iron chelation.
[189]
Sweet potatoYYIVS 643.2 Da
TYQTF 659.4 Da
SGQYFL
713.2 Da
YYDPL 669.3 Da
Alcalase
UF
RP-HPLC
Hydroxyl radical scavenging activity.[110]
Bovine plasma Alcalase
UF
IC
RP-HPLC
Free radical scavenging capacity.
High reductive power.
[190]
CN: casein; LA: lactoalbumin; LG: lactoglobulin; f: fragment; A: alanine; R: arginine; N: asparagine; D: aspartic acid; C: cysteine; E: glutamic acid; Q: glutamine; G: glycine; H: histidine; Hp: hydroxyproline; I: isoleucine; L: leucine; K: lysine; M: methionine; F: phenylalanine; P: proline; S: serine; T: threonine; W: tryptophan; Y: tyrosine; V: valine; UF: ultrafiltration; RP: reversed-phase; HPLC: high-performance liquid chromatography; MALDI: matrix-assisted laser desorption/ionization; TOF: time of flight; MS: mass spectrometry; UPLC: ultra-performance liquid chromatography; GPC: gel filtration chromatography; LTQ: linear ion trap mass spectrometer; MDLC: multidimensional liquid chromatography; IC: ion-exchange chromatography; SEC: high resolution size exclusion chromatography; ESI: electrospray ionization; Q: quadrupole.
Table 4. Relationship between antioxidant activity, amino acids, and mechanism of action.
Table 4. Relationship between antioxidant activity, amino acids, and mechanism of action.
Amino AcidMechanism of ActionExampleReference
CysteineSH groups are radical scavengers, protect tissues from oxidative stress and improve glutathione
peroxidase activity.
Tripeptides with C.[197,199]
Hydrophobic amino acidsIncrease solubility of peptides in lipids, facilitating access to hydrophobic radical species and polyunsaturated fatty acids.P, H or T, within sequences and V or L at N-terminus in peptides.
Terminal amino acids such as L or V and G and P residues in gluten peptide sequences.
[77,162,192]
Acidic and basic amino acidsThe carboxyl and amino groups
of the side chains chelate metal ions due to their ability to dissociate and be proton donors.
A (acidic amino acid) and H (basic amino acid).[103]
Aromatic amino acids (Y, H, W, and F)They stabilize radicals by
donating electrons, maintaining their own stability via the resonance
of their structure.
H at N-terminus.
H at C-terminus.
Tripeptides with W or Y at C-terminus.
[162]
Table 5. Peptide fractions with antimicrobial activity.
Table 5. Peptide fractions with antimicrobial activity.
SourceAntimicrobial PeptideReference
Green coconut water (Cocos nucifera
L).
CnAMP1,
CnAMP2
CnAMP3
[227]
Stems, seeds, and leaves of plantsThionins
Defensins
Snakins
[228]
Sardinella prepared by treatment with Bacillus subtilis A26
protease
Sardinella protein hydrolysate (SPH),[226]
Atlantic mackerel (Scomber scombrus)
Protamex protein hydrolysate
SIFIQRFTT, RKSGDPLGR,
AKPGDGAGSGPR
GLPGPLGPAGPK
[229]
Protein hydrolysate of anchovy
cooking wastewater
GLSRLFTALK[230]
FrogsAurein 1–2
Brevinin 1
Maculatins
Citropin
[231,232,233,234,235]
InsectsCecropin[236]
Horse CrabTachyplesins
Polyphemusin
[237,238]
PigsProtegrins
Tritrpticin
[236]
BovineCathelicidin
Indolicidin
[236]
CowBactenecin 1
b-defensin
Bac 5
Indolicidin
[238]
Mammalsα defensins
β defensins
[239,240]
HoneybeeApidaecin[238]
Trypsin digested Botrytis
cinerea protease
FPGSAD,
SCVGTDLNR,
VAHLT
VQGGDAYYLNR
SGSTASAVGASLCR
[241]
Lactococcus lactis, Bacillus subtilis, and Bacillus brevisNisin
Gramicidin
[242]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

López-García, G.; Dublan-García, O.; Arizmendi-Cotero, D.; Gómez Oliván, L.M. Antioxidant and Antimicrobial Peptides Derived from Food Proteins. Molecules 2022, 27, 1343. https://doi.org/10.3390/molecules27041343

AMA Style

López-García G, Dublan-García O, Arizmendi-Cotero D, Gómez Oliván LM. Antioxidant and Antimicrobial Peptides Derived from Food Proteins. Molecules. 2022; 27(4):1343. https://doi.org/10.3390/molecules27041343

Chicago/Turabian Style

López-García, Guadalupe, Octavio Dublan-García, Daniel Arizmendi-Cotero, and Leobardo Manuel Gómez Oliván. 2022. "Antioxidant and Antimicrobial Peptides Derived from Food Proteins" Molecules 27, no. 4: 1343. https://doi.org/10.3390/molecules27041343

Article Metrics

Back to TopTop