Open access peer-reviewed chapter

The Hydrolysates from Fish By-Product, An Opportunity Increasing

Written By

Jose Edgar Zapata Montoya and Angie Franco Sanchez

Submitted: 19 November 2021 Reviewed: 22 December 2021 Published: 14 February 2022

DOI: 10.5772/intechopen.102348

From the Edited Volume

Hydrolases

Edited by Sajjad Haider, Adnan Haider and Angel Catalá

Chapter metrics overview

297 Chapter Downloads

View Full Metrics

Abstract

The fishery industries have continuously increased over the last decade. This growth comes accompanied by a high volume of by-products released to environment, because these industries discard between 60 and 70% of their production as waste. This waste includes fish whole or part from these such as fillet remains (15–20%), skin and fins (1–3%), bones (9–15%), heads (9–12%), viscera (12–18%) and scales (5%). This by-products are rich in proteins and lipids which of several nature, which can be recovered to obtain compounds of high added value. In this chapter, some methods to recover compounds from fish by-products will be discussed. Among others, will be discussed topics about postharvest of fish, by-product releasing, enzymatic hydrolysis of by-product and bioactive peptide obtaining from fish waste.

Keywords

  • bioactives peptides
  • enzymatic hydrolysis
  • protein revaluation

1. Introduction

Production levels of fishery and aquaculture have been increasing for the last 30 years, as fish is an important protein source for human consumption and it is expected to reach a production of 196 mt by 2025 [1]. As a result, more and more people depend on fish or other fisheries production, capture, processing and marketing. By 2018, aquaculture production in the world was estimated to reach over 178 million tons [2], whereas marine capture fisheries have been around half the global production [3].

A huge waste volume has been produced along with that production increase, too. Around 70% of fish is processed before final sale, producing between 20 and 80% of fish waste, depending on the fish type and its transformation technology [4]. Furthermore, important amounts of water are required for those processes [5]. That situation represents a challenge from an environmental perspective because around 50% of that fish waste is discarded without being used [6]. Most of it is buried or deposited in water sources, either in the ocean, rivers, or streams. In the case of landfills, it can lead to saturations that cause odor and leachate problems. As for dumping in water sources, aerobic bacteria use organic matter by the action of oxygen, releasing large amounts of phosphorus, nitrogen and ammonium, affecting pH, causing algae growth, and turbidity. The absence of oxygen in water results in the release of hydrogen sulfide, carbon dioxide, organic acids, methane, and ammonium [7].

These wastes contain important nutrient levels [8] and their composition depends on species, source organs or obtaining processes, as seen in Table 1. On the other hand, some of those nutrients represent an opportunity from an economic perspective, as in the case of the protein, which can be recovered to obtain high added-value compounds.

ProcessOrganic byproduct%Goal
StunningN/ADecrease agony time to reduce undesirable compound production
ClassificationWhole fishSeparate by size or species
Slime removalAqueous fluidReduce microbial contamination surface
ScalingScales5Reduce bacterial contamination
WashingWashing water100Remove micro-organisms and contaminants
Head removalHeads9–32Remove non-edible or low-value parts
EviscerationViscera12–18Remove internal organs to reduce microbial contamination
Fin CuttingFins1–2Remove non-edible parts
SkinningSkin3Remove non-edible parts
FilletingFillet remains15–20Separation of dorsal and abdominal meat from fish
Bone/meat separationBones and skeletons9–15Separate meat from ribs and bones

Table 1.

Processes used for fish preparation after capture.

Among the methods used to add value to fish residues, there are protein hydrolysis, silage, and collagen recovery [9]. In the first hydrolysis tests evaluated, chemical processes and extraction with organic solvents were used, showing that they affected the nutritional quality of proteins and amino acids. For this reason, commercial enzymes have been increasingly applied to intend to obtain hydrolyzed protein of this substrate type [10]. These latter processes have moderate operating conditions, show greater reproducibility, and are more controllable and selective than chemical processes. Besides, they deliver products with techno-functional properties, excellent digestibility, rapid absorption, and amino acid balance, in addition to high levels of bioactive peptides [11].

This chapter will address the issue of protein residues used in fish processing aiming to obtain bioactive peptides through enzymatic hydrolysis using commercial enzymes. The basic concepts of fish processing, the characteristics of the waste generated, their use by enzymatic hydrolysis, and bioactive and functional peptide production will be addressed.

Advertisement

2. Fish post-harvest

Once the fish is harvested, it undergoes different processes intending to improve conservation conditions, separate the non-edible or low commercial value parts, and leave the product ready to deliver to the consumer. Table 1 lists, in general terms, the stages of fish processing, many of which release some type of organic by-product [3, 6].

Advertisement

3. Bromatological characteristics of the Main fish-farmed by-products

Fish by-products are made up of different compound types with food importance [12]. The major components are moisture, fat, and protein. However, the bromatological composition varies depending on the species, age, and gender of the fish, in addition to the part of the fish from which the by-product comes, or the processes to which it has been subjected [13]. Thus, Table 2 presents the bromatological composition of different fish by-products, for different species, fish parts, and processes.

Type of wasteProteinFatMoistureAshReference
Freeze-dried Viscera of Yamú (Brycon siebenthalae)19.1979.490.48[14]
Argentine hake (Merluccius hubbsi) gonad21.9510.9268.7211.61[15]
Raw Viscera Tilapia of (Oreochromis spp.)4.0332.9361.360.67[16]
Tilapia (Oreochromis nilotica) skeleton (D.B)50.630.665.315.3[17]
Raw Viscera of Trucha (Oncorhynchus mykiss)9.1431.1556.931.51[10]
Argentine hake (M. hubbsi) liver16.3829.7155.791.61[15]
Viscera of Catla Catla8.5212.4676.252.50[18]
Cape hake (Merluccius capensis) by products18.01.178.51.9[19]
Tilapia (Oreochromis spp.) scales (D.B)67.9615.1832.08[20]
Greenland halibut (Reinhardtius hippoglossoides) skin15.9510.6255.4417.63[21]
Tilapia (Oreochromis spp.) viscera7.8726.08621.19[22]
Blue shark (Prionace glauca) skin22.790.2476.034.24[21]
Rainbow trout (O. mykiss) viscera151371.72.7[23]
Atlantic salmon (Salmo salar) viscera844601[24]
Tilapia (Oreochromis spp.) defatted viscera10.041.8883.211.71[22]
Yellowfin tuna (Thunnus albacares) skin32.383.220.6762.57[21]
Red tilapia (Oreochromis niloticus) head, skeleton, and tail14.65.566.68.9[25]
Tilapia del Nilo (O. niloticus) skin29.6813.8954.911.61[26]
Black Sea anchovy (Engraulis encrasicholus) head13.3910.0270.945.00[27]
Atlantic salmon (S. salar) head (D.B.)13223944[24]
Tilapia (Oreochromis spp.) spines (D.B.)55.5453.4622.91[20]
Tilapia (Oreochromis nilótica) skeleton50.630.665.315.3[17]
Black Sea anchovy (Engraulis encrasicholus) frame16.4715.5059.727.60[27]
Tilapia (Oreochromis spp.) viscera4.57433.60262.6930.732[28]
Atlantic salmon (S. salar) skeleton1527424[24]
Tilapia (Oreochromis spp.) defatted viscera12.6442.52582.6071.462[28]
Black Sea anchovy (Engraulis encrasicholus) viscera12.0523.9061.502.09[27]
Cuttlefish (Sepia officinalis) viscera17.454.7874.991.95[29]
Salmon (S. salar) head1322394[24]
Tilapia (Oreochromis spp.) scales83.90.915.1[30]
Salmon (S. salar) skeleton1527424[24]

Table 2.

Bromatological composition of fish by-products D.B.: Dry base.

As Table 2 shows, these residues contain mainly proteins, fats and water, but they may also contain high added-value compounds such as collagen and gelatin, polyunsaturated fatty acids (EPA and DHA), monounsaturated such as palmitic and oleic, in addition to minerals and enzymes such as pepsin, trypsin, chymotrypsin and collagenase [3]. Because of their nutrient richness, inappropriate dumping of these residues affects not only the area where they are directly discharged, but it can also alter natural ecosystems in a wider area. In this sense, phosphorus and dissolved nitrogen release can be favored and thus increase biochemical demand for oxygen (BDO), because at least 80% of the nutrients in fish residues are potentially eutrophic substances. This leads to the higher growth of macroalgae in aquifers [31].

In some regions of the world, alternatives to use by-products have been sought. That is how the demand for complete fish heads and skeletons as food for humans has increased in Asia and Africa [32]. Bones, which contain the highest protein levels among the residues (41–84%), are a good source of collagen and gelatin. Besides, their mineral content has been used in the manufacture of food products for schoolchildren (85 mg/kg zinc, 350 mg/kg iron, and 84 g/kg calcium) [32]. Whereas skeletons contain significant amounts of meat remaining after filleting, whose protein is highly digestible and can be extracted for different purposes since it has better nutritional properties than plant proteins, and better essential amino acids balance than other animal protein sources [33] but are more sensitive to heat [34]. On the other hand, fish skin, provides gelatin [32], such as, Nile Tilapia skin has been used to produce collagen [35], which can be used for tissue regeneration [36].

A fish by-product that has gained the most attention in recent years is the viscera (12–20% of the fish), which comprise all organs of the main body cavity, including gills, heart, liver, spleen, swim bladder, stomach, gonads, intestines and their contents [6]. This residue has an average composition of 8–21% proteins, 2–12% lipids, 60–81% humidity and 1–5% ash [6]. The high protein content, in addition to being an excellent enzyme source, makes them a potential source of added-value products with exceptional properties for different industrial applications [37].

Between 70 and 80% of fish muscle is a structural protein, between 20 and 30% sarcoplasma proteins, and the remaining 2–3% of proteins are insoluble connective tissue. The main food protein is myofibrillar, with 66–77% of the total in fish meat. This protein comprises between 50 and 60% myosin and 15–30% actin [38]. Myosin fibers are connected by actin molecules and can be cut at one end by trypsin and chymotrypsin, while at the other end by papain, to form they divide into two forms of meromyosin, heavy and light, with different functional properties [39].

Fish proteins contain between 16 and 18 amino acids, which have an excellent balance, usually 8 essential and 8 non-essential. This makes this type of protein very widely used for animal feed, although they are also used for fertilizer production, silage and in recent decades, bioactive peptide production [30, 40]. Table 3 shows the aminograms of different residues of several fish species, some raw and others that have undergone hydrolysis processes [14], atomization drying [40] and membrane fractionation [11].

Amino AcidRed tilapiaMackerel fishYamú viscera
RTVHFRTVHWTWMDMPIDH9DH28
Histidine4.061.994.53.85.26.6295.0695.222
Isoleucine2.532.445.565.64.9194.0735.221
Leucine7.998.149.4108.85.4455.2545.267
Lysine7.689.917.987.63.543.332.437
Methionine1.320.142.74.62.80.9441.0180.656
Arginine3.974.447.65.97.1
Valine5.274.437.86.88.51.1081.9010.874
Phenylalanine0.911.174.243.12.4071.8982.406
Threonine8.046.065.75.55.51.2281.9271.898
Tryptophan
Ac Aspartic2.314.3911.812.211.41.8371.7992.135
Ac Glutamic6.65.8415.81815.64.3294.7975.045
Asparagine2.314.391.0540.9840.627
Serine4.263.944.55.24.13.4363.723.398
Glycine21.3817.6266.24.61.5161.8821.508
Alanine3.83.827.77.37.32.4983.182.89
Tyrosine3.114.233.53.93.417.5417.6613.824
Cystine3.2374.3023.706
Hydroxyproline
Proline1.54.61.5
Glutamine6.65.84
Total92.1488.79106.1112102.161.6662.7957.114
Reference[11][41][14]

Table 3.

Amino acids content of fish y-products.

RTVH: Red Tilapia Viscera Hydrolysates.

FRTVH: Fraction <3 kDa of Red Tilapia Viscera Hydrolysates.

PI: Yamú Protein Viscera Isolate.

DH9: Yamú Protein Viscera Hydrolysates with 9% of Degree of hydrolysis.

DH28: Yamú Protein Viscera Hydrolysates with 28% of Degree of hydrolysis.

Advertisement

4. Enzymatic hydrolysis of fish by-product proteins

4.1 Protein hydrolysis

Protein hydrolysis occurs when a peptide bond is broken by water action, in the presence of a catalyst that may be an enzyme or a chemical agent [42]. Low-cost chemical processes can be by acid or alkaline hydrolysis, but they are non-specific, not reproducible and lead to amino acid denaturation. On the other hand, enzymatic hydrolysis is more expensive but does not deteriorate amino acids [43].

Once the native protein is broken, fragments of the native protein (oligomers) form, which can be a substrate for the subsequent hydrolysis process, so it is a multi-substrate reaction [44], especially in mediums where no pure protein is available, but mixtures of innumerable proteins, such as in fish residues and in general in other agro-industrial waste. Due to the hydrolysis process, the molecular characteristics of the proteins change, because the average molecular weight of the protein fragments present decreases, this increases the surface load, causes the release of hydrophobic groups, and changes functional properties, among other effects [45].

4.2 Enzymatic hydrolysis of protein

This process consists of decomposing proteins into smaller fragments, whose catalysts are enzymes called proteases [11]. This is a set of simultaneous link break reactions, consisting of serial stages, with different species loaded in equilibrium, giving fragments of decreasing size as follows [46]:

proteousproteinspeptonespeptidesamino acid

The catalytic process that occurs is divided into three steps. First, the enzyme (E) should approach the substrate (S) and bind to form the enzyme-substrate complex (ES). Second, the rupture of the peptide bond results in the release of a peptide. Third, the remaining peptide is separated from the enzyme after a nucleophilic attack from a water molecule [11]. Each of these reactions has its speed as described in Eq. (1) [47]. This process can be repeated on any of the peptides formed [46].

E+SK1ESK2EP+HP´+H2OK3E+POH+HP´E1

E: Enzyme, S: Substrate, P and P´: Resulting peptides, kx: Constant reaction rate.

This procedure has advantages over chemical hydrolysis as they have high selectivity and low contamination. It is a specific process that is carried out under moderate pH and temperature conditions, which makes it easy to control [30]. The product obtained is called protein hydrolyzate and it consists of peptides generally between 2 and 20 amino acids [48]. However, there are also disadvantages such as the high enzyme costs and long processing times [49].

Critical operating conditions in protein enzymatic hydrolysis include temperature, pH, enzyme type and concentration, substrate and concentration, cofactors, coenzymes, hydrolysis time [50], agitation speed [51], and presence of inhibitors, like fat in fish by-products [11].

On the other hand, variations that enzyme activity may suffer during the reaction should be controlled, such as denaturation, aggregation, or enzyme inactivation, which can be produced by temperature effects, pH shear stress or other substances that interfere with catalysis [12].

4.2.1 Enzymatic hydrolysis kinetics

During the reaction, the enzyme attacks the peptide bond as follows [52, 53]:

Opening of the peptide bond

CHR´CONHCHR´´+H2OenzymeCHR´COOH+NH2CHR´´E2

Proton exchange

CHR´COOHNH2CHR´´CHR´COO+N+H3CHR´´E3

Tritation of amino group

N+H3CHR´´+OHNH2CHR´´+H2OE4

Under neutral or alkaline conditions, the dissociation of the amino group becomes significant, so a decrease in pH may occur due to the accumulation of the protons released, which makes it necessary to add a base to keep pH constant and prevent the enzyme from being affected in its activity [30]. The analysis of the equations above concludes that the amount of hydrolyzed protein is proportional to the amount of base required to neutralize the pH of the reaction medium [30].

4.2.2 Follow-up of hydrolysis reaction

The hydrolysis reaction progress is established by the Hydrolysis Degree (HD), expressed as a fraction or percentage of the number of broken peptide bonds at any given time (h) for the total peptide bonds in the intact protein (htot) (Eq. 5) [54]. Both can be expressed as protein meq/g or as protein mmol/g [30].

GH%=hhtot.100E5

Methods used to determine Hydrolysis Degree (HD) include the pH-stat method [52], O-phthaldialdehyde (OPA) [54], Trinitrobenesulfonic acid (TNBS) [55], formalin titration, and soluble nitrogen in trichloroacetic acid (TCA) [56]. The fundamental difference between these methods is in the principle that each one is based to measure the number of broken bonds (h) at any given time of the reaction, because htot is usually determined from the analysis of the total amino acid content in the intact protein [57].

4.2.2.1 pH-stat method

This method is based on the fact that in peptide bond hydrolysis, a carboxyl group and an amino group are released. In an aqueous solution, these groups will be more or less ionized depending on pH [55]. At neutral or alkaline pH, carboxyl groups are fully ionized and proton exchange occurs between the carboxyl group and the amino group. At alkaline pH, amino groups will also be partially or fully ionized depending on the pH and amino acid in question, since the pK of the free amino acids N-terminal amino group ranges from 9 to 10.8. The following equations show, in general, the chemical species involved in protein enzymatic hydrolysis [58].

P1CONHP2+H2OproteaseP1COOH+NH2P2E6
P1COOHP1COO+H+E7
NH2P2NH3+P2E8

The resulting free protons cause a pH decrease of the reaction mixture, and a base addition is required to keep pH constant. The amount of base required is directly related to the amount of hydrolyzed peptide bonds, and it can be used to estimate HD. Unfortunately, the relationship between HD and base consumption is not simple and depends on several variables, including pK of the α-amino group released, the temperature of the reaction mixture, and length of the peptide chain [52]. The relationship between the spent base volume and HD has been described by Adler-Nissen, 1986 [55] in Eq. (9).

GH%=BNBMpαhTot.100E9

where B is the base volume consumed in L to keep pH constant, MP is the protein mass in kg, NB is the base concentration, and α is the dissociation degree of the amino groups released in the reaction. α and pK are calculated with Eqs. (10) and (11), respectively, where T is the temperature (K) [59].

α=10pHpK1+10pHpKE10
pK=7.8+298T298T2400E11

4.2.2.2 O-phthaldialdehyde method (OPA) and Trinitrobencenesulfonic acid method (TNBS)

Both methods are spectrophotometric, based on the determination of the α-amino groups released, by derivatization with trinitro-bencenesulfonic acid or ortho-phthaldialdehyde, respectively [56]. They are detected in the ultraviolet–visible range for the TNBS method, or fluorescent for the OPA. The absorbance value obtained is then converted into quantitative values using a standard curve prepared with a free amino acid, usually leucine, calculating HD as the percentage proportion of the amino acid released in the hydrolyzed regarding the amino acid amount of the whole protein [54, 55]. In Figures 1 and 2, reactions of an amino group with TNBS and OPA, respectively, take place [56].

Figure 1.

Reaction of OPA with amino acids. Source [60].

Figure 2.

Reaction of TNBS with amino acids. Source [61].

However, in these methods, derivatization reagents exhibit different reactivity to some amino acids, affecting measurement accuracy. For example, in the case of the OPA method, it will not be accurate when applied on proline- and cysteine-rich hydrolyzates [57].

4.2.3 Proteases most important characteristics

Proteases are the enzymes responsible for catalyzing the hydrolysis reaction of protein-peptide bonds, also known as peptidases [62]. Although, they can be obtained from plants, animals or microorganisms, most commercially viable proteases are obtained from this latter [63], especially Bacillus species, such as Bacillus licheniformis, Bacillus subtilis, and Aspergillus fungal species such as Aspergillus niger, A. flavus, Ammophilus fumigatus, and A. oryzae [64]. Some of the commercial proteases that have been used to obtain hydrolyzates from fish residues include trypsin, chymotrypsin, pepsin, Alcalase® 2.4 L, Flavourzyme® 500 L, E Properase, pronase, collagenases, bromelain and papain [50].

Proteases belong to the hydrolases group, they constitute a large and complex group of enzymes that differ from each other in their specificity due to substrate, their selectivity, the nature of their active sites, their catalytic mechanism, their stability profile, their pH, and optimum temperature. For these reasons, proteases cannot be classified under the general enzyme nomenclature system, but are classified according to their catalytic action, the nature of their active site, and their optimal pH value [63]. From the point of view of functional groups that have their active site, proteases can be classified into four main groups as follows [62]: Serine Proteases, Aspartic Proteases, Cysteine Proteases, Metalloproteases. On the other hand, when considering its catalytic mode of action, i.e., the excision site of the polypeptide chain, proteases are classified into exopeptidases and endopeptidases [65]. While, based on their optimal pH range, proteases can be classified into alkaline, neutral and acidic.

Advertisement

5. Production of bioactive and techno-functional peptides of fishery by-products

According to the HD achieved, the hydrolyzate obtained will potentially have biological activities or techno-functional properties. HD less than 10% result in improved techno-functional properties, such as emulsification, foaming capacity and greater solubility, whereas a higher HD tend to deliver hydrolyzates with greater potential as bioactive peptide sources [66].

5.1 Bioactive peptides

A bioactive peptide is a sequence of amino acids that is encrypted in the intact protein, in which it remains inactive, but once released, it can interact with certain receptors and regulate the physiological functions of the organism [67]. This may express some kind of effect on metabolic behavior, either human or animal [65]. These peptides can be released from the protein by gastrointestinal digestion, enzyme hydrolysis, or fermentation [68].

Among the most widely studied biological activities, are antihypertensive [69] Antioxidant [11] Antimicrobial [70], antithrombotic [71], anticancer [11] metal chelating agent, anticoagulants, among others [72].

One of the methods currently applied for obtaining bioactive peptides is enzymatic hydrolysis using commercial enzymes, which represents a reproducible, scalable, and industrial-application-capable method [73]. In this technology, biological activities of the peptides obtained may be affected by the operating conditions applied to isolate proteins, hydrolysis degree, protease type, peptide structure, the amino acids sequence, concentration, and the molecular weight of the peptides obtained [74].

The relationship between the peptide’s biological activity and their molecular weight has been widely documented [73], so the search for conditions that maximize HD has been one of the priorities in many studies [75] Peptide fractions with molecular weights between 1 and 4 kDa are of the greatest interest for nutritional and/or pharmaceutical uses in particular [75].

5.1.1 Antioxidant peptides

Free radicals and reactive oxygen species ROS [76], can cause DNA, protein, or lipid damage, resulting in human body damage from neurodegenerative, inflammatory, cardiovascular, diabetes, and cancer diseases [76]. This type of effect can be counteracted by substances with antioxidant capacity, which have different mechanisms of action depending on the free radical reduction form, among which are SET (single electron transfer), and HAT (hydrogen atom transfer) [77]. Based on these mechanisms, some methods to evaluate the antioxidant capacity of different substances have been designed. SET-based methods detect the antioxidant’s ability to transfer a chemical species such as metals, carbonyls and electrons, the most commonly used methods of this type are ABTS and FRAP. In the case of HAT methods, the antioxidant ability to inactivate a free radical is measured through the donation of a hydrogen atom, in which one of the most commonly used methods is ORAC [77].

On the other hand, some metals such as iron and copper, which are of importance at the physiological level, may also participate in the formation of reactive oxygen species [78], as in the case of hydroxyl radicals (OH), that are formed by the Fenton reaction and can cause damage to different types of tissues (Canabady-Rochelle et al., 2018). In this sense, metal chelation can counteract the formation of metal-catalyzed radicals in some way, which has somewhat been considered as a form of antioxidant activity [79].

Thus, peptide antioxidant activity is related to metal chelating activity and electron donation activity, which facilitates interaction with free radicals and cuts the reaction chain in which they participate [80]. In addition, the presence of hydrophobic sequences in peptides can interact with lipid molecules, eliminating the donation of protons to result in lipid radicals [81]. Thus, the imidazole group in histidine residues participates in hydrogen atom transfer, electron transfer, active oxygen extinction and capture of hydroxyl radicals [82].

The antioxidant capacity in these hydrolyzates has been attributed to the presence at the N-terminal end of peptide sequences of non-polar hydrophobic amino acids, such as phenylalanine, alanine and proline, and hydrophilic amino acids such as tyrosine, histidine and valine [6]. Thus, capturing the activity of hydrogen peroxide, the chelating activity of Fe2+, and reducing the power of Abalone (Haliotis discus hannai) hydrolyzates was related to hydrophobic amino acids in their peptides [83]. The capturing capacity of radicals has also been attributed to the presence of aromatic residues [84]. While tryptophan and tyrosine have been attributed antioxidant activity mediated by their phenolic and indolic groups, capable of donating hydrogen atoms [85]. The Table 4 lists several sequences of antioxidant peptides, from different kinds of fish by-products.

SourceSequenceReference
Amur sturgeon (Acipenser schrenckii) skinPAGT[86]
COD (Gadus macrocephalus) gelatinTCSP, TGGGNV[87]
Hoki (Johnius belengerii) SkinHGPLGPL[88]
Tilapia (Oreochromis niloticus) skinEGL, TGDET[89]
Mackerel (Magalaspis cordvla) visceraACFL[90]
Biuefin leatherjacket (Navodon septentrionalis) HeadWEGPLK, GPP, GVPLT[91]
Salmon Pectoral finFLNEFLHV[92]
Black Pomfret (Parastromateus niger) VisceraAMTGLEA[93]
Skate (Raja porosa) CartilageF1MGPY, GPACDY, 1VAGPQ[94]
Grass carp (Ctenopharyngodon Idella) skinPYSFK, GFGPQLVGGRP[95]
Squid (Ommastrephes bartrami) VisceraWVAPLK[96]
Salmon (Salmo sp.) FinFLNEFLHV[92]

Table 4.

Amino acid sequence of antioxidant peptides from fish by-products.

5.1.2 Antihypertensive peptides

Hypertension is one of the most important cardiovascular risk factors worldwide, since high blood pressure currently affects about 20% of adults around the world [97]. In these blood pressure-increasing processes, the angiotensin I converter enzyme (ACE) plays a crucial role. This enzyme, a dipeptidyl carboxypeptidase (EC. 3.4.15.1), promotes the conversion of angiotensin I to a powerful angiotensin II vasoconstrictor, and inactivates the bradequinine vasodilator, which is a depressant of the renin-angiotensin system action [97]. Angiotensin II is also involved in the release of the steroid Na-retaining, which also tends to increase blood pressure [97]. For these reasons, a first step in the search for potentially useful substances to control high blood pressure is the ability test to inhibit ACE. In this sense, the search for peptides that can reach therapeutic tests as drugs for blood pressure control should initially be evaluated as ACE inhibitors [97]. The Table 5 lists several sequences of antihypertensive peptides, from different kind of fish by-products.

SourceSequenceReference
COD (Gadus macrocephalus) gelatinTCSP, TGGGNV[87]
Pollack (Theragra chalcogramma) SkinGPL, GPM[98]
Salmon (Oncorhynchus keta) SkinGLPLNLP[99]
Sea Bream ScaleGY, VY, GF, VIY[100]
Sardinella (Sardinella aurita) Head/visceraFRGLMHY[101]
Snakehead fish (Channidae sp) MuscleLYPPP, YSMYPP[102]
Small-spotted catshark (Scyliorhinus canicula) Muscle, viscera, skin, and frameVAMPF[103]
Lizard fish (Saurida elongate) MuscleRVCLP[104]
Lizardfish (Synodus macrops) Scale GelatinAGPPGSDGQPGAK[105]

Table 5.

Amino acid sequence of ACE inhibitor peptides from fish by-products.

5.1.3 Anti-carcinogenic peptides

Cancer (malignant tumor), one of the most common diseases in the world [106], consists of abnormal and uncontrolled growth of cells, with proliferation and spread in surrounding tissues [11]. Thus, inhibition of deregulated cell proliferation is one of the strategies for treating this type of disease [107]. Among the broad list of substances that have been evaluated for this purpose are luteinizing hormone-releasing hormone and Atrial natriuretic peptide, for the treatment of prostate and colorectal cancer, respectively [106].

Various fish-derived proteins have been reported as sources of anticancer peptides [11, 108], as in the case with the antiproliferative activity of protein hydrolyzates of 18 fish species against breast cancer cell lines [109]. In Table 6, different fishery sources that have been active against some types of cancer are shown.

SourceType of cancerReference
Ruditapes philippinarum hydrolysatesprostate, breast, and lung cancer[110]
Squid gelatin hydrolysatesCMF-7, U87[111]
Oyster protein and anchovy hydrolysatesColon and prostate cancer[112]
Blood clam muscleProstate cancer[91]
Krab subproductsProstate cancer[113]

Table 6.

Use of peptides from fish by-products in cancer treatment.

There are three ways in which antiproliferative peptides act on cancer cells, apoptosis, necrosis, and cell cycle disturbances [11]. These mechanisms of action change according to structural characteristics such as molecular weight and amino acid composition. Thus, smaller peptides have greater molecular mobility and diffusivity, so they can interact better with the components of cancer cells. This activity has been attributed to amino acid sequences between 3 and 25 residues, with the predominance of hydrophobic amino acids, and one or more residues of Lys, Pro, Arg, Ser, Glu, THR Leu, Gly, Ala and Tyr. Because hydrophobic amino acids improve interactions between peptides and the outer surface of the bilayer of the tumor cell membrane, due to their phospholipid content and thus, they exert selective cytotoxic activity on these cells to healthy cells [107].

In addition to the amino acid sequence, the anti-cancer peptide’s function is influenced by net load, amphipathicity, hydrophobicity, structural membrane folding (including secondary structure, dynamics and orientation), oligomerization, and peptide concentration [11]. The cationic amphibious structure predisposes them to interact with the cell membrane anion surfaces [114]. The α helix is a main structural characteristic of this peptide type, with lateral chains of hydrophilic and hydrophobic amino acids, forming clear hydrophilic and hydrophobic surfaces. On the other hand, they concentrate on the N-terminal and the C-terminal to form different hydrophilic and hydrophobic domains. Anti-cancer peptides with a β sheet structure are generally stabilized by disulfide bonds, and these sheets are in β antiparallel formation. Meanwhile [11]. The net charge and positive charge number also influence these peptides activity, since their association with the cancer cell membrane occurs through electrostatic interactions due to its cationic condition and the anion lipopolysaccharide on the external membrane that causes its disturbance [115].

5.1.4 Anticoagulant peptides

Blood clotting is a crucial process for human health, excessive clotting that leads to blocked blood vessels causes strokes, heart attacks, and pulmonary embolism [11]. This makes anticoagulant compounds vital to preserving life quality in modern times. The anticoagulant is a compound that will stop blood clotting by binding to one or more coagulation factors, preventing it from binding to the membrane phospholipids [11]. Heparin is currently the anticoagulant most commonly used, but heparin has several disadvantages, including thrombocytopenia and non-specific plasma binding. In addition, it can cause platelet dysfunction and aggregation [116]. Therefore, there is a marked interest in the search for new anticoagulant compounds with minor collateral risks for the medical treatment of thromboembolic events [11].

Anticoagulant activity is less investigated than other biological activities, and specifically, peptides with this activity isolated from fish-based by-products have not been reported [11]. This way, an oligopeptide from the blue mussel, with a molecular mass of approximately 2,5 kDa has been isolated, showing anticoagulant activity by the prolongation of both thrombin time and activated partial thromboplastin time, by interaction specifically with blood clotting factors IX, X, and II. Nasri et al. [71], in 2012 isolated four anticoagulant peptides from protein hydrolyzates of goby muscle proteins, in which they found that they had Arg in the C-terminal position. Thus, concluding that small peptides with an amino acid charged at the C-end are considered potential thrombin inhibitors and/or other factors involved in the coagulation process [71]. Anticoagulant peptides from yellow-sole fish skeleton have also been isolated [117].

5.1.5 Antimicrobial peptides

The excessive use of conventional antimicrobial products has caused the emergence of resistant strains, which poses a health threat. Therefore, the development of antimicrobials using mechanisms other than traditional antibiotics is needed [11]. In this context, antimicrobial peptides effectively promote toxicity against invading pathogenic microorganisms, and also modulate the immune response in superior organisms [118]. These peptides are produced in all kingdoms, from bacteria to fungi and plants to mammals. Their unique intrinsic properties make them attractive therapeutic agents, since they show high biological activities associated with low toxicity and high specificity, as well as potentially useful as ingredients of functional or health-promoting foods [119]. These peptides generally contain less than 50 amino acid residues, with a molecular weight less than 10 kDa [120]. Despite their structural diversity, they have common physico-chemical characteristics; they are positively charged (+2 to +9) under physiological conditions due to the presence of lysine, arginine and histidine residues; and contain a substantial portion (50% or more) of hydrophobic residues [118]. These peptides commonly adopt an amphipathic conformation in which positively charged and hydrophobic groups are segregated into opposite faces of a α-helix, a β-leaf, or some other tertiary structure. This gives them the ability to cross the phospholipid membrane. The spectrum of different chemical properties of the amino acid side chain provides a variety of peptide sequences to show a cationic amphibious helical peptide [121]. Having a positive net charge allows them to interact with the anionic phospholipids of the bacterial membrane or other pathogens, and their amphipathicity, i.e., presence of apolar regions (with hydrophobic amino acids), and positive loads regions (cationic amino acids, Arg, Lys or His), facilitates them that, after initial interaction, the polar regions interact with the polar chains of the phospholipids, achieving the insertion of the peptide into the microbial membrane [122]. They are also flexible, which allows their internalization toward the bacterial cytoplasm, and leads to cell death due to ion and metabolic substances loss [123].

The most common mechanisms of action recognized in peptide antimicrobial activity include (i) the barrel model, in which a water-filled channel and an ion channel protein are formed by the interaction of peptides, acting as pores that disrupt the structure of the cell membrane; (ii) toroidal pore, in which less organized pore structures are formed; (iii) carpet models, in which the destabilization of the cell membrane in mycellar structures is caused by the accumulation of peptides above the limit concentration; (iv) molecular electroporation, following the concept that molecular electroporation can be achieved not only by electrical fields externally applied, but also by highly charged molecules that bind to the membrane surface; (v) sinking raft model, product of the induction of the membrane curvature by adsorbed peptides, which is relieved by its aggregation in the bilayer, allowing the aggregate to be translocated into the lumen of the gallbladder by a sinking raft process; and after membrane permeation, intracellular targets activation or blocking occurs [11]. These peptides not only generate toxic effects on microorganisms, but also exert important effects on the host, including immunomodulation, angiogenesis induction, wound healing and gene expression modulation. These effects may complement each other during the control of infectious and inflammatory diseases, and may be highly desirable when considered an optimal combination of an antimicrobial compound and regeneration booster [118]. In recent decades, barbel muscle antimicrobial peptides have been obtained by enzymatic hydrolysis of proteins from aquatic organisms [124]. Mustelus viscera [125], sea cucumber by-products [126], and different fish species [120], among others.

5.2 Commercial peptides obtained from fish sources

Thanks to their potential to produce bioactive compounds, fish parts and their residues have been used to obtain different types of functional inputs that have reached the market in different countries (Table 7). It should be noted, however, that few countries in which these products are being marketed. Given that fish, production extends to a much larger number of countries and that waste from that industry is proportional to production, it is clear that there is a latent possibility of expanding the market for products derived from fish sources.

Commercial nameSourceFunctionalityCountry
Custom Collagen®TilapiaLiver and kidneyUS
Hydroiyzed Fish Collagen Type 1TilapiaSkin, tendons, and arteriesUK
Amizate®Atlantic salmonMuscle anabolismNorth America
Protizen®Stress, weight disorder, sleep troubleUK
Levenorm®SardaAntihypertensiveCanada
MOLVAL®MolvaCholesterol, stress, and cardiovascular healthUK
Norland Hydrolyzed Fish CollagenCodHair, skin and nailsUS
PeptACE®SardaVascular function and blood pressureJapan and US
Stabilium®200Molva dypterygiaStress, memory, and cognitive functionUK
Seacure®HakeGastrointestinal and bowel functionCanada and US
Seagest™White fishIntestinal lining and healthUS
Valtyron®SardineBlood pressure
Vasotensin®Tuna and verdelVascular function and blood pressureJapan and US
Nutripeptin®CodWeight and blood glucoseUS and UK
Liquamen®MolvaOxidative stress, glycemic index, and stressUK

Table 7.

Commercial products obtained by enzymatic Hydroiysis of fish protein by-products [37, 127].

References

  1. 1. Ishak NH, Sarbon NM. A review of protein hydrolysates and bioactive peptides deriving from wastes generated by fish processing. Food and Bioprocess Technology. 2018;11:2-16
  2. 2. FAO. The State of World Fisheries and Aquaculture: Meeting the Sustainable Development Goals. Rome: FAO; 2018
  3. 3. Ghaly AE, Ramakrishnan VV, Brooks MS, Budge SM, Dave D. Fish processing wastes as a potential source of proteins, amino acids and oils: A critical review. J Microb Biochem Technol. 2013;5:107-129. DOI: 10.4172/1948-5948.1000110
  4. 4. AMEC. Management of wastes from Atlantic seafood processing operations. Dartmouth, Nova Scotia, Canada: AMEC Earth and Environment Limited; 2003
  5. 5. Gaviria YS, Figueroa OA, Zapata JE. Aplicación de la metodología de huella ecológica como indicador de sostenibilidad en el uso de ensilaje de pescado en dietas para alimentación de aves. Información tecnológica. 2021;32(5):199–208. DOI: 10.4067/s0718-07642021000500199
  6. 6. Villamil O, Vaquiro H, Solanilla JF. Fish viscera protein hydrolysates: Production, potential applications and functional and bioactive properties. Food Chemistry. 2017;224:160-171. DOI: 10.1016/j.foodchem.2016.12.057
  7. 7. Tchoukanova N, Gonzalez M, Poirier S. Best Management Practices: Marine Products Processing. New Brunswick: The Fisheries and Marine Products Division of the Coastal Zones Research Institute Inc; 2012
  8. 8. Suarez LM, Montes JR, Zapata JE. Optimization acid content in silage víscera red tilapia (Oreochromis spp.) with Life cycle assessment of feeds derived. Technol Inf. 2018;29(6):83-94
  9. 9. Liu Y, Li X, Chen Z, Yu J, Wang F, Wang J. Characterization of structural and functional properties of fish protein hydrolysates from surimi processing by-products. Food Chemistry. 2014;151:459-465
  10. 10. Zapata JE, Moya M, Figueroa OA. Hidrólisis Enzimática de la Proteína de Vísceras de Trucha Arco Íris (Oncorhynchus mykiss): Efecto del tipo de Enzima, Temperatura, pH y Velocidad de Agitación. Información Tecnológica. 2019;30(6):63-72
  11. 11. Gomez LJ, Gomez NA, Zapata JE, Lopez-García G, Cilla A, Alegría A. In-vitro antioxidant capacity and cytoprotective/cytotoxic effects upon Caco-2 cells of red tilapia (Oreochromis spp.) viscera hydrolysates. Food Research International. 2019;120:52-61. DOI: 10.1016/j.foodres.2019.02.029
  12. 12. Esteban MB, García AJ, Ramos P, Márquez MC. Evaluation of fruitvegetable and fish wastes as alternative feedstuffs in pig diets. Waste Management. 2007;27:193-200
  13. 13. Ghaedian R, Coupland JN, Decker EA, McClemets JD. Ultrasonic determination of fish composition. Journal of Food Engineering. 1998;35:323-337
  14. 14. Carranza-Saavedra D, Zapata-Montoya JE, Váquiro-Herrera HA, Solanilla-Duque JF. Study of biological activities and physicochemical properties of Yamú (Brycon siebenthalae) viscera hydrolysates in sodium alginate-based edible coating solutions. International Journal of Food Engineering. 2021;17:677-691. DOI: 10.1515/ijfe-2021-0036
  15. 15. Leonarduzzi E, Rodrigues KA, Macchi GJ. Proximate composition and energy density in relation to Argentine hake females (Merluccius hubbsi) morphometrics and condition indices. Fisheries Research. 2014;160:33-40
  16. 16. Gaviria YS, Figueroa OA, Zapata JE. Efecto de la inclusión de ensilado químico de vísceras de tilapia roja (Oreochromis spp.) en dietas para pollos de engorde sobre los parámetros productivos y sanguíneos. Información tecnológica. 2021;32:79-88. DOI: 10.4067/S0718-07642021000300079
  17. 17. Chuesiang P, Sanguandeekul R. Protein hydrolysate from tilapia frame: Antioxidant and angiotensin I converting enzyme inhibitor properties. International Journal of Food Science and Technology. 2015;50(6):1436-1444
  18. 18. Bhaskar N, Mahendrakar NS. Protein hydrolysate from visceral waste proteins of Catla (Catla catla): Optimization of hydrolysis conditions for a commercial neutral protease. Bioresour Technology. 2008;99(10):4105-4111. DOI: 10.1016/j.biortech.2007.09.006
  19. 19. Pires C, Teixeira B, Cardoso C, Mendes R, Nunes ML, Batista I. Cape hake protein hydrolysates prepared from alkaline solubilised proteins pre-treated with citric acid and calcium ions: Functional properties and ACE inhibitory activity. Process Biochemistry. 2015;50(6):1006-1015
  20. 20. Quintero J, Zapata JE. Optimización de la extracción del colágeno soluble en Ácido de subproductos de tilapia roja (oreochromis spp) mediante un diseño de superficie de respuesta. Información tecnológica. 2017;28:109-120. DOI: 10.4067/s0718-07642017000100011
  21. 21. Sousa SC, Vazquez JA, Pérez-Martín RI, Carvalho AP, Gomes AM. Valorization of byproducts from commercial fish species: Extraction and chemical properties of skin gelatins. Molecules. 2017;22(1545):1-12
  22. 22. Sampedro LJG, Grimaldos NAG, Pereañez JA, Montoya JEZ. Lipids as competitive inhibitors of subtilisin carlsberg in the enzymatic hydrolysis of proteins in red tilapia (oreochromis sp.) viscera: Insights from kinetic models and a molecular docking study. Brazilian Journal of Chemical Engineering. 2019;36:647-655. DOI: 10.1590/0104-6632.20190362s20180346
  23. 23. Taheri A, Anvar SAA, Ahari H, Fogliano V. Comparison the functional properties of protein Hydrolysates from poultry byproducts and rainbow trout (Onchorhynchus mykiss) viscera. Iranian Journal of Fisheries Sciences. 2013;12(1):154-169
  24. 24. Opheim M, Šližytė R, Sterten H, Provan F, Larssen E, Kjos NP. Hydrolysis of Atlantic salmon (Salmo salar) rest raw materials- effect of raw material and processing on composition, nutritional value, and potential bioactive peptides in the hydrolysates. Process Biochemistry. 2015;50(8):1247-1257
  25. 25. Roslan J, Yunos KFM, Abdullah N, Kamal SMM. Characterization of fish protein hydrolysate from tilapia (Oreochromis Niloticus) by-product. Agriculture and Agricultural Science Procedia. 2014;2:312-319. DOI: 10.1016/j.aaspro.2014.11.044
  26. 26. Giraldo-Rios DE, Rios LA, Zapata-Montoya JE. Kinetic modeling of the alkaline deproteinization of Nile-tilapia skin for the production of collagen. Heliyon. 2020;6:1-8. DOI: 10.1016/j.heliyon.2020.e03854
  27. 27. Gencbay G, Turhan S. Proximate composition and nutritional profile of the black sea anchovy (Engraulis encrasicholus) whole fish, fillets, and by products. J Aquat Food Prod T. 2016;25(6):864-874
  28. 28. Arias L, Gómez LJ, Zapata JE. Efecto de Temperatura-Tiempo sobre los lípidos extraídos de vísceras de tilapia roja (oreochromis sp.) utilizando un proceso de Calentamiento-Congelación. Información tecnológica. 2017;28:131-142. DOI: 10.4067/s0718-07642017000500014
  29. 29. Soufi-Kechaou E, Derouiniot-Chaplin M, Ben Amar R, Jaouen P, Berge JP. Recovery of valuable marine compounds from cuttlefish by-product hydrolysates: Combination of enzyme bioreactor and membrane technologies. C. R. Chimie. 2017;20(9–10):975-985
  30. 30. Sierra-Lopera LM, Zapata-Montoya JE. Optimization of enzymatic hydrolysis of red tilapia scales (Oreochromis sp.) to obtain bioactive peptides. Biotechnology Reports. 2021;30:e00611. DOI: 10.1016/j.btre.2021.e00611
  31. 31. Morkore T, Vallet JL, Cardinal M, Guillen MCG, Montero P, et al. Fat content and fillet shape of Atlantic salmon: Relevance for processing yield and quality of raw and smoked products. Journal of Food Science. 2001;66:1348-1354
  32. 32. FAO/NACA/WHO J. The State of World Fisheries and Aquaculture [internet]. 2016. USA. Available from: www.fao.org/publications [Accessed: 29 Jan 2021]
  33. 33. Friedman K. Nutritional value of proteins from different food sources: A review. Journal of Agricultural and Food Chemistry. 1996;44:6-29
  34. 34. Dunajski E. Texture of fish muscle. Journal of Texture Studies. 1979;10:301-318
  35. 35. Chen J, Li L, Yi R, Xu N, Gao R, Hong B. Extraction and characterization of acid-soluble collagen from scales and skin of tilapia (Oreochromis niloticus). LWT - Food Science and Technology. 2016;66:453-459
  36. 36. Zhang Q, Wang Q, Lv S, Lu J, Jiang S, Regenstein JM, et al. Comparison of collagen and gelatin extracted from the skins of Nile tilapia (Oreochromis niloticus) and channel catfish (Ictalurus punctatus). Food Bioscience. 2016;13:41-48
  37. 37. Chalamaiah M, Dinesh Kumar B, Hemalatha R, Jyothirmayi T. Fish protein hydrolysates: Proximate composition, amino acid composition, antioxidant activities and applications: A review. Food Chemistry. 2012;135:3020-3038
  38. 38. Spinelli J, Dassow JA. Fish proteins: Their modification and potential uses in the food industry. In: Chemistry and Biochemistry of Marine Food Products. Westport, CT, USA: AVI Publishing Company; 1982. pp. 13-25
  39. 39. Kristinsson HG, Rasco BA. Fish protein hydrolysates: Production, biochemical, and functional properties. Critical Reviews in Food Science and Nutrition. 2000;40:43-81
  40. 40. Sepúlveda CT, Zapata JE, Martínez-Álvarez O, Alemán A, Montero MP, Gómez-Guillén MC. The preferential use of a soy-rapeseed lecithin blend for the liposomal encapsulation of a tilapia viscera hydrolysate. LWT. 2021;139:110530. DOI: 10.1016/j.lwt.2020.110530
  41. 41. Brooks MS, Ghaly AE, Ramakrishnan VV, Budge SM, Dave D. Fish processing wastes as a potential source of proteins, amino acids and oils: A critical review. Journal of Microbial & Biochemical Technology. 2013;5(4):107-129. DOI: 10.4172/1948-5948.1000110
  42. 42. Sierra M. Revaloración de escamas y esqueletos de tilapia roja (oreochromis spp.) para la obtención de péptidos bioactivos. Medellin: Universidad de Antioquia; 2021
  43. 43. Tavano OL. Protein hydrolysis using proteases: An important tool for food biotechnology. Journal of Molecular Catalysis B: Enzymatic. 2013;90:1-11. DOI: 10.1016/j.molcatb.2013.01.011
  44. 44. Trusek-Holownia A, Lech M, Noworyta A. Protein enzymatic hydrolysis integrated with ultrafiltration: Thermolysin application in obtaining peptides. Chemical Engineering Journal. 2016;305:61-68. DOI: 10.1016/j.cej.2016.05.087
  45. 45. Samaranayaka AGP, Li-Chan ECY. Food-derived peptidic antioxidants: A review of their production, assessment, and potential applications. Journal of Functional Foods. 2011;3(4):229-254. DOI: 10.1016/j.jff.2011.05.006
  46. 46. Benítez R, Ibarz A, Pagan J. Hidrolizados de proteína: procesos y aplicaciones Resumen. Acta Bioquímica Clínica Latinoamericana. 2008;42(2):227-237
  47. 47. Mc Auley MT. Computer Modelling for Nutritionists. Chester: Springer; 2019. DOI: 10.1007/978-3-319-39994-2
  48. 48. Zamora-Sillero J, Gharsallaoui A, Prentice C. Peptides from fish by-product protein hydrolysates and its functional properties: An overview. Marine Biotechnology. 2018;20(2):118-130. DOI: 10.1007/s10126-018-9799-3
  49. 49. He S, Franco C, Zhang W. Functions, applications and production of protein hydrolysates from fish processing co-products (FPCP). Food Research International. 2013;50(1):289-297. DOI: 10.1016/j.foodres.2012.10.031
  50. 50. Wang C, Chang T, Shi L, Yang H, Cui M, Tambalu L. Seafood processing by-products: Collagen and gelatin. In: Kim S-K, editor. Seafood Processing By-Products. 1st ed. New York: Springer; 2014. pp. 207-242. DOI: 10.1007/978-1-4614-9590-1_12
  51. 51. Gómez LJ, Zapata JE. Efecto del Nivel de Grasa y Velocidad de Agitación en la Hidrolisis Enzimática de Vísceras de Tilapia Roja (Orechromis sp.). Información tecnológica. 2017;28(4):47-56. DOI: 10.4067/s0718-07642017000400007
  52. 52. Adler-Nissen J. Enzymic hydrolysis of food proteins. London: Elsevier Applied Science Publishers; 1986
  53. 53. Márquez MC, Vázquez MA. Modeling of enzymatic protein hydrolysis. Process Biochemistry. 1999;35:111-117
  54. 54. Nielsen PM, Petersen D, Dambmann C. Improved method for determining food protein degree of hydrolysis. Journal of Food Science. 2001;66(5):642-646. DOI: 10.1111/j.1365-2621.2001.tb04614.x
  55. 55. Adler-Nissen J. Determination of the degree of hydrolysis of food protein hydrolysates by trinitrobenzenesulfonic acid. Journal of Agricultural and Food Chemistry. 1979;27(6):1256-1262
  56. 56. Rutherfurd SM. Methodology for determining degree of hydrolysis of proteins in hydrolysates: A review. Journal of Aoac International. 2010;93(5):1515-1522. DOI: 10.1093/jaoac/93.5.1515
  57. 57. Wubshet SG, Lindberg D, Veiseth-Kent E, Kristoffersen KA, Böcker U, Washburn KE, et al. Bioanalytical aspects in enzymatic protein hydrolysis of by products. In: Proteins: Sustainable Source, Processing and Applications. Cambridge: Academic Press; 2019. DOI: 10.1016/b978-0-12-816695-6.00008-8
  58. 58. Mat DJL, Cattenoz T, Souchon I, Michon C, Le Feunteun S. Monitoring protein hydrolysis by pepsin using pH-stat: In vitro gastric digestions in static and dynamic pH conditions. Food Chemistry. 2018;239:268-275. DOI: 10.1016/j.foodchem.2017.06.115
  59. 59. Valencia P, Espinoza K, Ceballos A, Pinto M, Almonacid S. Novel modeling methodology for the characterization of enzymatic hydrolysis of proteins. Process Biochemistry. 2015;50(4):589-597. DOI: 10.1016/j.procbio.2014.12.028
  60. 60. Dzwolak W, Ziajka S. Enzymatic hydrolysis of milk proteins under alkaline and acidic conditions. Journal of Food Science. 1999;64:393-395
  61. 61. Satake K, Okuyama T, Ohashi M, Shinoda T. Journal of Biochemistry. 1960;47:654-660
  62. 62. Rao MB, Tanksale AM, Ghatge MS, Deshpande VV. Molecular and biotechnological aspects of microbial proteases. Microbiology and Molecular Biology Reviews. 1998;62(3):597-635
  63. 63. Vazquez JA, Medina A, Duran AI, Nogueira M, Fernandez-Compas A, Perez Martin RI, et al. Production of valuable compounds and bioactive metabolites from by-products of fish discards using chemical processing, enzymatic hydrolysis, and bacterial fermentation. Marine Drugs. 2019;17(3):139. DOI: 10.3390/md17030139
  64. 64. Singh P, Kumar S. Microbial enzyme in food biotechnology. In: Kuddus M, editor. Enzymes in Food Biotechnology. Cambridge: Academic Press; 2019. pp. 19-28. DOI: 10.1016/B978-0-12-813280-7.00002-5
  65. 65. Martínez-Medina GA, Barragán AP, Ruiz HA, Ilyina A, Martínez Hernández JL, Rodríguez-Jasso RM, et al. Fungal proteases and production of bioactive peptides for the food industry. In: Kuddus M, editor. Enzymes in Food Biotechnology. Cambridge: Academic Press; 2019. pp. 221-246. DOI: 10.1016/B978-0-12-813280-7.00014-1
  66. 66. Beaubier S, Framboisier X, Ioannou I, Galet O, Kapel R. Simultaneous quantification of the degree of hydrolysis, protein conversion rate and mean molar weight of peptides released in the course of enzymatic proteolysis. Journal of Chromatography B: Analytical Technologies in the Biomedical and Life Sciences. 2019;1105:1-9. DOI: 10.1016/j.jchromb.2018.12.005
  67. 67. Tkaczewska J, Jamroz E, Kulawik P, Morawska M, Szczurowska K. Evaluation of the potential use of a carp (Cyprinus carpio) skin gelatine hydrolysate as an antioxidant component. Food & Function. 2019;10(2):1038-1048. DOI: 10.1039/c8fo02492h
  68. 68. Sabbione AC, Nardo AE, Añón MC, Scilingo A. Amaranth peptides with antithrombotic activity released by simulated gastrointestinal digestion. Journal of Functional Foods. 2016;20:204-214. DOI: 10.1016/j.jff.2015.10.015
  69. 69. Ketnawa S, Suwal S, Huang J, Liceaga AM. Selective separation and characterisation of dual ACE and DPP-IV inhibitory peptides from rainbow trout (Oncorhynchus mykiss) protein hydrolysates. International Journal of Food Science & Technology. 2018;54(4):1062-1073. DOI: 10.1111/ijfs.13939
  70. 70. Aissaoui N, Chobert JM, Haertle T, Marzouki MN, Abidi F. Purification and biochemical characterization of a neutral serine protease from Trichoderma harzianum. Use in antibacterial peptide production from a fish by-product hydrolysate. Applied Biochemistry and Biotechnology. 2017;182(2):831-845. DOI: 10.1007/s12010-016-2365-4
  71. 71. Nasri R, Amor IB, Bougatef A, Nedjar-Arroume N, Dhulster P, Gargouri J, et al. Anticoagulant activities of goby muscle protein hydrolysates. Food Chemistry. 2012;133(3):835-841. DOI: 10.1016/j.foodchem.2012.01.101
  72. 72. Agyei D, Ongkudon CM, Wei CY, Chan AS, Danquah MK. Bioprocess challenges to the isolation and purification of bioactive peptides. Food and Bioproducts Processing. 2016;98:244-256. DOI: 10.1016/j.fbp.2016.02.003
  73. 73. Gómez LJ, Figueroa OA, Zapata YJE. Actividad antioxidante de hidrolizados enzimáticos de plasma bovino obtenidos por efecto de alcalasa® 2.4 L. Información Tecnológica. 2013;24(1):33-42. DOI: 10.4067/S0718-07642013000100005
  74. 74. Sarmadi BH, Ismail A. Antioxidative peptides from food proteins: a review. Peptides. 2010;31(10):1949-1956. DOI: 10.1016/j.peptides.2010.06.020
  75. 75. Saidi S, Deratani A, Belleville MP, Amar RB. Production and fractionation of tuna by-product protein hydrolysate by ultrafiltration and nanofiltration: Impact on interesting peptides fractions and nutritional properties. Food Research International. 2014;65:453-461. DOI: 10.1016/j.foodres.2014.04.026
  76. 76. Ngo D, Vo T, Ngo D, Wijesekara I, Kim S. Biological activities and potential health benefits of bioactive peptides derived from marine organisms. International Journal of Biological Macromolecules. 2012;51:378-383
  77. 77. Santos-Sánchez NF, Salas-Coronado R, Villanueva-Cañongo C, Hernández Carlos B. Antioxidant compounds and their antioxidant mechanism. In: Antioxidants. London: IntechOpen; 2019
  78. 78. Canabady-Rochelle LLS, Selmeczi K, Collin S, Pasc A, Muhr L, Boschi-Muller S. SPR screening of metal chelating peptides in a hydrolysate for their antioxidant properties. Food Chemistry. 2018;239:478-485. DOI: 10.1016/j.foodchem.2017.06.116
  79. 79. Decker EA, Welch B. Role of ferritin as a lipid oxidation catalyst in muscle food. Journal of Agricultural and Food Chemistry. 1990;38(3):674-677. DOI: 10.1021/jf00093a019
  80. 80. You L, Zhao M, Regenstein JM, Ren J. In vitro antioxidant activity and in vivo anti-fatigue effect of loach (Misgurnus anguillicaudatus) peptides prepared by papain digestion. Food Chemistry. 2011;124(1):188-194. DOI: 10.1016/j.foodchem.2010.06.007
  81. 81. Je JY, Qian ZJ, Byun HG, Kim SK. Purification and characterization of an antioxidant peptide obtained from tuna backbone protein by enzymatic hydrolysis. Process Biochemistry. 2007;42(5):840-846. DOI: 10.1016/j.procbio.2007.02.006
  82. 82. Aluko RE. Functional Foods and Nutraceuticals. Biotechnology: New Ideas, New Developments (A Textbook of Modern Technology). New York, NY: Springer New York; 2012. DOI: 10.1007/978-1-4614-3480-1
  83. 83. Je JY, Park SY, Hwang JY, Ahn CB. Amino acid composition and in vitro antioxidant and cytoprotective activity of abalone viscera hydrolysate. Journal of Functional Foods. 2015;16:94-103. DOI: 10.1016/j.jff.2015.04.023
  84. 84. Ghanbari R. Review on the bioactive peptides from marine sources: Indication for health effects. International Journal of Peptide Research and Therapeutics. 2019;25(3):1187-1199. DOI: 10.1007/s10989-018-9766-x
  85. 85. Pihlanto A. Antioxidative peptides derived from milk proteins. International Dairy Journal. 2006;16(11):1306-1314. DOI: 10.1016/j.idairyj.2006.06.005
  86. 86. Ngo DH, Ryu B, Vo TS, Himaya SWA, Wijesekara I, Kim SK. Free radical scavenging and angiotensin-I converting enzyme inhibitory peptides from Pacific cod (Gadus macrocephalus) skin gelatin. International Journal of Biological Macromolecules. 2011;49:1110-1116
  87. 87. Nikoo M, Benjakul S, Ehsani A, Li J, Wu F, Yang N, et al. Antioxidant and cryoprotective effects of a tetrapeptide isolated from Amur sturgeon skin gelatin. Journal of Functional Foods. 2014;7:609-620
  88. 88. Mendis E, Rajapakse N, Kim SK. Antioxidant properties of a radicals scavenging peptide purified from enzymatically prepared fish skin gelatin hydrolysate. Journal of Agriculture and Food Chemistry. 2005;53:581-587
  89. 89. Zhang YF, Duan X, Zhuang YL. Purification and characterization of novel antioxidant peptides from enzymatic hydrolysates of tilapia (Oreochromis niloticus) skin gelatin. Peptides. 2012;38:13-21
  90. 90. Kumar S, Sutherland F, Rosso R, Teh AW, Lee G, Heck PM, et al. Effects of chronic omega-3 polyunsaturated fatty acid supplementation on human atrial electrophysiology. Heart Rhythm. 2011;8:562-568
  91. 91. Chi CF, Wang B, Wang YM, Zhang B, Deng SG. Isolation and characterization of three antioxidant peptides from protein hydrolysate of bluefin leatherjacket (Navodon septentrionalis) heads. Journal of Functional Foods. 2015;12:1-10
  92. 92. Ahn CB, Kim JG, Je JY. Purification and antioxidant properties of octapeptide from salmon byproduct protein hydrolysate by gastrointestinal digestion. Food Chemistry. 2014;147:78-83
  93. 93. Jaiganesh R, Nazeer RA, Sampath Kumar NS. Purification and identification of antioxidant peptide from black pomfret, Parastromateus niger (Bloch, 1975) viscera protein hydrolysate. Food Science and Biotechnology. 2011;20:1087-1094
  94. 94. Pan X, Zhao Y, Hu F, Wang B. Preparation and identification of antioxidant peptides from protein hydrolysate of skate (Raja porosa) cartilage. Journal of Functional Foods. 2016;25:220-230
  95. 95. Cai L, Wu X, Zhang Y, Li X, Ma S, Li J. Purification and characterization of three antioxidant peptides from protein hydrolysate of grass carp (Ctenopharyngodon idella) skin. Journal of Functional Foods. 2015;16:234-242. DOI: 10.1016/j.jff.2015.04.042
  96. 96. Song R, Zhang K, Wei R. In vitro antioxidative activities of squid (Ommastrephes bartrami) viscera autolysates and identification of active peptides. Process Biochemistry. 2016;51(10):1674-1682
  97. 97. Wijesekara I, Kim SK. Angiotensin-I-converting enzyme (ACE) inhibitors from marine resources: Prospects in the pharmaceutical industry. Marine Drugs. 2010;8(4):1080-1093. DOI: 10.3390/md8041080
  98. 98. Byun HG, Kim SK. Purification and characterization of angiotensin I converting enzyme (ACE) inhibitory peptides from Alaska pollack (Theragra chalcogramma) skin. Process Biochemistry. 2001;36:1155-1162
  99. 99. Lee JK, Jeon JK, Byun HG. Antihypertensive effect of novel angiotensin I converting enzyme inhibitory peptide from chum salmon (Oncorhynchus keta) skin in spontaneously hypertensive rats. Journal of Functional Foods. 2014;7:381-389
  100. 100. Fahmi A, Morimura S, Guob HC, Shigematsu T, Kida K, Uemura Y. Production of angiotensin I converting enzyme inhibitory peptides from sea bream scale. Process Biochemistry. 2004;39:1195-1200
  101. 101. Bougatef A, Nedjar-Arroume N, Ravallec-Plé R, Leroy Y, Guillochon D, Barkia A, et al. Angiotensin I-converting enzyme (ACE) inhibitory activities of sardinelle (Sardinella aurita) by-products protein hydrolysates obtained by treatment with microbial and visceral fish serine proteases. Food Chemistry. 2008;111:350-356
  102. 102. Ghassem M, Babji AS, Said M, Mahmoodani F, Arihara K. Angiotensin I-converting enzyme inhibitory peptides from snakehead fish sarcoplasmic protein hydrolysate. Journal of Food Biochemistry. 2014;38(2):14-19
  103. 103. García-Moreno PJ, Espejo-Carpio FJ, Guadix A, Guadix EM. Production and identification of angiotensin I-converting enzyme (ACE) inhibitory peptides from Mediterranean fish discards. Journal of Functional Foods. 2015;18:95-105
  104. 104. Wu X, Cai L, Zhang Y, Mi H, Cheng X, Li J. Compositions and antioxidant properties of protein hydrolysates from the skins of four carp species. International Journal of Food Science & Technology. 2015;50(12):2589-2597. DOI: 10.1111/ijfs.12927
  105. 105. Chen J, Liu Y, Wang G, Sun S, Liu R, Hong B, et al. Processing optimization and characterization of angiotensin-Ι-converting enzyme inhibitory peptides from lizardfish (Synodus macrops) scale gelatin. Marine Drugs. 2018;16(7):228. DOI: 10.3390/md16070228
  106. 106. Baig MH, Ahmad K, Saeed M, Alharbi AM, Barreto GE, Ashraf GM, et al. Peptide based therapeutics and their use for the treatment of neurodegenerative and other diseases. Biomedicine & Pharmacotherapy. 2018;103:574-581. DOI: 10.1016/j.biopha.2018.04.025
  107. 107. Chalamaiah M, Yu W, Wu J. Immunomodulatory and anticancer protein hydrolysates (peptides) from food proteins: A review. Food Chemistry. 2018;245:205-222. DOI: 10.1016/j.foodchem.2017.10.087
  108. 108. Correia-da-Silva M, Sousa E, Pinto MMM, Kijjoa A. Anticancer and cancer preventive compounds from edible marine organisms. Seminars in Cancer Biology. 2017;46:55-64. DOI: 10.1016/j.semcancer.2017.03.011
  109. 109. Picot L, Bordenave S, Didelot S, Fruitier-Arnaudin I, Sannier F, Thorkelsson G, et al. Antiproliferative activity of fish protein hydrolysates on human breast cancer cell lines. Process Biochemistry. 2006;41:1217-1222. DOI: 10.1016/j.procbio.2005.11.024
  110. 110. Kim EK, Kim YS, Hwang JW, Lee JS, Moon SH, Jeon BT, et al. Purification and characterization of a novel anticancer peptide derived from Ruditapes philippinarum. Process Biochemistry. 2013;48(7):1086-1090. DOI: 10.1016/j.procbio.2013.05.004
  111. 111. Alemán A, Pérez-santín E, Bordenave-juchereau S, Arnaudin I, Gómez-guillén MC, Montero P. Squid gelatin hydrolysates with antihypertensive, anticancer and antioxidant activity. Food Research International. 2011;44:1044-1051. DOI: 10.1016/j.foodres.2011.03.010
  112. 112. Song R, Wei R, Luo H, Yang Z. Isolation and identification of an antiproliferative peptide derived from heated products of peptic hydrolysates of halffin anchovy (Setipinna taty). Journal of Functional Foods. 2014;10:104-111
  113. 113. Doyen A, Beaulieu L, Saucier L, Pouliot Y, Bazinet L. Demonstration of in vitro anticancer properties of peptide fractions from a snow crab by-products hydrolysate after separation by electrodialysis with ultrafiltration membranes. Separation and Purification Technology. 2011;78(3):321-329. DOI: 10.1016/j.seppur.2011.01.037
  114. 114. Gianfranceschi GL, Gianfranceschi G, Quassinti L, Bramucci M. Biochemical requirements of bioactive peptides for nutraceutical efficacy. Journal of Functional Foods. 2018;47:252-263. DOI: 10.1016/j.jff.2018.05.034
  115. 115. Schweizer F. Cationic amphiphilic peptides with cancer-selective toxicity. European Journal of Pharmacology. 2009;625(1):190-194. DOI: 10.1016/j.ejphar.2009.08.043
  116. 116. Jin NZ, Gopinath SCB. Potential blood clotting factors and anticoagulants. Biomedicine and Pharmacotherapy. 2016;84:356-365. DOI: 10.1016/j.biopha.2016.09.057
  117. 117. Rajapakse N, Jung WK, Mendis E, Moon SH, Kim SK. A novel anticoagulant purified from fish protein hydrolysate inhibits factor XIIa and platelet aggregation. Life Sciences. 2005;76(22):2607-2619. DOI: 10.1016/j.lfs.2004.12.010
  118. 118. Alencar-Silva T, Braga MC, Santana GOS, Saldanha-Araujo F, Pogue R, Dias SC, et al. Breaking the frontiers of cosmetology with antimicrobial peptides. Biotechnology Advances. 2018;36(8):2019-2031. DOI: 10.1016/j.biotechadv.2018.08.005
  119. 119. Sila A, Hedhili K, Przybylski R, Ellouz-Chaabouni S, Dhulster P, Bougatef A, et al. Antibacterial activity of new peptides from barbel protein hydrolysates and mode of action via a membrane damage mechanism against Listeria monocytogenes. Journal of Functional Foods. 2014;11:322-329. DOI: 10.1016/j.jff.2014.10.006
  120. 120. Najafian L, Babji AS. A review of fish-derived antioxidant and antimicrobial peptides: Their production, assessment, and applications. Peptides. 2012;33(1):178-185. DOI: 10.1016/j.peptides.2011.11.013
  121. 121. Deslouches B, Phadke SM, Lazarevic V, Cascio M, Islam K, Montelaro RC, et al. De novo generation of cationic antimicrobial peptides: Influence of length and tryptophan substitution on antimicrobial activity. Antimicrobial Agents and Chemotherapy. 2005;49(1):316-322. DOI: 10.1128/AAC.49.1.316-322.2005
  122. 122. Vogel HJ, Schibli DJ, Jing W, Lohmeier-Vogel EM, Epand RF, Epand RM. Towards a structure-function analysis of bovine lactoferricin and related tryptophanand arginine-containing peptides. Biochemistry and Cell Biology. 2002;80(1):49-63
  123. 123. Villarruel R, Huizar R, Corrales M, Sánchez T, Islas A. Péptidos naturales antimicrobianos: Escudo esencial de la respuesta inmune. Investigación en Salud. 2004;6(3):170-179
  124. 124. Sila A, Nedjar-Arroume N, Hedhili K, Chataigné G, Balti R, Nasri M, et al. Antibacterial peptides from barbel muscle protein hydrolysates: Activity against some pathogenic bacteria. LWT-Food Science and Technology. 2014;55(1):183-188. DOI: 10.1016/j.lwt.2013.07.021
  125. 125. Abdelhedi O, Jridi M, Jemil I, Mora L, Toldrá F, Aristoy MC, et al. Combined biocatalytic conversion of smooth hound viscera: Protein hydrolysates elaboration and assessment of their antioxidant, anti-ACE and antibacterial activities. Food Research International. 2016;86:9-23. DOI: 10.1016/j.foodres.2016.05.013
  126. 126. Tripoteau L, Bedoux G, Gagnon J, Bourgougnon N. In vitro antiviral activities of enzymatic hydrolysates extracted from byproducts of the Atlantic holothurian Cucumaria frondosa. Process Biochemistry. 2015;50:867-875. DOI: 10.1016/j.procbio.2015.02.012
  127. 127. Sierra Lopera LM, Sepúlveda Rincón CT, Vásquez Mazo P, Figueroa Moreno OA, Zapata Montoya JE. Byproducts of aquaculture processes: Development and prospective uses. Review. Revista Vitae. 2018;25(3):128-140. DOI: 10.17533/udea.vitae.v25n3a03

Written By

Jose Edgar Zapata Montoya and Angie Franco Sanchez

Submitted: 19 November 2021 Reviewed: 22 December 2021 Published: 14 February 2022