Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 4676 2024-02-19 09:09:47 |
2 Reference format revised. Meta information modification 4676 2024-02-20 01:36:19 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Nikoo, M.; Regenstein, J.M.; Yasemi, M. Protein Hydrolysates from Fishery Processing By-Products. Encyclopedia. Available online: https://encyclopedia.pub/entry/55147 (accessed on 04 July 2024).
Nikoo M, Regenstein JM, Yasemi M. Protein Hydrolysates from Fishery Processing By-Products. Encyclopedia. Available at: https://encyclopedia.pub/entry/55147. Accessed July 04, 2024.
Nikoo, Mehdi, Joe M. Regenstein, Mehran Yasemi. "Protein Hydrolysates from Fishery Processing By-Products" Encyclopedia, https://encyclopedia.pub/entry/55147 (accessed July 04, 2024).
Nikoo, M., Regenstein, J.M., & Yasemi, M. (2024, February 19). Protein Hydrolysates from Fishery Processing By-Products. In Encyclopedia. https://encyclopedia.pub/entry/55147
Nikoo, Mehdi, et al. "Protein Hydrolysates from Fishery Processing By-Products." Encyclopedia. Web. 19 February, 2024.
Protein Hydrolysates from Fishery Processing By-Products
Edit

Fish processing by-products such as frames, trimmings, and viscera of commercial fish species are rich in proteins. Thus, they could potentially be an economical source of proteins that may be used to obtain bioactive peptides and functional protein hydrolysates for the food and nutraceutical industries. The structure, composition, and biological activities of peptides and hydrolysates depend on the freshness and the actual composition of the material. Peptides isolated from fishery by-products showed antioxidant activity. Changes in hydrolysis parameters changed the sequence and properties of the peptides and determined their physiological functions. The optimization of the value of such peptides and the production costs must be considered for each particular source of marine by-products and for their specific food applications. 

marine protein hydrolysates hydrolysis variables structure-function relations antioxidant activity

1. Introduction

In response to the recognition of the limits of marine sources and the crisis of food security, the production of aquatic food products is increasing and reached 179 million tonnes in 2018, of which 22 million tonnes (or 12%) were not used for human consumption [1]. To achieve the maximum use of by-products from fish processing industries (e.g., heads, frames, skin, trimmings, and viscera from fish; cephalothorax and shells from shrimp; heads and tentacles from squids; and shells and byssus threads from oysters and mussels), which account for 40–60% of the total weight, it is necessary to retain them in the food chain in line with a sustainable circular economy through the production of high-value biomolecules [2][3][4][5]. Of the different types of by-products that are produced after filleting, canning, packaging, etc., the heads, frames, and trimmings constitute >75% of the by-products’ weight and contain significant amounts of muscle residue that can be used for direct human consumption or can be converted to functional food ingredients. Despite this necessity, within the seafood sector, the increased use by-products is happening relatively slowly because the seafood industry focuses on its primary raw materials and products that require minimal processing [6].
One focus has been peptides from marine sources (fish, shellfish, and invertebrates) with antioxidant activity [7][8][9][10]. Antioxidant peptides have an essential role in inhibiting oxidation and scavenging free radicals. In the body, they may help fight aging and reduce food oxidation. These peptides have been prepared using enzymatic or autolytic hydrolysis or microbial fermentation [2]. Most peptides studied contain 2–10 amino acids (AAs), although some contain up to 20 AAs and had a molecular weight (MW) of 0.2 to 2 kDa [11][12][13][14]. Several proteases with different specificities for peptide bond cleavage and different hydrolysis conditions (temperature, time, pH, enzyme-to-substrate ratio, water/by-product ratio, and stirring rate) have been used to produce these marine by-product peptides [15][16][17][18]. Changes in hydrolysis conditions changed their antioxidant activities [19][20][21]. The antioxidant activity of peptides in foods has been associated with the scavenging of the free radicals formed during lipid peroxidation and metal chelation [22], which are dependent on the AA composition and sequence, size, hydrophobicity, and N- or C-terminal residues [23][24][25].
Marine protein hydrolysates derived using enzymatic hydrolysis showed antioxidant activity against free radicals and pro-oxidative metal ions. Thus, they may potentially be used as alternative antioxidants in foods and the human body to fight against free radical-mediated aging [26][27]. The structure of antioxidant peptides and protein hydrolysates from marine sources are highly variable depending on the types of by-products used as the initial protein source and the various operating parameters that affect the hydrolysis and the functional outcome. These by-products could be stabilized with non-oxidized products to preserve their freshness, in order to ensure lower deteriorative reactions during enzymatic hydrolysis, which is needed to obtain protein hydrolysates with acceptable organoleptic properties and storage stability. Marine protein hydrolysates and peptides have been shown to reduce the oxidation of both the lipids and proteins of seafood during storage, thus indicating antioxidant and anti-freezing effects [21][28][29][30].

2. Fish Protein Hydrolysate: Production and Processing Factors

The fish protein hydrolysates markets are anticipated to reach USD 558 million by 2025 with a compound annual growth rate (CAGR) of over 5% due to increasing demand for protein-based supplements, food formulations, infant nutrition, fertilizers, and aquafeeds due to their higher absorption or digestion [2]. These hydrolysates contain peptides with 2–10 amino acids or up to 20 amino acids and a molecular weight of <3 kDa, especially between 0.2 to 2 kDa [2][14]. The by-products of several commercial species such as tuna [17][20][31]; tilapia [12][32][33][34][35]; marine bony species [19][36][37]; small pelagics [38][39][40]; salmonids [15][16][41][42]; shrimp [18][43][44]; marine invertebrates such as mollusks [45][46][47], squid, and cuttlefish [48][49][50][51][52][53]; and underutilized fish [54][55] have been used as the sources for producing protein hydrolysates. By-products from these species showed differences in their compositions (i.e., different amounts of lipids, blood, proteins, undigested feed in their stomach or intestines, etc.), emphasizing the need to adjust the hydrolysis conditions for each source according to the inherent characteristics of the by-products (such as sorting by-products to obtain different fractions). Depending on the farming conditions or ocean water quality, there are some safety issues with by-products from different categories of aquatic species that should be considered. For example, undesirable metabolites and contaminants may enter the liquid phase in which peptides are formed, affecting the safety of the protein hydrolysates.
Enzymatic and autolytic hydrolysis and microbial fermentation have been used for converting the proteins of by-products into peptides with varying sizes and bioactivities [2][7][8][56]. The process of hydrolysis with commercial proteases can be controlled by selecting appropriate enzymes and adjusting the hydrolysis conditions (time, temperature, enzyme concentrations, water ratio, etc.). The final hydrolysates should be stable in terms of structure (such as peptide profile) and functions. Autolytic hydrolysis mediated by endogenous enzymes such as acid/aspartyl proteases (such as pepsin), serine proteases (trypsin and chymotrypsin), thiol/cysteine protease (cathepsin B, L, and S), and muscle proteases (lysosomal cathepsin, alkaline, and neutral protease) has been used to hydrolyze marine proteins [57][58]. However, acidic proteases have a lesser role in autolysis because most studies have shown that hydrolysis by endogenous proteases usually occurs at around neutral pH [2] unless this group of proteases is first isolated. Then, their effects are investigated in acidic pH [59][60][61], although the cost associated with purification of viscera proteases, their stability and activity may be a challenge compared to highly stable commercial enzymes.

2.1. By-Product Composition, Quality, Storage and Handling

The type of by-products (viscera, heads, frames, trimmings, or their mixtures) and their composition (different amounts of lipids, blood, hemoglobin, and metal ions as pro-oxidants) influence the composition, nutritional, and antioxidant properties of the protein hydrolysates [15][62]. Frames, trimmings and heads are clean by-products with potential uses for direct human consumption or as functional hydrolysates. A recent study has shown that 37.2, 56.7, and 81.0% of the weight of heads, frames, and trimmings of Atlantic salmon, respectively, are edible as direct human food while for the viscera, the edible yield is normally thought of as 0% [10], due to the presence of lipids, bile acids, blood, non-digested feed in the intestine, etc., which is associated with a high rate of oxidation and the formation of undesirable metabolites during hydrolysis [3]. According to a survey of methods to utilize Scottish salmon by-products, 15% of the final utilization (frames, trimmings, heads) is food, 75% (frames, heads, viscera, mixed by-products, skins) is for feed purposes, and 10% (blood) is for fuel and fertilizer [6]. Therefore, cleaner by-products (frames, heads, trimmings) with acceptable freshness are ideal for food purposes, including the production of protein hydrolysates, while viscera (alone) is used for producing hydrolyzed protein concentrate and oil or fish meal and oil (rendering) in the form of mixed by-products. Some of the oxidation products can add carbonyl derivatives to the peptides, decreasing their antioxidant activity and possibly their safety.
The freshness of by-products is one factor affecting the formation of oxidative compounds during hydrolysis, and eventually the structures, functions, and shelf life of protein hydrolysates [2][3]. Endogenous muscle proteases including matrix metalloprotease (MMP), and serine and cysteine proteases degrade myofibrillar proteins and microfibrillar networks, resulting in a significant decrease in quality [63][64][65]. These proteases are active in trimmings, frames, and heads, which with inappropriate storage temperatures, degrade myofibrillar proteins, lowering the initial quality of proteins for further hydrolysis. A recent survey of fishery by-products handling practices in Europe indicated that the sorting of by-product fractions was performed in >60% of seafood processing plants, while the remainder did not handle their by-products properly and only 25% of the surveyed plants managed by-products as food-grade. According to that survey, the majority of the processing companies used their by-products for non-food purposes, mainly as feed [66]
The pretreatments of by-products may lower the concentration of pro-oxidants while resulting in purer protein substrates for enzymatic hydrolysis. The rinsing or incubation of herring by-products (heads, backbone with caudal fin, skin, intestines, and eggs) with antioxidant solutions including Duralox-MANC, isoascorbic acid, isoascorbic acid + ethylenediaminetetraacetic acid) decreased the rate of lipid oxidation and hemoglobin levels during storage at 4 °C for up to 12 days while extending the shelf-life of by-products from <1 to >12 days with a rinsing strategy or to >7 days with direct additions into the by-products after mincing [67]. However, due to the high rate of lipid oxidation in such a highly sensitive system, upgrading to food grade with proper safety is a big challenge. The washing or defatting of underutilized Sind sardine muscle mince significantly decreased total pigments (600 and 140 μg/g dry sample) and heme iron (5.3 and 1.2 mg/100 g dry sample) content in washed and defatted mince substrates, respectively, compared with non-pretreated mince (2570 μg/g dry sample and 23 mg/100 g dry sample for total pigments and heme iron, respectively), resulting in the lower formation of TBARS but increased the DPPH radical scavenging and ferrous chelating activities of protein hydrolysates, especially with defatted mince [54]. The pretreatment of cape hake by-products with 8 mM CaCl2 + 5 mM citric acid followed by the alkaline solubilization of proteins (pH 11) resulted in significantly lower phospholipids (1 AU/g) and lipids (0.39%) but higher solubility in hydrolysates compared to samples directly produced from by-products (4.2 AU/g and 0.87%, respectively). However, these hydrolysates were characterized by a higher yellowness and redness that was attributed to the alkaline solubilization of heme proteins during protein isolation [68]. Despite the relative improvement in by-product quality, the economic issue of carrying out such pretreatments at the tonnage scale (i.e., industrially) is a challenge that has not been studied. To tackle technological problems associated with heme proteins and to lower oxidation, the use of antioxidative extracts from agricultural waste, including lingonberry press-cake; apple, oat, barley, and shrimp by-products; and seaweed (ulva) extracts as helpers, all at 30% of the dry weight of the by-product, decreased the formation of MDA and the oxidation product 4-hydroxy-(E)-2-hexenal (HHE) in herring and salmon heads and backbone protein isolates, resulting in more stable substrates. Of all the helpers, lingonberry press-cake followed by apple peel and ulva were the most effective in reducing lipid oxidation during alkaline solubilization/acid precipitation and 9 days of ice storage. The new color (dark purple) in the resulting protein isolates with lingonberry press-cake might be advantageous for increasing the acceptance of the color by consumers [69]

2.2. Proteolytic Enzymes

The composition and sequence of peptides in whole hydrolysates from the same source of protein may differ depending on the type of enzyme used. Endopeptidases (e.g., trypsin, chymotrypsin, pepsin, pancreatin, papain, and Alcalase) act separately from the N- or C-terminus, while exopeptidases (e.g., carboxypeptidase Y, aminopeptidase M, and Flavourzyme) break peptide bonds at the terminus of polypeptide chains [70]. Since different enzymes have specific cleavage sites (papain: Arg-, Lys- and Phe-; Alcalase: Ala-, Leu-, Val-, Tyr-, Phe- and Try-; trypsin: Arg- and Lys-; pepsin: Phe- and Leu-), different cleavage sites will affect the AA composition and the sequence of peptides [71]. Salmon skin gelatin hydrolyzed with Alcalase showed a higher content of hydrophobic AAs and a degree of hydrolysis (DH), surface hydrophobicity, and peptides with MW < 1 kDa than hydrolysates produced using Neutrase, Protamex, and Flavourzyme. This was associated with significantly higher OH and O2•− scavenging and Fe2+ chelating activity [42]. From tilapia skin gelatin, different peptides, including Gly-Pro-Ala [12], Glu-Gly-Leu (317 Da) and Tyr-Gly-Asp-Glu-Tyr (645 Da) [32], Asp-Pro-Ala-Leu-Ala-Thr-Glu-Pro-Asp-Pro-Met-Pro-Phe (1383 Da) [33], Leu-Ser-Gly-Tyr-Gly-Pro (592 Da) [34], and Tyr-Gly-Thr-Gly-Leu (509 Da) and Leu-Val-Phe-Leu (490 Da) [35] were obtained depending on the enzyme used despite starting with the same protein (tilapia skin gelatin (although the method of production of the gelatin may have differed)). In abalone viscera, different enzymes resulted in different peptide sequences: Alcalase—Gln-Ser-Cys-Ala-Arg-Phe (711 Da), Ala-Ala-Pro-Ala-Val-Ser-Gly-Arg (728 Da), Asn-Arg-Phe-Gly-Val-Ser-Arg (834 Da), and Pro-Val-Pro-Pro-Tyr-Lys-Ala (770 Da); Neutrase—Ala-Ala-Gln-Tyr-Ser-Arg-Asn (808 Da), Val-His-Ala-Glu-Pro-Thr-Lys (780 Da), Gly-Cys-Tyr-Val-Pro-Lys-Cys (769 Da), and Asn-Ser-His-Val-Val-Arg (711 Da); papain—Ala-Ala-Asn-Asn-Ser-Thr-Arg (732 Da), Thr-Ile-Asp-Cys-Asp-Arg (722 Da), Cys-Ile-Gly-Tyr-Asp-Arg (725 Da), Asp-Asp-Ile-Thr-Arg-Asp (734 Da), and Asp-Val-Ala-Phe-Met-Arg (738.3 Da); and trypsin—Met-Glu-Thr-Tyr (543.3 Da), Tyr-His-Gly-Phe (523 Da), Gln-Cys-Val-Arg (505 Da) [45]. Tyr-Pro-Pro-Ala-Lys (574 Da) [46] and Pro-Ile-Ile-Ser-Val-Tyr-Trp-Lys (1005 Da) [47] were purified from blue mussels using Neutrase and pepsin, respectively. Despite the influence of the structure of peptides on the selectivity of the enzymes used, the high cost of commercial proteases suggests the minimal use of enzymes for hydrolysis.

2.3. Operating Parameters

The temperature and pH are adjusted according to the selected proteases to ensure high hydrolytic activity. Thus, other factors, such as the water-to-by-products ratio, type of propeller, stirring rate, enzyme deactivation step, use of nitrogen gas, antioxidant addition, time, etc., should be optimized to ensure a stable end product.
The Abalone food muscle hydrolyzed with papain (HPP), Protamex® (HP), or an animal protease (HA) for 0.5 or 4 h showed differences in physicochemical and structural properties governed by the time of the hydrolysis and enzyme type. The fluorescence emission spectra of all hydrolysates showed a redshift of 10–12 nm compared with that of control, while fluorescence intensity was higher in hydrolysates than non-hydrolyzed proteins (AFP). Hydrolysis for four hours resulted in higher intensities in HPP, HP, and HA compared to lower hydrolysis time. Among all samples, HA-4, HPP-0.5, and HPP-4 had higher absolute ζ-potential values than AFP, indicating a higher number of ionizable groups on the protein surface that inhibited the formation of protein aggregates, consistent with solubility, S-S bonds and free –SH groups [72].
During hydrolysis, especially at the industrial scale, the high water addition increases the production costs due to the heating and drying needed to obtain a powder or a concentrated liquid [3]. Using less water during hydrolysis can be beneficial if it does not affect other processes and efficiency. For cod head hydrolysates with different water ratios (1:1, 1:0.75, and 1:0.5 kg/kg), the ratio was found to have little effect on hydrolysis yield, protein content, and MW distribution of peptides and, thus, high water addition might be unnecessary. In addition, the hydrolysis reaction time of 1 h was suitable to obtain hydrolysates with desirable properties [73].
Lipid oxidation is one of the main challenges during the enzymatic/autolytic hydrolysis of by-products, resulting in unpleasant odors and flavors, dark colorations, and the formation of oxidative products in the FPH [2][3]. Due to the low lipid content of heads (1–4%), a separated oil faction was not formed in the cod head hydrolysates (0.65% lipids in the FPH) [73]. When working with a mixture of cod viscera and trimmings, a separate oil faction was formed after enzymatic hydrolysis, and a minimum of 6 g of lipids/100 g wet weight by-products was required to form an oil fraction [74]. The intensity of lipid oxidation during hydrolysis is different among by-products with different compositions, which may affect the structure and safety of the resulting peptides and their antioxidant activity.
Enzyme deactivation is the last step of the hydrolysis process. The high temperature (80–100 °C, 10–15 min) is often used to deactivate enzymes [11][12][13][14][15][16]. This temperature may lead to structural changes in the protein hydrolysates and peptides, decreasing antioxidant or other bioactivities.

2.4. Process Scale-Up

One roadblock to the industrial production of protein hydrolysates from by-products is that most studies with by-products were performed at the laboratory scale, which limits their industrial adaptation [75]. Some studies attempted pilot trials to confirm the technical feasibility of the laboratory scale at an industrial scale. In this sense, the process of producing peptides from hake by-catches was scaled up from a 0.5 to a 150 L reactor using the optimized hydrolysis conditions (2% enzyme, two h, 50% solids, pH 9, 70 °C) that were identified in the laboratory [76]. The authors found similar results at the pilot plant scale in terms of protein extraction yield (60.0% pilot and 61.4% lab), antioxidant capacity (172 mg TEAC/g protein in pilot and 224 mg TEAC/g protein in lab), and antioxidant capacity yield (103 mg TEAC/g protein in pilot and 132 mg TEAC/g protein in lab). Furthermore, liquid, solid, and bone yield did not show any significant differences from the results of the laboratory trials. Monkfish by-products (heads and viscera) hydrolysis was scaled up from 100 mL to a 5 L glass reactor at optimized laboratory conditions: 57 °C, pH 8.3, solid to liquid (S/L) ratio of 1:1 (w/w), 0.05% Alcalase, and a 200 rpm stirring rate for three hours. Following hydrolysis, the hydrolysates were filtered (100 μm) to remove non-hydrolyzed materials.

3. Antioxidant Activity of Fishery By-Products Protein Hydrolysates and Peptides

The antioxidant peptides had 2–10 amino acids, although some had up to 20 amino acids and had a MW of 0.2 to 2 kDa. The antioxidant activity of peptides is mainly related to the presence and position of specific amino acid residues in the peptide chain. Primary structure, amino acid composition, hydrophobicity, spatial conformation, etc., are characteristics that are affected by enzymatic hydrolysis and determine its antioxidant activity [77]. Peptides show antioxidant activity due to the presence of one or more hydrophobic (Pro, Ala, Gly, Leu, Ile, Met, Trp, Phe, Val) and aromatic (Tyr, Trp, Phe) amino acids that can quench free radicals via various mechanisms, including hydrogen atom transfer (HAT), single electron transfer followed by proton transfer (SET-PT), and sequential proton loss by electron transfer (SPLET) mechanisms [78]. Scavenging free radicals and oxidants using HAT, SET-PT, and SPLET generally leads to the same end results, although the kinetics and potential for side reactions vary. SPLET, SET-PT, and HAT may occur in parallel; however, the dominant mechanism depends on the antioxidant’s conformational and geometrical features, solubility, partition coefficient, and the type of solvents [79]. The antioxidant activity of hydrophobic amino acids has been attributed to their ability to interact with lipid molecules by increasing the solubility of peptides in lipids and scavenging lipid-derived radicals through electron-donating substituents such as OH and NH2 on amino acid side chains in peptides. Hydrophobic amino acids can improve the antioxidant activity of peptides by providing a potential pool of free electrons.
Peptides containing proline-rich sequences have been identified to possess antioxidant properties. Proline has an electron-rich nitrogen-containing pyrrolidone ring that stabilizes the radical peptide formed after electron donation [80][81]. Peptides with Pro at the C-terminus (e.g., Pox: Tyr-Tyr-His-Pro) were the most potent antioxidant (0.8 TE at 2.5 μM). The modification of the structure by moving Pro into positions X1 (Pro-Tyr-Tyr-His), X2 (Tyr-Pro-Tyr-His), and X3 (Tyr-Tyr-Pro-His), but leaving the other residues in the same order as in Pox, resulted in a significant difference in ORAC, which was 0.2, 0.1, and 0.55 TE at 2.5 μM for X1, X2, and X3, respectively [82]. Intense antioxidant activity has been reported for small peptides containing amino acid residues such as Tyr, His, and Pro [83][84]. The dipeptide Tyr-Tyr at the N-terminal position of Tyr-Tyr-His-Pro and Tyr-Tyr-Pro-His was the portion responsible for stronger antioxidant activity. However, Tyr-Pro-Tyr-His showed the weakest ORAC and inhibited ROS production by 36% at 0.07 μM in human keratinocyte cells (HaCat) after treatment with H2O2 when compared to Tyr-Tyr-His-Pro (with the highest ORAC), which showed the similar inhibition of ROS production at 2.5 μM (40%) [82]. It is believed that besides amino acid composition and sequence, the changes in secondary structure have a significant impact on the capture and dissipation of free radicals. The nanopeptide Val-Leu-Leu-Tyr-Lys-Asp-His-Cys-His (1127 Da) produced using the self-assembly of pine nut Val-Leu-Leu-Tyr (506 Da) and sea cucumber Lys-Asp-His-Cys-His (638 Da) had significantly higher antioxidant activity compared to individual peptides due to changes in the secondary structure as seen in the lower electron paramagnetic resonance (EPR) signal, higher random crimp degree, and increased supply of hydrogen protons (i.e., the higher exposure to active hydrogen) from Raman spectroscopy and 1H NMR spectrum analysis in the nanopeptide than the tetrapeptide and the pentapeptide
The presence of Tyr at the N-terminal position of peptide Tyr-Ala-Glu-Glu-Arg-Tyr-Pro-Ile-Leu has been reported as the residue that most contributed to antioxidant activity (3.8 μM TE/mg protein). However, Tyr-Pro-Ile and Tyr-Gln-Ile-Gly-Leu with Tyr at the same position showed lower ORAC (1.6 and 1.7 μM TE/mg protein, respectively), indicating the role of adjacent amino acids and chain length on antioxidant activity [85]. Tyr-containing peptides from abalone viscera showed strong ABTS radical scavenging activity in the order of Cys-Ile-Gly-Tyr-Asp-Arg (0.144 mg/mL) > Tyr-His-Gly-Phe (0.268 mg/mL) > and Gly-Cys-Tyr-Val-Pro-Lys-Cys (0.389 mg/mL). The first and last peptides, which contained both Tyr and Cys, showed similar trends for scavenging DPPH radicals (IC50 of 0.207 and 0.405 mg/mL, respectively). Despite the observed high ABTS radical scavenging activity, peptides Met-Glu-Thr-Tyr and Tyr-His-Gly-Phe, which had Tyr at the C- or N-terminal position, respectively, had weak scavenging activity against DPPH radicals (<20%), which was attributed to the lack of Cys in their sequence. Despite having different sizes or amino acids residues, peptides Gln-Cys-Val-Arg and Gln-Ser-Cys-Ala-Arg-Phe showed similar DPPH radical scavenging activity (IC50 of 0.392 and 0.416 mg/mL, respectively), indicating the complexity of the relationship between the peptides’ structures and function. Regarding the number of amino acid residues within peptide sequence, the peptide Gly-Cys-Tyr-Val-Pro-Lys-Cys, containing two Cys residues, showed lower free radical scavenging activity (IC50 of 0.389 and 0.405 mg/mL for scavenging ABTS and DHHP radicals, respectively) than Cys-Ile-Gly-Tyr-Asp-Arg, which contained only one Cys (IC50 of 0.144 and 0.207 mg/mL for scavenging ABTS and DPPH radicals, respectively [86]
It was shown that the presence of a Tyr, Trp, Cys, or Met residue with electron/hydrogen donating ability was the driving force for dipeptides to scavenge radicals. The presence of Tyr, Trp, and Cys in the sequence was required for dipeptides to scavenge ABTS•+, while the presence of Tyr, Trp, Cys, and Met was needed for dipeptides to scavenge ROO when using the ORAC assay. Structure–activity relationships showed that Tyr- and Trp-containing dipeptides with Tyr/Trp residue at the N-terminus (Tyr/Trp-X; Tyr-Gly, Tyr-Ser, Tyr-Gln, Tyr-Glu) had stronger ORAC and ABTS•+ scavenging activity than that at the C-terminus (X-Tyr/Trp, Gly-Tyr, and Glu-Tyr) and the steric effects, hydrophobicity, and hydrogen bonding also affected the neighboring AAs. Tyr-containing dipeptides showed higher ABTS•+ scavenging activity. In contrast, Trp dipeptides (Trp-Gly, Trp-Ser, Trp-Gln, Trp-Glu, Gly-Trp, Glu-Trp) had higher ORAC, and only Cys-containing dipeptides showed moderate reducing power activities [87]. The calculation of BDE, IP, PA, and ETE of Tyr/Trp-X and X-Tyr/Trp (where X was Gly, Leu, Pro, Phe, Ser, Thr, Asn, Gln, Asp, Glu, Lys, and Arg) showed that there were few differences among dipeptides, indicating that the neighboring AAs did not affect the intrinsic hydrogen or electron-donating ability of the dipeptides studied. Thus, the differences in their radical scavenging activity can be attributed to other factors (such as steric effects and inter/intra-molecular hydrogen bonds).

4. Application of Fish By-Products Protein Hydrolysates to Control Oxidative Deteriorations of Seafood

The oxidation of lipids is often the major cause of the quality loss of foods during storage, as seen in the changes in color, texture, flavor, and aroma, which impairs sensory and nutritional properties and the shelf-life of foods [28]. The decomposition of the hydroperoxides formed by pro-oxidative metal ions is a driving factor for lipid oxidation, producing highly reactive alkoxyl lipid radicals and hydroxyl ions. Alkoxyl radicals degrade rapidly to form volatile decomposition products, often with off-odors [29]. Furthermore, protein carbonyls can be introduced into proteins using a covalent linkage of lipid carbonyls (e.g., protein-bound malondialdehyde). Protein oxidation leads to functional property changes such as decreased solubility, digestibility, and water-holding capacity [30]. The loss of nutrients and myofibrillar water and the changes in texture are inevitable during frozen storage [88][89]. The formation of ice crystals, associated with cell membrane rupture and muscle fibers, often leads to protein denaturation and undesirable reactions such as aggregation and decrease in solubility, solute concentration (macromolecular crowding), lipid oxidation, and instability of proteins at the ice–water interphase [90][91]. Protein hydrolysates and peptides may be potential antioxidants to reduce oxidation during food storage, thus extending the shelf life [28]. The antioxidant activity of protein hydrolysates was related to amino acid composition, sequence, size, and the amino acid residues at the C- or N-terminal positions [77]. Enzymatic hydrolysis disrupts the tertiary structure of food proteins, leading to the increase in the solvent accessibility of peptides to scavenge free radicals and chelate pro-oxidative metal ions. Protein hydrolysates and peptides have been reported to control food oxidation through various mechanisms, including inactivating ROS, scavenging free radicals, chelating pro-oxidative metal ions, reducing lipid hydroperoxides, and changing the physical state of foods.

5. Conclusions

Marine by-products have been studied as a source of antioxidant peptides for food, feed, and nutraceutical applications. Those studies recommended these peptides as potential functional ingredients to enhance health and nutrition. However, the differences in the composition of by-products, the type of proteases used, and the different hydrolysis parameters resulted in various end products from the same protein. To ensure consistency, the process of upgrading at three levels, i.e., by-products, enzymes, and operating parameters must be optimized for each by-product source.
Different peptides were produced by different enzymes. However, it has not been determined which peptide is a more potent antioxidant in controlling oxidation in which food system. Different foods, due to inherent composition differences (i.e., different amounts of pro-oxidants, oxidation-prone substances, and internal antioxidant enzymes), will probably have different reactions to antioxidant peptides and specific structures during storage. Therefore, the effect of peptides with specific structures or protein hydrolysates produced using a specific condition in different food matrices needs to be investigated.
Although the structure of peptides was influenced by the specificity of the proteases used, most studies used fresh by-products with acceptable initial quality. It is less known whether the same peptides structure and function can be obtained using previously stored by-products and how the quality of proteins in refrigerated or frozen by-products, as well as associated chemical reactions during storage, affect hydrolysis and product structure, function, and stability. This area needs to be further investigated, especially in by-products with high lipid and blood contents, such as herring and salmonid by-products, and to understand which fractions are more oxidized and contribute more significantly to undesirable biochemical reactions during hydrolysis. Several researchers have tried to stabilize by-products before upgrading them using antioxidants. There is a price to adding synthetic antioxidants or to maintaining the initial quality and safety of agricultural wastes as sources of antioxidative extracts. The practical ability to undertake this process for large quantities of by-products and the space and energy consumption required to create low temperature storage are among the issues that make this application more complicated.
By-product processing should be carried out near fish production and processing centers so the hydrolysis of by-products can be performed within the shortest possible time. When the quantity of by-products exceeds a center’s capacity, they should be stored frozen. Therefore, the impact of frozen storage at varying lengths of time on the structure of peptides (amino acid composition, sequence, and size) and the occurrence of undesirable oxidative deteriorations and biochemical changes that may affect biological functions needs further study. There is limited information about the effect of initial protein quality due to processing and storage on the functional and biological activities of hydrolysates and peptides for food applications.
Many studies are underway regarding the use of fish protein hydrolysates in food. Nevertheless, the supply of fresh raw materials with acceptable safety, competitive prices with other commercial ingredients from plants and other sources, and the lack of efficient and standardized techniques to transform fish by-products into marketable forms limit their utilization.
Although the role of protein hydrolysates in maintaining the quality of seafood products has been shown, a standard method for its production from a specific source of marine by-products on a pilot or an industrial scale and its industrial application has not yet been undertaken. However, the industry prefers to use synthetic preservatives with lower prices to maintain the quality of seafood products during storage.
Protein hydrolysates may also affect the sensory characteristics of food. So, how to mask or remove the fish smell using encapsulating methods and the cost of such pretreatments on protein hydrolysates and the market must be addressed. Furthermore, hygroscopicity, the development of bitterness during enzymatic hydrolysis, low storage stability, and hydrolysis in the gastrointestinal tract (GIT) may pose several challenges to the application of fish protein hydrolysates in the food industry [92][93]. Several encapsulating techniques, including liposomes, nanohydrogels, emulsions, and diphasic gel double emulsions, have been used to improve the storage and gastrointestinal stability of protein hydrolysates [94][95]

References

  1. FAO. The state of world fisheries and aquaculture 2020. In Sustainability in Action; FAO: Rome, Italy, 2020; Available online: https://www.fao.org/3/ca9231en/CA9231EN.pdf (accessed on 10 December 2023).
  2. Nikoo, M.; Benjakul, S.; Ahmadi Gavlighi, H. Protein hydrolysates derived from aquaculture and marine byproducts through autolytic hydrolysis. Compr. Rev. Food Sci. Food Saf. 2022, 21, 4872–4899.
  3. Nikoo, M.; Regenstein, J.M.; Haghi Vayghan, A.; Walayat, N. Formation of oxidative compounds during enzymatic hydrolysis of byproducts of the seafood industry. Processes 2023, 11, 543.
  4. Wassef, E.; Saleh, N.; Abde-Meguid, N.; Abdel-Mohsen, H. Utilization of fish waste biomass as a fishmeal alternative in European seabass (Dicentrarchus labrax) diets: Effects on immuno-competence and liver and intestinal histomorphology. Int. Aquat. Res. 2023, 15, 145–161.
  5. Mohd Noor, M.I.; Azra, M.N.; Lim, V.C.; Zaini, A.A.; Dali, F.; Hashim, I.M.; Hamzah, H.C.; Abdullah, M.F. Aquaculture research in Southeast Asia—A scientometric analysis (1990–2019). Int. Aquat. Res. 2021, 13, 271–288.
  6. Stevens, J.R.; Newton, R.W.; Tlusty, M.; Little, D.C. The rise of aquaculture by-products: Increasing food production, value, and sustainability through strategic utilisation. Mar. Policy 2018, 90, 115–124.
  7. Mutalipassi, M.; Esposito, R.; Ruocco, N.; Viel, T.; Costantini, M.; Zupo, V. Bioactive compounds of nutraceutical value from fishery and aquaculture discards. Foods 2021, 10, 1495.
  8. Siddiqui, S.A.; Schulte, H.; Pleissner, D.; Schönfelder, S.; Kvangarsnes, K.; Dauksas, E.; Rustad, T.; Cropotova, J.; Heinz, V.; Smetana, S. Transformation of seafood side-streams and residuals into valuable products. Foods 2023, 12, 422.
  9. Henriques, A.; Vázquez, J.A.; Valcarcel, J.; Mendes, R.; Bandarra, N.M.; Pires, C. Characterization of protein hydrolysates from fish discards and by-products from the North-West Spain fishing fleet as potential sources of bioactive peptides. Mar. Drugs 2021, 19, 338.
  10. Malcorps, W.; Newton, R.W.; Sprague, M.; Glencross, B.D.; Little, D.C. Nutritional characterisation of European aquaculture processing by-products to facilitate strategic utilisation. Front. Sustain. Food Syst. 2021, 5, 378.
  11. Nikoo, M.; Benjakul, S.; Ehsani, A.; Li, J.; Wu, F.; Yang, N.; Xu, B.; Jin, Z.; Xu, X. Antioxidant and cryoprotective effects of a tetrapeptide isolated from Amur sturgeon skin gelatin. J. Funct. Foods 2014, 7, 609–620.
  12. Zheng, L.; Yu, H.; Wei, H.; Xing, Q.; Zou, Y.; Zhou, Y.; Peng, J. Antioxidative peptides of hydrolysate prepared from fish skin gelatin using ginger protease activate antioxidant response element-mediated gene transcription in IPEC-J2 cells. J. Funct. Foods 2018, 51, 104–112.
  13. Nong, N.T.P.; Hsu, J.L. Bioactive peptides: An understanding from current screening methodology. Processes 2022, 10, 1114.
  14. Cunha, S.A.; Pintado, M.E. Bioactive peptides derived from marine sources: Biological and functional properties. Trends Food Sci. Technol. 2022, 119, 348–370.
  15. Opheim, M.; Šližytė, R.; Sterten, H.; Provan, F.; Larssen, E.; Kjos, N.P. 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. Biochem. 2015, 50, 1247–1257.
  16. Vázquez, J.A.; Sotelo, C.G.; Sanz, N.; Pérez-Martín, R.I.; Rodríguez-Amado, I.; Valcarcel, J. Valorization of aquaculture by-products of salmonids to produce enzymatic hydrolysates: Process optimization, chemical characterization and evaluation of bioactives. Mar. Drugs 2019, 17, 676.
  17. Vázquez, J.A.; Pedreira, A.; Durán, S.; Cabanelas, D.; Souto-Montero, P.; Martínez, P.; Mulet, M.; Pérez-Martín, R.I.; Valcarcel, J. Biorefinery for tuna head wastes: Production of protein hydrolysates, high-quality oils, minerals and bacterial peptones. J. Clean Prod. 2022, 357, 131909.
  18. Nikoo, M.; Xu, X.; Regenstein, J.M.; Noori, F. Autolysis of Pacific white shrimp (Litopenaeus vannamei) processing by-products: Enzymatic activities, lipid and protein oxidation, and antioxidant activity of hydrolysates. Food Biosci. 2021, 39, 100844.
  19. Valcarcel, J.; Sanz, N.; Vázquez, J.A. Optimization of the enzymatic protein hydrolysis of by-products from seabream (Sparus aurata) and seabass (Dicentrarchus labrax), chemical and functional characterization. Foods 2020, 9, 1503.
  20. Zhang, S.Y.; Zhao, Y.Q.; Wang, Y.M.; Yang, X.R.; Chi, C.F.; Wang, B. Gelatins and antioxidant peptides from Skipjack tuna (Katsuwonus pelamis) skins: Purification, characterization, and cytoprotection on ultraviolet-A injured human skin fibroblasts. Food Biosci. 2022, 50, 102138.
  21. Ngo, D.H.; Kim, S.K. Marine bioactive peptides as potential antioxidants. Curr. Protein Pept. Sci. 2013, 14, 189–198.
  22. Wang, T.; Zhao, Q.; Wang, Q. Production and antioxidant properties of marine-derived bioactive peptides. In Marine Proteins and Peptides: Biological Activities and Applications; Wiley Online Library: Hoboken, NJ, USA, 2013; pp. 385–406.
  23. Rodríguez-Forero, A.; Medina-Lambraño, K.; Acosta-Ortíz, E. Variations in the proximate composition of the sea cucumber, Isostichopus sp. aff badionotus. Int. Aquat. Res. 2021, 13, 241–252.
  24. Aluko, R.E. Amino acids, peptides, and proteins as antioxidants for food preservation. In Handbook of Antioxidants for Food Preservation; Woodhead Publishing: Sawston, UK, 2015; pp. 105–140.
  25. Nikoo, M.; Benjakul, S. Potential application of seafood-derived peptides as bifunctional ingredients, antioxidant–cryoprotectant: A review. J. Funct. Foods 2015, 19, 753–764.
  26. Saiwong, S.; Autsavapromporn, N.; Siriwoharn, T.; Techapun, C.; Wangtueai, S. Enzymatic hydrolysis optimization for preparation of sea cucumber (Holothuria scabra) hydrolysate with an antiproliferative effect on the HepG2 liver cancer cell line and antioxidant properties. Int. J. Mol. Sci. 2023, 24, 9491.
  27. Liu, H.; Liang, J.; Xiao, G.; Vargas-De-La-Cruz, C.; Simal-Gandara, J.; Xiao, J.; Wang, Q. Active sites of peptides Asp-Asp-Asp-Tyr and Asp-Tyr-Asp-Asp protect against cellular oxidative stress. Food Chem. 2022, 366, 130626.
  28. Tang, T.; Wu, N.; Tang, S.; Xiao, N.; Jiang, Y.; Tu, Y.; Xu, M. Industrial application of protein hydrolysates in food. J. Agric. Food Chem. 2023, 71, 1788–1801.
  29. Jacobsen, C.; Horn, A.F.; Sørensen, A.D.M.; Farvin, K.S.; Nielsen, N.S. Antioxidative strategies to minimize oxidation in formulated food systems containing fish oils and omega-3 fatty acids. In Antioxidants and Functional Components in Aquatic Foods; Wiley Online Library: Hoboken, NJ, USA, 2014; pp. 127–150.
  30. Estévez, M.; Xiong, Y. Protein oxidation in foods: Mechanisms, consequences, and antioxidant solutions. Foods 2021, 10, 2346.
  31. Zhang, J.B.; Zhao, Y.Q.; Wang, Y.M.; Chi, C.F.; Wang, B. Eight collagen peptides from hydrolysate fraction of Spanish mackerel skins: Isolation, identification, and in vitro antioxidant activity evaluation. Mar. Drugs 2019, 17, 224.
  32. Zhang, Y.; Duan, X.; Zhuang, Y. Purification and characterization of novel antioxidant peptides from enzymatic hydrolysates of tilapia (Oreochromis niloticus) skin gelatin. Peptides 2012, 38, 13–21.
  33. Ngo, D.H.; Qian, Z.J.; Ryu, B.; Park, J.W.; Kim, S.K. In vitro antioxidant activity of a peptide isolated from Nile tilapia (Oreochromis niloticus) scale gelatin in free radical-mediated oxidative systems. J. Funct. Foods 2010, 2, 107–117.
  34. Sun, L.; Zhang, Y.; Zhuang, Y. Antiphotoaging effect and purification of an antioxidant peptide from tilapia (Oreochromis niloticus) gelatin peptides. J. Funct. Foods 2013, 5, 154–162.
  35. Bingtong, L.; Yongliang, Z.; Liping, S. Identification and characterization of the peptides with calcium-binding capacity from tilapia (Oreochromis niloticus) skin gelatin enzymatic hydrolysates. J. Food Sci. 2020, 85, 114–122.
  36. Chi, C.F.; Wang, B.; Wang, Y.M.; Zhang, B.; Deng, S.G. Isolation and characterization of three antioxidant peptides from protein hydrolysate of bluefin leatherjacket (Navodon septentrionalis) heads. J. Funct. Foods 2015, 12, 1–10.
  37. Zhao, W.H.; Luo, Q.B.; Pan, X.; Chi, C.F.; Sun, K.L.; Wang, B. Preparation, identification, and activity evaluation of ten antioxidant peptides from protein hydrolysate of swim bladders of miiuy croaker (Miichthys miiuy). J. Funct. Foods 2018, 47, 503–511.
  38. Ovissipour, M.; Rasco, B.; Shiroodi, S.G.; Modanlow, M.; Gholami, S.; Nemati, M. Antioxidant activity of protein hydrolysates from whole anchovy sprat (Clupeonella engrauliformis) prepared using endogenous enzymes and commercial proteases. J. Sci. Food Agric. 2013, 93, 1718–1726.
  39. Rivero-Pino, F.; Espejo-Carpio, F.J.; Guadix, E.M. Production and identification of dipeptidyl peptidase IV (DPP-IV) inhibitory peptides from discarded Sardine pilchardus protein. Food Chem. 2020, 328, 127096.
  40. Chiodza, K.; Goosen, N.J. Evaluation of handling and storage stability of spray dried protein hydrolysates from sardine (Sardina pilchardus) processing by-products: Effect of enzymatic hydrolysis time, spray drying temperature and maltodextrin concentration. Food Bioprod. Process. 2023, 141, 1–22.
  41. Nikoo, M.; Benjakul, S.; Yasemi, M.; Gavlighi, H.A.; Xu, X. Hydrolysates from rainbow trout (Oncorhynchus mykiss) processing by-product with different pretreatments: Antioxidant activity and their effect on lipid and protein oxidation of raw fish emulsion. Lwt 2019, 108, 120–128.
  42. Zhang, X.; Dai, Z.; Zhang, Y.; Dong, Y.; Hu, X. Structural characteristics and stability of salmon skin protein hydrolysates obtained with different proteases. LWT Food Sci. Technol. 2022, 153, 112460.
  43. Tkaczewska, J.; Kulawik, P.; Jamróz, E.; Čagalj, M.; Matas, R.F.; Šimat, V. Valorization of prawn/shrimp shell waste through the production of biologically active components for functional food purposes. J. Sci. Food Agric. 2023, 104, 707–715.
  44. Leiva-Portilla, D.; Martínez, R.; Bernal, C. Valorization of shrimp (Heterocarpus reedi) processing waste via enzymatic hydrolysis: Protein extractions, hydrolysates and antioxidant peptide fractions. Biocatal. Agric. Biotechnol. 2023, 48, 102625.
  45. Hu, Y.; Yang, J.; He, C.; Wei, H.; Wu, G.; Xiong, H.; Ma, Y. Fractionation and purification of antioxidant peptides from abalone viscera by a combination of Sephadex G-15 and Toyopearl HW-40F chromatography. Int. J. Food Sci. Technol. 2022, 57, 1218–1225.
  46. Wang, B.; Li, L.; Chi, C.F.; Ma, J.H.; Luo, H.Y.; Xu, Y.F. Purification and characterisation of a novel antioxidant peptide derived from blue mussel (Mytilus edulis) protein hydrolysate. Food Chem. 2013, 138, 1713–1719.
  47. Park, S.Y.; Kim, Y.S.; Ahn, C.B.; Je, J.Y. Partial purification and identification of three antioxidant peptides with hepatoprotective effects from blue mussel (Mytilus edulis) hydrolysate by peptic hydrolysis. J Funct. Foods 2016, 20, 88–95.
  48. Mendis, E.; Rajapakse, N.; Byun, H.G.; Kim, S.K. Investigation of jumbo squid (Dosidicus gigas) skin gelatin peptides for their in vitro antioxidant effects. Life Sci. 2005, 77, 2166–2178.
  49. Sukkhown, P.; Jangchud, K.; Lorjaroenphon, Y.; Pirak, T. Flavored-functional protein hydrolysates from enzymatic hydrolysis of dried squid by-products: Effect of drying method. Food Hydrocoll. 2018, 76, 103–112.
  50. Shavandi, A.; Hu, Z.; Teh, S.; Zhao, J.; Carne, A.; Bekhit, A.; Bekhit, A.E.D.A. Antioxidant and functional properties of protein hydrolysates obtained from squid pen chitosan extraction effluent. Food Chem. 2017, 227, 194–201.
  51. Soufi-Kechaou, E.; Derouiniot-Chaplin, M.; Amar, R.B.; Jaouen, P.; Berge, J.P. Recovery of valuable marine compounds from cuttlefish by-product hydrolysates: Combination of enzyme bioreactor and membrane technologies: Fractionation of cuttlefish protein hydrolysates by ultrafiltration: Impact on peptidic populations. C. R. Chim. 2017, 20, 975–985.
  52. Cudennec, B.; Balti, R.; Ravallec, R.; Caron, J.; Bougatef, A.; Dhulster, P.; Nedjar, N. In vitro evidence for gut hormone stimulation release and dipeptidyl-peptidase IV inhibitory activity of protein hydrolysate obtained from cuttlefish (Sepia officinalis) viscera. Food Res. Int. 2015, 78, 238–245.
  53. Kchaou, H.; Jridi, M.; Benbettaieb, N.; Debeaufort, F.; Nasri, M. Bioactive films based on cuttlefish (Sepia officinalis) skin gelatin incorporated with cuttlefish protein hydrolysates: Physicochemical characterization and antioxidant properties. Food Packag. Shelf Life 2020, 24, 100477.
  54. Sarteshnizi, R.A.; Sahari, M.A.; Gavlighi, H.A.; Regenstein, J.M.; Nikoo, M. Antioxidant activity of Sind sardine hydrolysates with pistachio green hull (PGH) extracts. Food Biosci. 2019, 27, 37–45.
  55. Tadesse, S.A.; Emire, S.A.; Barea, P.; Illera, A.E.; Melgosa, R.; Beltrán, S.; Sanz, M.T. Valorisation of low-valued ray-finned fish (Labeobarbus nedgia) by enzymatic hydrolysis to obtain fish-discarded protein hydrolysates as functional foods. Food Bioprod. Process. 2023, 141, 167–184.
  56. Nasri, R.; Abdelhedi, O.; Nasri, M.; Jridi, M. Fermented protein hydrolysates: Biological activities and applications. Curr. Opin. Food Sci. 2022, 43, 120–127.
  57. Simpson, B.K.; Nollet, L.M.; Toldrá, F.; Benjakul, S.; Paliyath, G.; Hui, Y.H. (Eds.) Food Biochemistry and Food Processing; John Wiley & Sons: Hoboken, NJ, USA, 2012.
  58. Klomklao, S.; Benjakul, S.; Simpson, B.K. Seafood enzymes: Biochemical properties and their impact on quality. In Food Biochemistry and Food Processing; Simpson, B.K., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2012; pp. 263–274.
  59. Kuepethkaew, S.; Zhang, Y.; Kishimura, H.; Kumagai, Y.; Simpson, B.K.; Benjakul, S.; Damodaran, S.; Klomklao, S. Enzymological characteristics of pepsinogens and pepsins purified from lizardfish (Saurida micropectoralis) stomach. Food Chem. 2022, 366, 130532.
  60. Nalinanon, S.; Benjakul, S.; Kishimura, H.; Osako, K. Type I collagen from the skin of ornate threadfin bream (Nemipterus hexodon): Characteristics and effect of pepsin hydrolysis. Food Chem. 2011, 125, 500–507.
  61. Ashaolu, T.J.; Lee, C.C.; Ashaolu, J.O.; Tarhan, O.; Pourjafar, H.; Jafari, S.M. Pepsin: An excellent proteolytic enzyme for the production of bioactive peptides. Food. Rev. Int. 2023, 1–38.
  62. Vázquez, J.A.; Valcarcel, J.; Sapatinha, M.; Bandarra, N.M.; Mendes, R.; Pires, C. Effect of the season on the production and chemical properties of fish protein hydrolysates and high-quality oils obtained from gurnard (Trigla spp.) by-products. LWT 2023, 177, 114576.
  63. Huang, X.; Wang, H.; Tu, Z. A comprehensive review of the control and utilization of animal aquatic products by autolysis-based processes: Mechanism, process, factors, and application. Food Res. Int. 2022, 164, 112325.
  64. Liu, B.; Liu, Z.Q.; Li, D.Y.; Yu, M.M.; Liu, Y.X.; Qin, L.; Zhou, D.-Y.; Shahidi, F.; Zhu, B.W. Action of endogenous proteases on texture deterioration of the bay scallop (Argopecten irradians) adductor muscle during cold storage and its mechanism. Food Chem. 2020, 323, 126790.
  65. Liu, Z.Q.; Zhou, D.Y.; Liu, Y.X.; Yu, M.M.; Liu, B.; Song, L.; Dong, X.-P.; Qi, H.; Shahidi, F. Inhibitory effect of natural metal ion chelators on the autolysis of sea cucumber (Stichopus japonicus) and its mechanism. Food Res. Int. 2020, 133, 109205.
  66. Abdollahi, M.; Hyldig, G.; Sørensen, A.D.M.; Sørensen, R.; Iñarra, B.; Chastagnol, C.B.; Gutierrez, M.; San Martin, D.; Simonsen, A.; Dam, L.; et al. Hurdles and Bottlenecks in Maintaining and Value Adding of Seafood Side-Streams. Waseabi. 2020, pp. 1–45, Grant agreement No 837726. Available online: https://www.waseabi.eu/media/zq5mau35/waseabi-deliverable-d1-1_industry_version-1-2_april-2021.pdf (accessed on 10 December 2023).
  67. Wu, H.; Ghirmai, S.; Undeland, I. Stabilization of herring (Clupea harengus) by-products against lipid oxidation by rinsing and incubation with antioxidant solutions. Food Chem. 2022, 316, 126337.
  68. Pires, C.; Teixeira, B.; Cardoso, C.; Mendes, R.; Nunes, M.L.; 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. Biochem. 2015, 50, 1006–1015.
  69. Zhang, J.; Abdollahi, M.; Alminger, M.; Undeland, I. Cross-processing herring and salmon co-products with agricultural and marine side-streams or seaweeds produces protein isolates more stable towards lipid oxidation. Food Chem. 2022, 382, 132314.
  70. Irankunda, R.; Camaño Echavarría, J.A.; Paris, C.; Stefan, L.; Desobry, S.; Selmeczi, K.; Muhr, L.; Canabady-Rochelle, L. Metal-chelating peptides separation using immobilized metal ion affinity chromatography: Experimental methodology and simulation. Separations 2022, 9, 370.
  71. Sun, K.L.; Gao, M.; Wang, Y.Z.; Li, X.R.; Wang, P.; Wang, B. Antioxidant peptides from protein hydrolysate of marine red algae Eucheuma cottonii: Preparation, identification, and cytoprotective mechanisms on H2O2 oxidative damaged HUVECs. Front. Microbiol. 2022, 13, 791248.
  72. Li, G.; Zuo, X.; Luo, X.; Chen, Z.; Cao, W.; Lin, H.; Qin, X.; Wu, L.; Zheng, H. Functional, physicochemical, and structural properties of the hydrolysates derived from the abalone (Haliotis discus subsp hannai Ino) foot muscle proteins. Food Chem. X 2023, 19, 100841.
  73. Remme, J.; Tveit, G.M.; Toldnes, B.; Slizyte, R.; Carvajal, A.K. Production of protein hydrolysates from cod (Gadus morhua) heads: Lab and pilot scale studies. J. Aquat. Food Prod. Technol. 2022, 31, 114–127.
  74. Šližytė, R.; Rustad, T.; Storrø, I. Enzymatic hydrolysis of cod (Gadus morhua) by-products: Optimization of yield and properties of lipid and protein fractions. Process. Biochem. 2005, 40, 3680–3692.
  75. Sasidharan, A.; Rustad, T.; Cusimano, G.M. Tuna sidestream valorization: A circular blue bioeconomy approach. Environ. Sci. Pollut. Res. 2023, 1–19.
  76. Iñarra, B.; Bald, C.; Gutierrez, M.; San Martin, D.; Zufía, J.; Ibarruri, J. Production of bioactive peptides from hake by-catches: Optimization and scale-up of enzymatic hydrolysis process. Mar. Drugs 2023, 21, 552.
  77. Elias, R.J.; Kellerby, S.S.; Decker, E.A. Antioxidant activity of proteins and peptides. Crit. Rew. Food. Sci. Nutr. 2008, 48, 430–441.
  78. Lončar, A.; Negrojević, L.; Dimitrić-Marković, J.; Dimić, D. The reactivity of neurotransmitters and their metabolites towards various nitrogen-centered radicals: Experimental, theoretical, and biotoxicity evaluation. Comput. Biol. Chem. 2021, 95, 107573.
  79. Shahidi, F.; Zhong, Y. Measurement of antioxidant activity. J. Funct. Foods 2015, 18, 757–781.
  80. Bamdad, F.; Shin, S.H.; Suh, J.W.; Nimalaratne, C.; Sunwoo, H. Anti-inflammatory and antioxidant properties of casein hydrolysate produced using high hydrostatic pressure combined with proteolytic enzymes. Molecules 2017, 22, 609.
  81. Basilicata, M.G.; Pepe, G.; Adesso, S.; Ostacolo, C.; Sala, M.; Sommella, E.; Scala, M.C.; Messore, A.; Autore, G.; Marzocco, S.; et al. Antioxidant properties of buffalo-milk dairy products: A β-Lg peptide released after gastrointestinal digestion of buffalo ricotta cheese reduces oxidative stress in intestinal epithelial cells. Int. J. Mol. Sci. 2018, 19, 1955.
  82. Sandomenico, A.; Severino, V.; Apone, F.; De Lucia, A.; Caporale, A.; Doti, N.; Russo, A.; Russo, R.; Rega, C.; Del Giacco, T.; et al. Trifluoroacetylated tyrosine-rich D-tetrapeptides have potent antioxidant activity. Peptides 2017, 89, 50–59.
  83. Mirzapour-Kouhdasht, A.; Moosavi-Nasab, M.; Lee, C.W.; Yun, H.; Eun, J.B. Structure–function engineering of novel fish gelatin-derived multifunctional peptides using high-resolution peptidomics and bioinformatics. Sci. Rep. 2021, 11, 7401.
  84. Zhang, S.; Dong, L.; Bao, Z.; Lin, S. C-terminal modification on the immunomodulatory activity, antioxidant activity, and structure–activity relationship of pulsed electric field (PEF)-treated pine nut peptide. Foods 2022, 11, 2649.
  85. Davalos, A.; Miguel, M.; Bartolome, B.; Lopez-Fandino, R. Antioxidant activity of peptides derived from egg white proteins by enzymatic hydrolysis. J. Food Prot. 2004, 67, 1939–1944.
  86. Liu, J.; Wu, G.; Yang, J.; He, C.; Xiong, H.; Ma, Y. Abalone visceral peptides containing Cys and Tyr exhibit strong in vitro antioxidant activity and cytoprotective effects against oxidative damage. Food Chem. X 2023, 17, 100582.
  87. Zheng, L.; Zhao, Y.; Dong, H.; Su, G.; Zhao, M. Structure–activity relationship of antioxidant dipeptides: Dominant role of Tyr, Trp, Cys and Met residues. J. Funct. Foods. 2016, 21, 485–496.
  88. Ma, X.; Mei, J.; Xie, J. Effects of multi-frequency ultrasound on the freezing rates, quality properties and structural characteristics of cultured large yellow croaker (Larimichthys crocea). Ultrason. Sonochem. 2021, 76, 105657.
  89. Liu, Y.; Tan, Y.; Luo, Y.; Li, X.; Hong, H. Evidence of myofibrillar protein oxidation and degradation induced by exudates during the thawing process of bighead carp fillets. Food Chem. 2024, 434, 137396.
  90. Saki, N.; Ghaffari, M.; Nikoo, M. Effect of active ice nucleation bacteria on freezing and the properties of surimi during frozen storage. LWT 2023, 176, 114548.
  91. Jia, H.; Roy, K.; Pan, J.; Mraz, J. Icy affairs: Understanding recent advancements in the freezing and frozen storage of fish. Compr. Rev. Food Sci. Food Saf. 2022, 21, 1383–1408.
  92. Sharma, K.; Nilsuwan, K.; Ma, L.; Benjakul, S. Effect of liposomal encapsulation and ultrasonication on debittering of protein hydrolysate and plastein from salmon frame. Foods 2023, 12, 761.
  93. Xu, J.; Jiang, S.; Liu, L.; Zhao, Y.; Zeng, M. Encapsulation of oyster protein hydrolysates in nanoliposomes: Vesicle characteristics, storage stability, in vitro release, and gastrointestinal digestion. J. Food Sci. 2021, 86, 960–968.
  94. Gao, Y.; Li, X.; Xie, Y.; Huang, X.; Cheng, C.; McClements, D.J.; Zhang, L.; Chen, X.; Zou, L.; Wei, L. Encapsulation of bitter peptides in diphasic gel double emulsions: Bitterness masking, sustained release and digestion stability. Food Res. Int. 2022, 162, 112205.
  95. Gao, Y.; Wu, X.; McClements, D.J.; Cheng, C.; Xie, Y.; Liang, R.; Liu, J.; Zou, L.; Liu, W. Encapsulation of bitter peptides in water-in-oil high internal phase emulsions reduces their bitterness and improves gastrointestinal stability. Food Chem. 2022, 386, 132787.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , ,
View Times: 98
Revisions: 2 times (View History)
Update Date: 20 Feb 2024
1000/1000
Video Production Service