Acessibilidade / Reportar erro

Production factors affecting antioxidant peptides from tilapia processing byproducts

Abstract

This research aimed to elucidate significant factors affecting antioxidant capacity of protein hydrolysates from tilapia processing byproducts. Effects of protein type, substrate concentration (0.4-1.2%) and time of hydrolysis (0-60 min) on antioxidant abilities were investigated. Antioxidant activity of the alkaline-aided protein hydrolysate (APH) hydrolyzed by Protease G6 at 1.2% and 60 min hydrolysis was comparable to the control (minced tilapia muscle hydrolysates) and was more effective than the myofibrillar protein and sarcoplasmic protein hydrolysates. Principal component analysis showed that the APH exert their antioxidant capacity by peroxyl radical quenching ability. These findings provide evidence that the APH from fish byproducts can be alternatively used as a natural antioxidant.

Keywords:
tilapia byproducts; myofibrillar protein; sarcoplasmic protein; alkaline-aided protein hydrolysate; natural antioxidant

1 Introduction

Byproducts from fish processing waste have attracted considerable attention especially for protein hydrolysate production. In 2016, the farmed Nile tilapia (Oreochromis niloticus ) production in Thailand was about 176,400 MT and the export volume was 7,975 MT with around 38.1% exported as chilled and frozen fillet. Approximately 50-70% byproducts are generated from the filleting process of tilapia ( Subdivision of Research and Analysis of Fisheries Statistics, 2017 Subdivision of Research and Analysis of Fisheries Statistics (2017). Nile tilapia production statistics. Thailand: Fisheries Development Policy and Strategy Division, Department of Fisheries, Ministry of Agriculture and Cooperatives. Retrieved from https://www.fisheries.go.th/strategy/UserFiles/files/tilapia%206-60(1).pdf.
https://www.fisheries.go.th/strategy/Us...
). With around 35% crude protein (dry basis), the tilapia byproducts can be used as a valuable and low-cost source for human food proteins.

Due to the harmful effects on consumer health, human consumption of synthetic antioxidants is restricted in many countries. Accordingly, several researchers have made more efforts to discover alternative natural antioxidants including antioxidant peptides and their possible use in food products ( Hraš et al., 2000 Hraš, A. R., Hadolin, M., Knez, Ž., & Bauman, D. (2000). Comparison of antioxidative and synergistic effects of rosemary extract with α-tocopherol, ascorbyl palmitate and citric acid in sunflower oil. Food Chemistry, 71(2), 229-233. http://dx.doi.org/10.1016/S0308-8146(00)00161-8.
http://dx.doi.org/10.1016/S0308-8146(00...
). Protein hydrolysates from aquatic food byproducts have been reported to exhibit antioxidant properties ( García-Moreno et al., 2014 García-Moreno, P. J., Batista, I., Pires, C., Bandarra, N. M., Espejo-Carpio, F. J., Guadix, A., & Guadix, E. M. (2014). Antioxidant activity of protein hydrolysates obtained from discarded Mediterranean fish species. Food Research International , 65, 469-476. http://dx.doi.org/10.1016/j.foodres.2014.03.061.
http://dx.doi.org/10.1016/j.foodres.201...
; Chi et al., 2015 Chi, C. F., Hu, F. Y., Wang, B., Li, Z. R., & Luo, H. Y. (2015). Influence of amino acid compositions and peptide profiles on antioxidant capacities of two protein hydrolysates from Skipjack tuna (Katsuwonus pelamis) dark muscle. Marine Drugs , 13(5), 2580-2601. http://dx.doi.org/10.3390/md13052580. PMid:25923316.
http://dx.doi.org/10.3390/md13052580 ...
). Peptides derived from suitably hydrolyzed traditional tilapia also showed a specific bioactivity or multi-functional bioactivities ( Dekkers et al., 2011 Dekkers, E., Raghavan, S., Kristinsson, H. G., & Marshall, M. R. (2011). Oxidative stability of mahi mahi red muscle dipped in tilapia protein hydrolysats. Food Chemistry , 124(2), 640-645. http://dx.doi.org/10.1016/j.foodchem.2010.06.088.
http://dx.doi.org/10.1016/j.foodchem.20...
). Various factors such as protein source, substrate concentration, enzyme type as well as condition of hydrolysis have been extensively reported to influence functionalities and bioactivities of fish protein hydrolysates. However, information regarding the comparison of antioxidant peptides derived from different protein types extracted from fish byproducts is limited. Isoelectric solubilisation-precipitation is an efficient process to recover protein from fish byproducts with high yield and simultaneously removes lipids and other impurities (skin, bone, scale, and connective tissue etc.) ( Hultin & Kelleher, 1999 Hultin, H. O., & Kelleher, S. D. (1999). U.S. Pat. No. 6,005,073. Process for isolating a protein composition from a muscle source and protein composition. Advanced Protein Technologies Inc. ). The obtained protein extract contains both myofibrillar and sarcoplasmic proteins, which would yield different antioxidant activity from each protein alone after being hydrolyzed.

The research objectives were to evaluate the effects of various protein types (sarcoplasmic, myofibrillar and alkali-aided extracted proteins from the tilapia byproducts), substrate concentration and time of hydrolysis on antioxidant properties of protein hydrolysates and to correlate them using principal component analysis (PCA).

2 Materials and methods

2.1 Materials

Frozen Nile tilapia fillet and byproducts (TB) including trimmed meat, belly flap meat, fin and head-frame were purchased from Grobest Thailand, Co., Ltd. (Nakhonphanom province, Thailand). The samples (in polyethylene bags) were stored with ice in a polystyrene box and transported by a temperature controlled truck (-12 ± 2 °C) to the Department of Food Technology, Khon Kaen University, within 5-6 h. Protease G6 (DuPont Genencor® Science, USA, a commercial alkaline endopeptidase from Bacillus licheniformis with a declared activity of 580,000 DU/g) was purchased from Siam Victory Chemicals Co., Ltd, Thailand.

2.2 Preparation of protein extracts

Both frozen tilapia fillet (control) and tilapia byproducts (TB) (head-frame, fin, trimmed meat and belly flap meat) were cut into pieces (about 1 x 1 inch2) after thawing (at 5°C, 1 h), and then minced by grinder with 5 mm diameter sieve (BIRO 8-22 E97, The BIRO MFG. Co., USA). All of TB types were mixed in 1:1:1:1 ratio (w:w:w:w) for 30 s (Crypto Peerless 10, Crypto Peerless LTD., France). Moisture contents of the fillet and mixed TB were 79.3% and 59.4%, respectively. Other compositions i.e. protein, fat, and ash contents of the fillet and mixed TB were 77.4 and 43.0%, 12.5 and 53.5%, 5.2 and 8.7% (dry basis), respectively ( Association of Official Analytical Chemists, 1999 Association of Official Analytical Chemists – AOAC. (1999). Official Methods of Analysis (16th ed.). Washington: Association of Official Analytical Chemist. ). After freeze drying (Freeze Dryer, Alpha 2-4 LD plus, Martin Christ, Germany), the mixed TB and fillet (control) powders were vacuum-packed (Henkovac Table Top 1500, Europac Co., LTD., Netherlands) in an aluminum foil bag and kept at -30 °C until used for protein extraction.

For the sarcoplasmic protein (SP) and myofibrillar protein (MP) extractions ( Dyer et al., 1950 Dyer, W. J., French, H. V., & Snow, J. M. (1950). Proteins in fish muscle. I. Extraction of protein fraction in fresh fish. Journal of the Fisheries Research Board of Canada , 7(10), 585-593. http://dx.doi.org/10.1139/f47-052.
http://dx.doi.org/10.1139/f47-052 ...
with slight modification), the TB powder was homogenized (Ace, Nihonseiki Kaisha Co., Japan) with 5 parts of sodium phosphate buffer (50 mM, pH 7.0) and with 7 parts of 0.6 M NaCl solution in sodium phosphate buffer (50 mM, pH 7.0), respectively. The supernatants were collected and freeze dried (Freeze dryer, Alpha 2-4 LD plus, Martin Christ, Germany). The alkali-aided extracted protein (AP) was prepared from the TB using the method of Hultin & Kelleher (1999) Hultin, H. O., & Kelleher, S. D. (1999). U.S. Pat. No. 6,005,073. Process for isolating a protein composition from a muscle source and protein composition. Advanced Protein Technologies Inc. with slight modifications. In brief, TB powder was mixed with cold deionized water (1:9) and homogenized at 8,000 rpm for 1 min (at 4 °C). The protein solubilization was conducted at pH 11.0 using 2 N NaOH (digital pH meter, Mettler-Toledo FE20-I, Switzerland). After centrifugation (10,000 x g, 20 min, 4 °C), the attained supernatant was adjusted to 5.5 using 2 N HCl to precipitate proteins. The protein concentrate (sediment) was adjusted to pH 7.0 with 50 mM sodium phosphate buffer and lyophilized. The three protein extracts powders were each vacuum-packed in an aluminum foil bag and kept at -30 °C until used.

2.3 Preparation of protein hydrolysates

The three lyophilized protein extracts and minced fillet (control) were hydrolyzed by Protease G6. Effect of hydrolysis times (0-60 min) and substrate concentrations (0.4, 0.6, 0.8, 1.0 & 1.2%) on antioxidant activities were determined. Other hydrolysis conditions were fixed at 1.5 ml Protease G6/100 g protein (E/S), pH 9.5 (glycine-NaOH buffer), and 65 °C in water bath (Comfort, Heto Master Shake, France). These hydrolysis conditions were chosen according to the preliminary experiment by varying pHs (7.0-10.0) and temperatures (25-65 °C) of hydrolysis to obtain the highest degree of hydrolysis and protein content. The reaction was inactivated by immersion in boiling water for 10 min. After being cooled and centrifuged (1-15 K, Higher-Speed Benchtop Microcentrifuge, Sigma-Sartorius, Germany), the supernatant was removed and filtered through 0.22 µm membrane. The aliquot of protein hydrolysate solution was kept at 4 °C until analyzed.

2.3.1 The release of TCA soluble peptides

The releases of trichloroacetic acid (TCA) soluble peptides during hydrolysis were determined ( Hoyle & Merritt, 1994 Hoyle, N. T., & Merritt, J. H. (1994). Quality of fish protein hydrolysates from herring (Clupea harengus). Journal of Food Science, 59(1), 76-79. http://dx.doi.org/10.1111/j.1365-2621.1994.tb06901.x.
http://dx.doi.org/10.1111/j.1365-2621.1...
, with slight modifications). An aliquot of the hydrolysate taken at 0, 5, 10, 15, 30, and 60 min was mixed with 20% TCA solution to a final concentration of 10%. The mixture was stood for 30 min and centrifuged at 10,000 xg for 15 min (4 °C). The total soluble protein content was determined using an automatic microtiter plate spectrophotometer (Varioskan flash reader, Thermo fisher, Finland) ( Fryer et al., 1986 Fryer, H. J., Davis, G. E., Manthorpe, M., & Varon, S. (1986). Lowry protein assay using an automatic microtiter plate spectrophotometer. Analytical Biochemistry , 153(2), 262-266. http://dx.doi.org/10.1016/0003-2697(86)90090-4. PMid:3706709.
http://dx.doi.org/10.1016/0003-2697(86)...
) as described in 2.3.2.

The rate of TCA soluble peptides released was calculated according to this equation:

Rate of soluble peptide released (mg/ml/min) = [protein content in 10% TCA at t min - protein content in 10% TCA at 0 min] / t

when t = time at the hydrolysate taken to be analyzed (min).

2.3.2 Protein content in protein hydrolysates

Fryer et al. (1986) Fryer, H. J., Davis, G. E., Manthorpe, M., & Varon, S. (1986). Lowry protein assay using an automatic microtiter plate spectrophotometer. Analytical Biochemistry , 153(2), 262-266. http://dx.doi.org/10.1016/0003-2697(86)90090-4. PMid:3706709.
http://dx.doi.org/10.1016/0003-2697(86)...
described a Lowry protein assay using an automatic microtiter plate spectrophotometer. One hundred microliters of protein hydrolysate were pipetted into each well of a flat bottom polystyrene 96-well plate (flat bottom polystyrene without lid and sterile (Corning, USA)) and 25 µl solution C (10% Na2CO3 in 0.5 N NaOH; 10% K, Na-tartrate; 5% Cu2SO4) was added, then mixed for 30s using plate-shaker and incubated for 10 min at room temperature. Ten µl of solution D (1 N, Folin-Ciocalteu phenol reagent) were added to each well and mixed for 30s. The reaction was incubated for 20 min at room temperature in the dark and then the absorbance was read at 750 nm by the multi-mode microplate reader. Bovine serum albumin was used to establish a standard curve.

2.3.3 Oxygen Radical Absorbance Capacity (ORAC) assay

The hydrophilic- and lipophilic-ORAC values of protein hydrolysates were analyzed according to the method of Huang et al. (2002) Huang, D., Ou, B., Hampsch-Woodill, M., Flanagan, J. A., & Prior, R. L. (2002). High-throughput assay of oxygen radical absorbance capacity using a multichannel liquid handing system coupled with a Microplate fluorescence reader in 96-well format. Journal of Agricultural and Food Chemistry, 50(16), 4437-4444. http://dx.doi.org/10.1021/jf0201529. PMid:12137457.
http://dx.doi.org/10.1021/jf0201529 ...
with some modifications. Freshly diluted fluorescein disodium (FL) in 75 mM phosphate buffer (pH 7.4) was prepared before adding to sample or standard. In the hydrophilic-ORAC, 75 mM phosphate buffer (pH 7.4) was used as a solubilizing medium. For the lipophilic-ORAC assay, sample and standard were diluted with the 7% randomly methylated β-cyclodextrin (RMCD) made in a 50 ml/100 ml acetone-water.

The 25 µl diluted samples or standards (trolox) were added with the FL working solution (120 µM, 175 µl) in each well of a 96-well black plate and pre-incubated at 37 °C for 15 min. Fifty µl for hydrophilic or 75 µl for lipophilic of 65.5 mM AAPH (2,2′-azobis (2-amidinopropane)) was added into each well using automatic dispenser to activate the reaction. The AAPH was freshly prepared in 75 mM phosphate buffer (pH 7.4) to generate free radicals for each run of assay. The fluorescence excitation/emission was performed at 485/530 nm. The FL intensity changes representing the peroxyl radicals inhibition was instantly recorded at 1 min interval for 35 min, at 37 °C in the multi-mode microplate reader. Phosphate buffer (75 mM, pH 7.4) was used as a negative control and subtracted from each peroxyl-treated sample (the net area under the fluorescence decay curve; net AUC).

The ORAC values (mg trolox equivalents per 100 µg BSA) were obtained from plotting a series of trolox concentrations (0-100 µg/ml; y) and the net AUC (x) using a linear regression equation (y = ax + b) ( Cao & Prior, 1999 Cao, G. H., & Prior, R. L. (1999). Measurement of oxygen radical absorbance capacity in biological samples. Oxidants and antioxidants. Methods in Enzymology , 299, 50-62. http://dx.doi.org/10.1016/S0076-6879(99)99008-0. PMid:9916196.
http://dx.doi.org/10.1016/S0076-6879(99...
).

2.3.4 ABTS radical cation decolorization (ABTS) assay

The ABTS radical scavenging activity of protein hydrolysates were estimated following the method of Arnao et al. (2001) Arnao, M., Cano, A., & Acosta, M. (2001). The hydrophilic and lipophilic contribution to total antioxidant activity. Food Chemistry, 73(2), 239-244. http://dx.doi.org/10.1016/S0308-8146(00)00324-1.
http://dx.doi.org/10.1016/S0308-8146(00...
. The ABTS radical cation (ABTS+•) was generated by the reaction between 7.4 mM ABTS solution and 2.6 mM potassium persulfate in the dark at room temperature for 12-16 h. The fresh ABTS+• working solution was prepared by diluting 1 ml of ABTS+• solution with 53 ml of phosphate buffer (10 mM, pH 7.4) containing 0.15 M NaCl to an absorbance reading of 1.1 ± 0.02 units (734 nm). The reaction between the protein hydrolysate or trolox as a standard (25 µl) and the diluted ABTS +• solution (200 µl) took place in the dark for 2 h. Reduction of blue-green ABTS radical colored solution was then recorded at 734 nm using the microplate reader. The ABTS radical scavenging activity was calculated from a standard curve of trolox ranging from 0 to 60 µg/ml and expressed as mg trolox equivalent per 100 µg BSA.

2.3.5 Assay of metal ion chelation

The method of Boyer & McCleary (1987) Boyer, R. F., & Mccleary, C. J. (1987). Superoxide ion as a primary reductant in ascorbate-mediated ferritin iron release. Free Radical Biology & Medicine, 3(6), 389-395. http://dx.doi.org/10.1016/0891-5849(87)90017-7. PMid:2828195.
http://dx.doi.org/10.1016/0891-5849(87)...
was modified to measure the metal chelating ability of protein hydrolysates. Five µl of 4 mM Fe2SO4 were added to 150 µl of sample solutions and blank (distilled water) in a microplate. Subsequently, the mixture was added with 10 µl of 10 mM ferrozine (3-(2-pyridyl)-5,6-bis(4-phenyl-sulfonic acid)-1,2,4-triazine) at room temperature for 30 min. The absorbance of the reaction was read at 562 nm using the microplate reader. EDTA was taken as the standard. The chelating activity was expressed as mg EDTA per 100 µg protein.

2.4 Statistical analysis

Analysis of variance (ANOVA) was carried out and mean comparison was determined by Duncan’s multiple range test (DMRT) at the 95% significant level (P<0.05) using SPSS statistical program, version 19. Two replications of experiment were performed and all analyses were carried out in (at least) triplicate.

To find out the percentage of the variance explaining the factors (protein types, substrate concentration and hydrolysis time) that influence the antioxidant activity of protein hydrolysates, principal component analysis (PCA) was performed by XLSTAT program, version 2017.

3 Results and discussion

3.1 Antioxidant properties of TB protein hydrolysates

The use of different hydrolysis conditions should release peptides with various abilities, chain sizes and chemical compositions. ANOVA test represents effects of factors and their interactions on protein content and antioxidant activities of hydrolysates ( Table 1 ). In general, all factors have significant effects on all responses. An interaction effect of protein type and substrate concentration was found in ABTS assay, metal chelating ability and release of TCA soluble peptides (P<0.05). The metal chelating ability and release of TCA soluble peptides were also influenced by interaction effect of protein type and hydrolysis time (P<0.05).

Table 1
ANOVA test representing effects of factors and their interactions on protein content and antioxidant activities of protein hydrolysates.

The high release of TCA soluble peptides showed a high release of peptide fragments. The protein content of APH was higher than other hydrolysates ( Table 2 ), which was agreed with its highest degree of hydrolysis considered as the release of peptides during hydrolysis ( Figure 1 ). The MP showed the lowest rate of hydrolysis possibly because it contained high NaCl content used in the extraction process resulting in ~1.5 M NaCl in the MP powder after freeze drying. It has been known that protein becomes aggregates at high salt concentration (salting out effect) ( Nelson & Cox, 2001 Nelson, D. L., & Cox, M. M. (2001). Lehninger Principles of Biochemistry (3rd ed.). New York: Worth Publisher Inc. http://dx.doi.org/10.1007/978-3-662-08289-8.
http://dx.doi.org/10.1007/978-3-662-082...
), thus being more difficult to be accessed by protease. All proteins were dramatically hydrolyzed during the first 15 min of incubation and being constant afterward. The lower hydrolysis degree of SPH was concomitant with the study by Toopcham et al. (2017) Toopcham, T., Mes, J. J., Wichers, H. J., Roytrakul, S., & Yongsawatdigul, J. (2017). Bioavailability of angiotensin I-converting enzyme (ACE) inhibitory peptides derived from Virgibacillus halodenitrificans SK1-3-7 proteinases hydrolyzed tilapia muscle proteins. Food Chemistry, 220, 190-197. http://dx.doi.org/10.1016/j.foodchem.2016.09.183. PMid:27855889.
http://dx.doi.org/10.1016/j.foodchem.20...
in tilapia muscle proteins. Their explanation was that sarcoplasmic proteins contained protease inhibitors with ability of serine protease inhibition. Thus, the Protease G6, an alkaline serine endopeptidase, used in our research might be partially inhibited during the hydrolysis of SP. The alkali-aided extracted proteins was experienced chemical unfolding, thus it might be more exposed for protease to attack. Wachirattanapongmetee et al. (2009) Wachirattanapongmetee, K., Thawornchinsombut, S., Pitirit, T., Yongsawatdigul, J., & Park, J. W. (2009). Functional properties of protein hydrolyses prepared from alkali-aided protein extraction of hybrid catfish frame. Trends in Food Science & Technology , 1, 71-81. reported the higher degree of hydrolysis of alkali-extracted proteins obtained from hybrid catfish frame compared to that of the control (minced hybrid catfish frame). These results demonstrated that different types of protein and preparation methods have different susceptibility to proteolysis.

Table 2
Effect of protein type on antioxidant activities and protein content of protein hydrolysates.
Figure 1
The release of soluble peptides of various protein types hydrolyzed by Protease G6 at selected substrate concentration (0.4 and 1.2%). CH, APH, MPH, and SPH; Hydrolysates of freeze dried minced fillet, alkali-aided extracted proteins, myofibrillar protein extract, and sarcoplasmic protein extract, respectively. Values expressed as mean of duplicates (each in triplicate) ± standard deviation.

Antioxidant activities of protein hydrolysates were influenced by the protein type used in hydrolysis (Table 2 ). The two ORAC assays used in this study allowed to monitor the antioxidant efficiency of both lipophilic and hydrophilic compounds in the protein hydrolysates independently by quenching the similar peroxyl free radical ( Wu et al., 2003 Wu, H. C., Chen, H. M., & Shiau, C. Y. (2003). Free amino acids and peptides as related to antioxidant properties in protein hydrolysates of mackerel (Scomber austriasicus ). Food Research International, 36(9-10), 949-957. http://dx.doi.org/10.1016/S0963-9969(03)00104-2.
http://dx.doi.org/10.1016/S0963-9969(03...
). The lipophilic-ORAC were typically lower in all samples compared to hydrophilic-ORAC ( Table 2 , 3 - 4 ) indicating the protein hydrolysates are more efficient in an aqueous system. Higher antioxidant potency regarding both hydrophilic and lipophilic ORAC assays of the control hydrolysates (CH) and AP hydrolysates (APH) was notified. In contrast, MP and SP hydrolysates (MPH & SPH) showed higher metal chelating ability than APH and CH (P<0.05). When comparing the antioxidant activities of all assays between MPH and SPH, they were comparable ( Table 2 ). Chen et al. (2015) Chen, X.X., Hu, X., Li, L.H., Yang, X.Q., Wu, Y.Y., Lin, W.L., Zhao, Y.Q., Ma, H.X. & Wei, Y. (2015). Antioxidant properties of tilapia component protein hydrolysates and the membrane ultrafiltration fractions. Advanced Materials Research, 1073-1076, 1812-1817. http://dx.doi.org/10.4028/www.scientific.net/amr.1073-1076.1812
http://dx.doi.org/10.4028/www.scientifi...
stated that SPH exhibited strongest antioxidant properties including DPPH, hydroxyl, superoxide anion radicals scavenging abilities and the total reducing power when compared with myofibrillar and stroma proteins from tilapia. Nevertheless, in this research, the fresh tilapia mince was used for the protein extractions unlike which, the freeze-dried tilapia byproducts were employed in our work. In addition, the protein hydrolysis was carried out from freeze-dried protein extracts whereas Chen and coworkers’ hydrolysates were prepared from protein extract solutions and different protease. Thus, with different process and enzyme used, different results were observed. Wang et al. (2013) Wang, H., Zhang, F., Cao, J., Zhang, Q., & Chen, Z. (2013). Proteolysis characteristics of sarcoplasmic, myofibrillar, and stromal proteins separated from grass carp and antioxidant properties of their hydrolysates. Food Science and Biotechnology, 22(2), 531-540. http://dx.doi.org/10.1007/s10068-013-0111-z.
http://dx.doi.org/10.1007/s10068-013-01...
reported the comparable OH scavenging activities of sarcoplasmic and myofibrillar protein hydrolysates from grass carp digested with Alcalase 2.4L. The discrepancy of these findings could be explained by the differences in substrate, protease type as well as enzyme-to-substrate ratio, all of which were directly affect the extent of hydrolysis and subsequently antioxidant activities. These results illustrated that each mechanism possibly relies on different molecular size and/or peptide composition and sequence to give the specific antioxidant activity ( Wu et al., 2003 Wu, H. C., Chen, H. M., & Shiau, C. Y. (2003). Free amino acids and peptides as related to antioxidant properties in protein hydrolysates of mackerel (Scomber austriasicus ). Food Research International, 36(9-10), 949-957. http://dx.doi.org/10.1016/S0963-9969(03)00104-2.
http://dx.doi.org/10.1016/S0963-9969(03...
; Cheung et al., 2012 Cheung, I. W. Y., Cheung, L. K. Y., Tan, N. Y., & Li-Chan, E. C. Y. (2012). The role of molecular size in antioxidant activity of peptide fractions from Pacific hake (Merluccius productus) hydrolysates. Food Chemistry, 134(3), 1297-1306. http://dx.doi.org/10.1016/j.foodchem.2012.02.215. PMid:25005946.
http://dx.doi.org/10.1016/j.foodchem.20...
).

Table 3
Effect of substrate concentration on antioxidant activities and protein content of protein hydrolysates.
Table 4
Effect of time of hydrolysis on antioxidant activities and protein content of protein hydrolysates.

Table 3 and 4 revealed effects of substrate concentration and hydrolysis time, respectively on antioxidant activities and protein content of protein hydrolysates. All responses increased with increasing substrate content and time of digestion. Contrarily, a decline trend of metal chelating ability was distinguished when hydrolysis time increased ( Table 4 ) implying that this property fades when the peptides size was decreased upon hydrolysis time. A similar trend found in protein hydrolysates from defatted salmon backbones ( Slizyte et al., 2016 Slizyte, R., Rommi, K., Mozuraityte, R., Eck, P., Five, K., & Rustad, T. (2016). Bioactivities of fish protein hydrolysates from defatted salmon backbones. Biotechnology Reports , 11, 99-109. http://dx.doi.org/10.1016/j.btre.2016.08.003. PMid:28352546.
http://dx.doi.org/10.1016/j.btre.2016.0...
). The akin correlations among those factors and responses were also observed in the principal component analysis (PCA) ( Figure 2 ). It can be concluded that the APH produced at 1.2% substrate concentration and 60 min hydrolysis possessed great antioxidant activities with the highest protein content.

Figure 2
Principal component analysis (PCA) on the measured parameters ABTS, ORAC and metal chelating ability of protein hydrolysates affected by protein type, substrate concentration and time of hydrolysis. Freeze dried minced fillet hydrolysate (CH) (solid circle), alkali-aided protein hydrolysate (APH) (open circle), myofibrillar protein hydrolysate (MPH) (solid square), and sarcoplasmic protein hydrolysate (SPH) (open square). Enlarged symbols represent hydrolysates of each protein obtained at 1.2 % substrate concentration and 60 min hydrolysis time. ABTS: ABTS radical scavenging activity; H-ORAC and L-ORAC: oxygen radical absorbance capacity in water and fat, respectively; metal: metal chelating ability.

3.2 Principal component analysis

PCA was performed to know how the production factors; protein type (4 levels), substrate concentration (5 levels) and hydrolysis time (6 levels) were correlated to antioxidant capacity of the protein hydrolysates. Dimensional distribution of all 120 treatments can be clustered according to their similarities and differences using PCA. Different antioxidant activities of samples were then grouped allowing understanding the predominant samples with high antioxidant potency.

Principal component (PC) plot ( Figure 2 ) comprises of two main axes explaining 84.1% (PC1 = 61.6 and PC2 = 22.5%) of the antioxidant capacity, protein content, and release of peptide fragments. It was noted that PC1 was associated positively with parameters which reflected hydrolysis of various tilapia proteins as protein content, release of peptides, as well as with antioxidant capacities (ABTS, hydrophilic-ORAC, and lipophilic-ORAC). While PC2 was associated positively with only metal chelating ability. Furthermore, it was evident that hydrophilic-ORAC, lipophilic-ORAC, and protein content were related with each other. The APH and CH especially at 1.2% substrate concentration and 60 min hydrolysis, which exhibited the highest antioxidant activities (ABTS and lipophilic-ORAC for APH and hydrophilic-ORAC for CH) closely located among most of attribute vectors in the biplot), indicating positive influence on the dependent components of PC1 i.e. hydrophilic-ORAC, lipophilic-ORAC, and protein content. Thus, their antioxidant activities were primarily caused by peptides with peroxyl radical quenching ability. Whereas, MPH and SPH were appeared near ABTS vector. Oppositely, less samples were located in the same quadrant (upper left) with metal chelating ability, which was isolated from other attribute vectors exhibiting that most protein hydrolysates were less associated with chelating mechanism. These results demonstrated that the APH exhibited comparative antioxidant abilities to the CH. The results were in consistent with the results obtained from ANOVA analysis.

4 Conclusions

Antioxidant activities of protein hydrolysates produced from tilapia byproducts significantly influenced by protein type, substrate concentration and hydrolysis time. The alkaline-aided protein hydrolysate (APH) demonstrated comparative antioxidant activity to the control (hydrolysate from tilapia fillet), which was more proficient than the myofibrillar and sarcoplasmic protein hydrolysates. The APH’s antioxidant activity was highly related to peptides with peroxyl radical quenching ability. Therefore, the potential of APH as the alternative natural antioxidant may be applied in food products. Nevertheless, identification of antioxidant peptides and their mechanisms are needed to be further investigated.

Acknowledgements

The authors would like to thank the Higher Education Research Promotion and National Research University Project of Thailand, Office of the Higher Education Commission, through the Food and Functional Food Research Cluster of Khon Kaen University (PhD-54115) for the scholarship and research fund. The authors thank Mr. Peter Bint for assistance with the English language presentation.

  • Practical Application: The alkaline-aided proteins were recovered from fish byproducts revealing antioxidant capacity after being hydrolysis under suitable conditions. The alkaline-aided protein hydrolysates (APH) hydrolyzed by Protease G6 can be used as a valuable and low-cost source for a natural antioxidant for human consumption.

References

  • Arnao, M., Cano, A., & Acosta, M. (2001). The hydrophilic and lipophilic contribution to total antioxidant activity. Food Chemistry, 73(2), 239-244. http://dx.doi.org/10.1016/S0308-8146(00)00324-1.
    » http://dx.doi.org/10.1016/S0308-8146(00)00324-1
  • Association of Official Analytical Chemists – AOAC. (1999). Official Methods of Analysis (16th ed.). Washington: Association of Official Analytical Chemist.
  • Boyer, R. F., & Mccleary, C. J. (1987). Superoxide ion as a primary reductant in ascorbate-mediated ferritin iron release. Free Radical Biology & Medicine, 3(6), 389-395. http://dx.doi.org/10.1016/0891-5849(87)90017-7. PMid:2828195.
    » http://dx.doi.org/10.1016/0891-5849(87)90017-7
  • Cao, G. H., & Prior, R. L. (1999). Measurement of oxygen radical absorbance capacity in biological samples. Oxidants and antioxidants. Methods in Enzymology , 299, 50-62. http://dx.doi.org/10.1016/S0076-6879(99)99008-0. PMid:9916196.
    » http://dx.doi.org/10.1016/S0076-6879(99)99008-0
  • Chen, X.X., Hu, X., Li, L.H., Yang, X.Q., Wu, Y.Y., Lin, W.L., Zhao, Y.Q., Ma, H.X. & Wei, Y. (2015). Antioxidant properties of tilapia component protein hydrolysates and the membrane ultrafiltration fractions. Advanced Materials Research, 1073-1076, 1812-1817. http://dx.doi.org/10.4028/www.scientific.net/amr.1073-1076.1812
    » http://dx.doi.org/10.4028/www.scientific.net/amr.1073-1076.1812
  • Cheung, I. W. Y., Cheung, L. K. Y., Tan, N. Y., & Li-Chan, E. C. Y. (2012). The role of molecular size in antioxidant activity of peptide fractions from Pacific hake (Merluccius productus) hydrolysates. Food Chemistry, 134(3), 1297-1306. http://dx.doi.org/10.1016/j.foodchem.2012.02.215. PMid:25005946.
    » http://dx.doi.org/10.1016/j.foodchem.2012.02.215
  • Chi, C. F., Hu, F. Y., Wang, B., Li, Z. R., & Luo, H. Y. (2015). Influence of amino acid compositions and peptide profiles on antioxidant capacities of two protein hydrolysates from Skipjack tuna (Katsuwonus pelamis) dark muscle. Marine Drugs , 13(5), 2580-2601. http://dx.doi.org/10.3390/md13052580. PMid:25923316.
    » http://dx.doi.org/10.3390/md13052580
  • Dekkers, E., Raghavan, S., Kristinsson, H. G., & Marshall, M. R. (2011). Oxidative stability of mahi mahi red muscle dipped in tilapia protein hydrolysats. Food Chemistry , 124(2), 640-645. http://dx.doi.org/10.1016/j.foodchem.2010.06.088.
    » http://dx.doi.org/10.1016/j.foodchem.2010.06.088
  • Dyer, W. J., French, H. V., & Snow, J. M. (1950). Proteins in fish muscle. I. Extraction of protein fraction in fresh fish. Journal of the Fisheries Research Board of Canada , 7(10), 585-593. http://dx.doi.org/10.1139/f47-052.
    » http://dx.doi.org/10.1139/f47-052
  • Fryer, H. J., Davis, G. E., Manthorpe, M., & Varon, S. (1986). Lowry protein assay using an automatic microtiter plate spectrophotometer. Analytical Biochemistry , 153(2), 262-266. http://dx.doi.org/10.1016/0003-2697(86)90090-4. PMid:3706709.
    » http://dx.doi.org/10.1016/0003-2697(86)90090-4
  • García-Moreno, P. J., Batista, I., Pires, C., Bandarra, N. M., Espejo-Carpio, F. J., Guadix, A., & Guadix, E. M. (2014). Antioxidant activity of protein hydrolysates obtained from discarded Mediterranean fish species. Food Research International , 65, 469-476. http://dx.doi.org/10.1016/j.foodres.2014.03.061.
    » http://dx.doi.org/10.1016/j.foodres.2014.03.061
  • Hoyle, N. T., & Merritt, J. H. (1994). Quality of fish protein hydrolysates from herring (Clupea harengus). Journal of Food Science, 59(1), 76-79. http://dx.doi.org/10.1111/j.1365-2621.1994.tb06901.x.
    » http://dx.doi.org/10.1111/j.1365-2621.1994.tb06901.x
  • Hraš, A. R., Hadolin, M., Knez, Ž., & Bauman, D. (2000). Comparison of antioxidative and synergistic effects of rosemary extract with α-tocopherol, ascorbyl palmitate and citric acid in sunflower oil. Food Chemistry, 71(2), 229-233. http://dx.doi.org/10.1016/S0308-8146(00)00161-8.
    » http://dx.doi.org/10.1016/S0308-8146(00)00161-8
  • Huang, D., Ou, B., Hampsch-Woodill, M., Flanagan, J. A., & Prior, R. L. (2002). High-throughput assay of oxygen radical absorbance capacity using a multichannel liquid handing system coupled with a Microplate fluorescence reader in 96-well format. Journal of Agricultural and Food Chemistry, 50(16), 4437-4444. http://dx.doi.org/10.1021/jf0201529. PMid:12137457.
    » http://dx.doi.org/10.1021/jf0201529
  • Hultin, H. O., & Kelleher, S. D. (1999). U.S. Pat. No. 6,005,073. Process for isolating a protein composition from a muscle source and protein composition. Advanced Protein Technologies Inc.
  • Nelson, D. L., & Cox, M. M. (2001). Lehninger Principles of Biochemistry (3rd ed.). New York: Worth Publisher Inc. http://dx.doi.org/10.1007/978-3-662-08289-8.
    » http://dx.doi.org/10.1007/978-3-662-08289-8
  • Slizyte, R., Rommi, K., Mozuraityte, R., Eck, P., Five, K., & Rustad, T. (2016). Bioactivities of fish protein hydrolysates from defatted salmon backbones. Biotechnology Reports , 11, 99-109. http://dx.doi.org/10.1016/j.btre.2016.08.003. PMid:28352546.
    » http://dx.doi.org/10.1016/j.btre.2016.08.003
  • Subdivision of Research and Analysis of Fisheries Statistics (2017). Nile tilapia production statistics Thailand: Fisheries Development Policy and Strategy Division, Department of Fisheries, Ministry of Agriculture and Cooperatives. Retrieved from https://www.fisheries.go.th/strategy/UserFiles/files/tilapia%206-60(1).pdf.
    » https://www.fisheries.go.th/strategy/UserFiles/files/tilapia%206-60(1).pdf
  • Toopcham, T., Mes, J. J., Wichers, H. J., Roytrakul, S., & Yongsawatdigul, J. (2017). Bioavailability of angiotensin I-converting enzyme (ACE) inhibitory peptides derived from Virgibacillus halodenitrificans SK1-3-7 proteinases hydrolyzed tilapia muscle proteins. Food Chemistry, 220, 190-197. http://dx.doi.org/10.1016/j.foodchem.2016.09.183. PMid:27855889.
    » http://dx.doi.org/10.1016/j.foodchem.2016.09.183
  • Wachirattanapongmetee, K., Thawornchinsombut, S., Pitirit, T., Yongsawatdigul, J., & Park, J. W. (2009). Functional properties of protein hydrolyses prepared from alkali-aided protein extraction of hybrid catfish frame. Trends in Food Science & Technology , 1, 71-81.
  • Wang, H., Zhang, F., Cao, J., Zhang, Q., & Chen, Z. (2013). Proteolysis characteristics of sarcoplasmic, myofibrillar, and stromal proteins separated from grass carp and antioxidant properties of their hydrolysates. Food Science and Biotechnology, 22(2), 531-540. http://dx.doi.org/10.1007/s10068-013-0111-z.
    » http://dx.doi.org/10.1007/s10068-013-0111-z
  • Wu, H. C., Chen, H. M., & Shiau, C. Y. (2003). Free amino acids and peptides as related to antioxidant properties in protein hydrolysates of mackerel (Scomber austriasicus ). Food Research International, 36(9-10), 949-957. http://dx.doi.org/10.1016/S0963-9969(03)00104-2.
    » http://dx.doi.org/10.1016/S0963-9969(03)00104-2

Publication Dates

  • Publication in this collection
    13 Dec 2018
  • Date of issue
    2019

History

  • Received
    10 Jan 2018
  • Accepted
    02 Oct 2018
Sociedade Brasileira de Ciência e Tecnologia de Alimentos Av. Brasil, 2880, Caixa Postal 271, 13001-970 Campinas SP - Brazil, Tel.: +55 19 3241.5793, Tel./Fax.: +55 19 3241.0527 - Campinas - SP - Brazil
E-mail: revista@sbcta.org.br