Vous êtes sur la page 1sur 18

Received: 14 July 2019 Revised: 13 October 2019 Accepted: 17 December 2019

DOI: 10.1111/jfpe.13370

ORIGINALARTICLE

Optimization of the porcine liver enzymatic hydrolysis


conditions
1 1 1
José U. Maluf | Mônica L. Fiorese | Keiti L. Maestre |
1 2 3
Fernanda R. Dos Passos | Joana K. Finkler | Jessica F. Fleck | Carlos E. Borba1

1
Chemical Engineering Postgraduate
Program, State University of West Paraná, Abstract
UNIOESTE, Toledo, Paraná, Brazil Worldwide porcine slaughterhouses produce significant amounts of by-products, such as
2
Fishing Resources and Fishing
porcine liver, daily. Even though the liver presents a very restricted market demand
Engineering Postgraduate Program, State
University of West Paraná, UNIOESTE, when raw, this product is composed of an exceptional nutritional content, with emphasis
Toledo, Paraná, Brazil
on protein, being, therefore, characterized as a potential raw material to produce protein
3
State University of West Paraná,
UNIOESTE, Toledo, Paraná, Brazil hydrolysates. This article aimed to study the production of protein hydrolysate using the
commercial enzymes Alcalase 2.4L™ and Novo Pro-D™. For the development of this
Correspondence
Mônica L. Fiorese, Chemical Engineering research, a central composite rotatable design with 4 axial points was applied for each
Postgraduate Program, State University enzyme. The factors evaluated were protein/water and enzyme/substrate ratios. The
of West Paraná, UNIOESTE, Campus
Toledo, Faculdade St. 645, Jd. La Salle, best hydrolysis results indicate average percentages of hydrolysis degree of 27.5% for
85903-000 Toledo, Paraná, Brazil. Email: both enzymes. The presence of low-molecular-weight peptides was also evidenced by
mlfiorese@gmail.com
the electrophoresis technique. It is, therefore, con-cluded that there is positive viability
Funding information for the use of hydrolyzed porcine liver in differ-entiated food applications with an
BRF S.A.; Coordination of Improvement of
Higher Education Personnel (CAPES), Grant/ aggregate market value.
Award Number: 001; State of Parana Research
Foundation (Araucária Foundation), Grant/ Practical Applications
Award Number: 50027 At present, the worldwide growing dissemination of food scarcity has become a
stim-ulus for the food industry to rethink the efficient use of raw materials to avoid
waste and maximize the use of existing nutritional resources. A potential way of
exploiting the nutritional value of the protein by-products with no commercial
applications is their transformation into protein hydrolysates by means of enzymatic
processes. Pro-tein hydrolysates have been in focus in the scientific community
with industrial appli-cations as beneficial products to the ingesting body, for acting
as either nutraceuticals or functional bioactive foods. Commercialization of this type
of prod-uct has become a growing market niche. Thus, this research, which
presents an inves-tigation regarding the conditions of the porcine liver hydrolysis,
fits within current issues, with innovative potential of obtainment of bioactive
peptides from by-products.

1 | I N T RO DU CT I O N Drummond, 2017). Porcine liver is considered to be a by-product of


swine slaughter, which is rich in proteins and preeminent nutrients,
The animal protein industry generates large amounts of by-products, such as Vitamins B12 and A (Nollet & Toldra, 2011; Seong et al.,
ranging from 33 to 49% of the weight of the living animal, depending 2015; Shimizu et al., 2006). The porcine liver can be commercialized

on the slaughtered species (Mullen, Alvarez, Zeugolis, Neill, & raw; however, it has a small consumer market. According to

J Food Process Eng. 2020;e13370. wileyonlinelibrary.com/journal/jfpe © 2020 Wiley Periodicals, Inc. 1 of 12


https://doi.org/10.1111/jfpe.13370
2 of 12 MALUF ET AL.

Srebernich, Gonçalves, and Domene (2017)), the powdered porcine bioactivities, such as antioxidant, antimicrobial, antihypertensive,
liver stands out as an excellent fortifier to be used in mixtures for antithrombotic, immunomodulatory, lipid-lowering, and anticancer
food preparations, which can be destined, for instance, to school (Kannan, Hettiarachchy, Marshall, Raghavan, & Kristinsson, 2011)
meals in soups, creams, cooked meat, and especially beans. activities.
Among the technological alternatives for the meat processing Enzymatic hydrolysis is very efficient in terms of process-related
industry regarding the transformation of by-products are the biotech- costs, and the enzyme is one of the most important economic vari-
nological processes, specifically the enzymatic hydrolysis. The use ables in the production of protein hydrolysates. In general, proteases
of enzymatic hydrolysis has grown considerably in recent years in of microbial origin with high hydrolytic activity and substrate specific-
the industrial segment, not only in the meat processing industry, but ity, in other words, high activity/cost ratio (Aspevik, Egede-Nissen, &
also in the pharmaceutical, food, and beverage industries (Martínez- Oterhals, 2016) are employed in industrial processes.
Alva-rez, Chamorro, & Brenes, 2015; Mora, Reig, & Toldrá, 2014; Among the enzymes of microbial origin, the most widely used in the
Mullen et al., 2017; Toldrá, Mora, & Reig, 2016). production of protein hydrolysates, either commercially or for research
The attractiveness of enzymatic hydrolysis lies on the high effi- studies, is Alcalase 2.4L™, a bacterial endopeptidase devel-oped by the
ciency of enzymes to act as biocatalysts, allowing hydrolytic reac- company Novozymes, originated in the submerged fer-mentation
tions to occur under mild conditions, and consequently, with lower process with the microorganism Bacillus licheniformis (Alavi, Jamshidian,
energy requirements. Moreover, this process presents notable prod- & Rezaei, 2019; Bernardi et al., 2016; Corîci et al., 2011). This enzyme is
uct selectivity through simplified production routes, low generation of classified as a subtilisin, a serine protease that is known to operate
undesirable bitter-tasting peptides (Shen, Guo, Dai, & Zhang, 2012), under moderate conditions of alkalinity (6–10) (Roslan et al., 2014)
lower environmental and physiological toxicity, and conse-quently, compared to the neutral or acidic enzymes, and temperatures of 35–70
high quality and productivity (Chapman, Ismail, & Dinu, 2018; Choi, C (Schmidt & Salas-Mellado, 2009) with opti-mum temperature of 55 C
Han, & Kim, 2015; Fernandes, 2010; Kapoor, Rafiq, & Sharma, (Bhaskar, Benila, Radha, & Lalitha, 2008; See, Hoo, & Babji, 2011) and
2017; Madhavan, Sindhu, Binod, Sukumaran, & Pandey, 2017; 60 C (Amiza, Nurul Ashikin, & Faazaz, 2011; Normah, Jamilah, Saari, &
Roslan, Yunos, Abdullah, & Kamal, 2014; Sun, Zhang, Ang, & Zhao, Yaakob, 2005), acting in the peptide bonds contained in the hydrophobic
2018). residues. The prominence of Alcalase 2.4L™ is found to be one of the
Protein hydrolysates can be obtained from different animal sources highly efficient enzymes for the protein hydrolysis due to its ability to
and their by-products, such as fish viscera (Feltes et al., 2010; Roslan, catalyze protein hydrolysis reaction with high degree of hydrolysis (DH)
Mustapa Kamal, Md. Yunos, & Abdullah, 2015; Roslan et al., 2014; Silva, in a relatively short period (Adler-Nissen, 1986; Aspmo, Horn, & Eijsink,
Ribeiro, Silva, Cahú, & Bezerra, 2014); sardine viscera (Venturin et al., 2005). Another enzyme that has presented application potential is Novo
2017); blue whiting (Micromesistius poutassou) (Harnedy et al., 2018); Pro-D™ produced by genetically modified B. licheniformis in submerged
salmon backbone (Slizyte et al., 2016); bovine liver (Di Bernardini et al., fermentation. The Novo Pro-D™ is also a subtilisin, but classified as a
2011); porcine liver (Shimizu et al., 2006); swine tissues (Damgaard, serine-endoprotease, which acts by hydrolyzing the internal peptide
Lametsch, & Otte, 2015); swine blood (Chang, Wu, & Chiang, 2007); bonds (de Miranda 2012; Rojas, Siqueira, Miranda, Tardioli, & Giordano,
chicken blood (Zheng, Si, Ahmad, Li, & Zhang, 2018); swine plasma (Liu, 2014; Zhou et al., 2019).
Kong, Xiong, & Xia, 2010); cattle plasma (Bah, Carne, McConnell, Mros,
& Bekhit, 2016); shrimp viscera (Katsuwonus pelamis) (Klomklao & The amino acid composition and DH during hydrolysis depend
Benjakul, 2016); and boneless chicken carcasses (Oliveira, Franzen, on various factors such as enzyme type, enzyme specificity, sub-
Terra, & Kubota, 2015). For rep-resenting potentially inexpensive strate composition, and on the characteristics of the process, such
feedstocks, these by-products have become industrially relevant to the as protein/water and enzyme/protein ratios, pH, temperature, and
meat industry, which has come to invest in attractive technology time of reaction. Thus, this study aimed to broaden the state of the
solutions that are efficient and able to deliver products with high levels of art regarding the knowledge of the basic principles of the factors that
protein and essential amino acids (FAO, 2017; Mora et al., 2014; Toldrá influence the hydrolysis reaction, which are crucial to obtain
et al., 2016). adequate process control which represent the maximum efficiency
In the meat packing industry, the enzymatic hydrolysis is of the raw material conversion into the desired product. The bio-
employed for protein fragmentation, that is, a parent protein can be transformation of the porcine liver by-product in the production of
hydrolyzed to produce a variety of different peptides with diverse protein hydrolysates was performed by optimizing the hydrolysis
activities. The peptides generated may be bioactive or functional, conditions of protein/water (%) and enzyme/protein (%) using the
depending on the type of enzyme and the operating conditions experimental planning tool. The enzyme Alcalase 2.4L™ was
employed (Sánchez & Vázquez, 2017). Several studies (Chang et selected, as it is the most widely used in the industrial segment of
al., 2007; Guo, Kouzuma, & Yonekura, 2009; Jang, Jo, Kang, & Lee, protein hydrolysates due to its high hydrolytic efficiency and low
2008; Liu, Xing, Fu, Zhou, & Zhang, 2016; Martínez-Alvarez et al., cost, compared to the enzyme Novo-ProD™. Although the last still
2015; Mullen et al., 2017; Sánchez & Vázquez, 2017; Villamil, lacks literature results, it was reported by Da Rosa et al. (2018) as
Váquiro, & Solanilla, 2017) showed that the ingestion of peptides being similar to Alcalase 2.4L™, besides presenting high cleavage
can improve health and highlight some of the major possibility.
MALUF ET AL. 3 of 12

2 | MATERIALSANDMETHODS
2 2
Y = b0 + b1X1b2X2 + b11X1 + b11X2 + b12X1X2: ð1Þ
2.1 | Feedstock

The real values for the levels of each independent variable (X i)


The porcine liver was donated by the company Brasil Food (BRF-Food), were represented by the factors, defined from literature searches
Toledo Unit, Paraná. The sample collection took place after the slaughter, in regarding meat raw materials (Schmidt & Salas-Mellado, 2009), and
the stage of animal evisceration. The liver used in this study came from similar industrial processes from BRF-Foods.
animals of the WH species (genetics information owned by the company), The amount of water added to the hydrolysis reaction medium
slaughtered after 190 days by carbon dioxide injection, ensuring the prin- was calculated from Equations (2) and (3):
ciples of animal welfare. The by-product porcine liver was ground in a helical protein ðgÞ
Water addition ðgÞ = ðX1Þ ð%Þ 2
mill, packed in polyethylene bags and stored in freezer (−18 ± 1 C) until raw material
transported under refrigeration (2 ± 1 C) in thermal boxes to the Food Quality Þ
ð 100 %
Control Laboratory located at the West Paraná State Uni-versity (Campus protein g ð Þ
Toledo, Paraná, Brazil), for further use. − raw material ð%Þ ,
ð Þ

raw material ðgÞ moisture ð%Þ


Total water ðgÞ = Water addition ðgÞ + ,
100 ð%Þ
2.2 | Chemical composition of the porcine liver ð3Þ

The porcine liver was characterized regarding total protein (928.08)


(MAPA, 1999; AOAC, 2016), lipids (991.36) (MAPA, 1999; AOAC, where X1 = protein/water (%) (CCRD).
2016), mineral matter (920.153) (AOAC, 2016), and moisture The hydrolysis was performed with 750 g of porcine raw
(950.46) (AOAC, 2016). The analyses were performed in triplicate. material for 60 min, under 100 rpm, the same pH of the raw material,
Two enzymes were used in the hydrolysis process: Alcalase around 6.4, and temperatures of 60 C for Alcalase 2.4L™ and 64 C
2.4L™ (Novozymes Latino Americana Ltd, Paraná, Brazil), a serine for Novo Pro-D™. After the reaction time, the enzyme was
pro-tease of microbial origin (B. licheniformis) with endogenous deactivated by heating the solution in a water bath at 85 C for 15
activity; and Novo Pro-D™, an alkaline protease produced by min (García-Moreno et al., 2014; Roslan et al., 2015; Verma, Chatli,
genetically modi-fied B. licheniformis. Kumar, & Mehta, 2017). The samples were then lyophilized to be
employed in molecular weight distribution and quantification of
amino acid composition.
2.3 | Central composite rotatable design for The experimental data were evaluated by using the software
enzymatic hydrolysis of porcine liver Statistica™ version 10 (Statsoft, Inc.), according to the variance analy-
sis (ANOVA) and the response surface analysis. The statistical signifi-
A central composite rotatable design (CCRD) was used to evaluate cance of the model was evaluated by the F test at a 5% level.

the conditions protein/water (%) (X1) and enzyme/protein (%) (X2) in


the porcine liver enzymatic hydrolysis. The experimental conditions
employed for the enzymes Alcalase 2.4L™ and Novo Pro-D™ are 2.4 | Enzymatic kinetics of porcine liver
pres-ented in Table 1. hydrolysate
For the two-factor system, the polynomial equation (Equation 1)
represented by the degree of protein hydrolysis (dependent variable) Enzymatic kinetics was performed for the two enzymes in the condi-
as a function of the terms of linear and quadratic effects was used, tion of maximum response for protein/water (%) and enzyme/protein
where Y is the predicted response, b 0 is the intercept, b1 and b2 are (%). The samples were collected at 5, 10, 15, 20, 30, 45, 60, 120,
the linear terms, and b11 and b12 are the quadratic terms. Y1 is the 180, 240, 300, 360, 420, and 480 min and frozen for further quantifi-
DH for the enzyme Alcalase 2.4L™ and Y2 for Novo Pro-D™. cation of the DH (%).

coded levels of the central composite rotatable design for the enzymes Alcalase 2.4L™ and
Novo Pro-D™

T A B L E 1 Real values of the


independent variables at different
Independet variable symbol −1.41 −1 0 1 1.41
X
Protein/water % (p/p) (wt/wt) 1 8 10 15 20 22.1
Coded Enzyme/protein % (p/p) X2 0.3 0.5 1 1.5 1.7
level
Vari
(wt/wt)
able
a
Coded level: define the factor levels in an experimental design.
4 of 12 detected by spectrophotometry after the enzymatic hydro-lysis of the
sample, with pancreatin solution at 45 C for 24 hr, followed by
2.5 | Degree of hydrolysis (%) colorimetric reaction with diamino benzaldehyde acid in sulfuric acid
medium sheltered from light for 6 hr. Then, a solution

The DH was determined according to the method described by


Hoyle and Merrit (1994)) and Baek and Cadwallader (1995).
Samples of hydrolyzed porcine liver were thawed and centrifuged
(Excelsa Baby I Centrifuge) at 2,600 rpm. Then, 20% trichloroacetic
acid was added in the ratio of 1:1 and held for 30 min. Subsequently,
the samples were filtered using a vacuum pump and Whatman
.
no. 40 filter paper. Finally, the absorbance of the samples were read
in a spectrophotom-eter (Thermo Fisher Scientific, model Spectronic
200E) at a wave-length of 750 nm. The results were expressed in
milligrams of albumin, according to Equation (4).

DH % p=p = Soluble protein ðmgÞ*100ð %Þ :


ð Þ

ð ÞðÞ Total protein mg

2.6 | Molecular weight distribution

This procedure was performed according to Laemmli's methodology


(Laemmli, 1970), where the Tris–HCl–glycine system buffer varied
from 4 to 17% Tris–tricine gradient gel. The samples (5 μL) were dis-
solved in 2% sodium dodecyl sulfate (SDS), 10% glycerol, 0.01%
−1
brom-ophenol blue, 0.0625 mol L Tris–HCl pH 6.8, and 5% β-
mercaptoethanol and heated for 5 min. Electrophoresis was con-
ducted using Mini PROTEAN 3 Cell electrophoresis system (Bio-Rad
−1
Laboratories, Hercules, CA) at 200 V with buffer (3 g L Tris-base,
−1 −1
14 g L glycine, and 1 g L SDS). After electrophoresis, the gel
pro-teins were stained with Coomassie Blue R-250 dye and
decolorized in a 50% methanol and 10% acetic acid solution. The
lmw marker from GE lifesciences, with molecular weight range
between 14.4 and 97 kDa was used as a control.

2.7 | Amino acid composition

The amino acid compositions of the porcine liver and protein hydro-
lysates were quantified by acid hydrolysis with 6 N hydrochloric acid,
3% (wt/vol) phenol, at 110 C for 24 hr. After hydrolysis, deriva-
tization occurred with the solution (4:4:2) of 0.2 N sodium acetate
trihydrate, high-performance liquid chromatograph (HPLC)-grade
methanol and 99–100% triethylamine. Amino acid identification was
performed by HPLC reverse phase column, with injection of 30 μL,
−1
detection at 254 nm, pH 6.40, flow of 1 mL min , 58 C, Eluent A
with 0.14 N sodium acetate, 0.14 N acetonitrile, and 0.14 N
triethylamine buffer and eluent B acetonitrile 60% (vol/vol) according
to the methodology proposed by White, Hart, and Fry (1986) and
Hagen, Frost, and Augustin (1989). The tryptophan con-tent was
MALUF ET AL.

of sodium nitrite was added and after 30 min the transmittance was
read at 590 nm in a spectrophotometer with quartz cuvettes.

3 | RESULTSANDDISCUSSION

3.1 | Chemical composition of porcine liver

The chemical composition of the raw porcine liver is reported in


Table 2.
From the results presented in Table 2, it can be verified that the
values found agree with the study of (Honikel, 2011), in which the
author reported variations in the porcine liver composition for protein
(18.9–21.6%) and lipids (2.4–6.8%).
Seong et al. (2015) evaluated the aspects of the chemical and
nutritional composition of chicken by-products, such as liver, gizzard,
stomach, small intestine, cecum, and duodenum. The results indicated
that the approximate range (minimum and maximum) of these by-
products were moisture (76.68–83.23%), fat (0.81–4.53%), protein
(10.96–17.70%). Although liver and gizzard presented higher levels of
protein, these values are lower than those found in the present study.
It should be noted that factors such as feeding, climate, and
time of growth of the animal to be slaughtered influence the
constitution of the nutrients present in the organ, however, the
values found are similar to those of this research.
In this context, by comparing the protein content found in the
meat by-products with the values present in the animal muscle, for
example, bovine loin (21%), porcine loin (22.2%) (De Castro
Cardoso Pereira & Dos Reis Baltazar Vicente, 2013), porcine chops
(17.3%), and duck meat (19.3%) (Ockerman & Basu, 2004), it is
noted that these values are similar to those found in by-products of
animal origin, especially the liver.
Thus, the potential presented by raw porcine liver as a source of
protein of high biological value, able to meet daily nutrient require-
ments is verified, with the possibility of being used in hydrolytic
indus-trial processes to obtain new high-value products, because
biopeptides are obtained from enzymatic hydrolysis.

3.2 | Porcine liver enzymatic hydrolysis

The results of the variables used in the CCRD regarding the produc-
tion of porcine liver protein hydrolysates using the enzymes Alcalase

T A B L E 2 Chemical composition of the raw porcine liver

Composition Raw porcine livera


Ashes (%) 1.4

Lipids (%) 2.0


Total protein (%) 19.7

Moisture (%) 74.3


a
Analysis performed in triplicate, mean value expressed.
MALUF ET AL. 5 of 12

2.4L™ and Novo Pro-D™ are presented in Table 3. The results of the model can represent 95.44% of the response variability, demon-
the analysis of variance concerning the DH are reported in Table 4. strating that the regression model is significant for both enzymes.
Within the range studied, linear (X 1) and (X2) variables were sig- The equation obtained by the statistical analysis of the regression
nificant (p value <.05) for the two enzymes used. Results of the response model with the coded variables is presented in Equations (3) and
surface of the model (Figure 1a) for the enzyme Alcalase 2.4L™ were (4), respectively.
evaluated by the ANOVA and F test (Table 4). The F value calculated for
a 95% confidence interval (p value <.05) was Fc (2;5;0.05) = 11.09, which Y1 = 18:66 + 3:80ðX1Þ + 2:58ðX2Þ, ð3Þ

was greater than the tabulated value (F tab2;5;0.05 = 5.79). The


2 Y2 = 18:81 + 7:26ðX1Þ + 7:56ðX2Þ:
regressions obtained provided R of 0.8607, indicating that the model ð4Þ

can explain 86.07% of the response vari-ability (Y 1). While for the
enzyme Novo Pro-D™ (Figure 2a), the F value calculated for a 95% The DH (Table 3) ranged from 10.93 to 27.86% (wt/wt), with Assay 4
confidence interval (p value <.05) was Fc (2;5;0.05) = 18.54, which was as the best response for both enzymes (20% X 1 and 1.5% X2), providing
higher than the tabulated F (F tab 2;5;0.05 = 5.79) (Barros Neto, Bruns, & results of 27.86 and 27.30% (wt/wt) for Y 1 and Y2, respec-tively. Table 5
Scarminio, 2010). In addition, presents an overview of the hydrolytic ability of

T A B L E 3 Experimental design for the degree of hydrolysis variable response with enzyme Alcalase 2.4L™ (Y 1) and Novo Pro-D™ (Y2) in 1 hr

Coded level Real level Response

Run (X1) (X2) Protein/water (%) (wt/wt) Enzyme/protein (%) (wt/wt) Y1 Y2


1 −1 0.5 10 0.5 12.43 12.56

2 1 0.5 20 0.5 22.47 15.57


3 −1 1.5 10 1.5 17.05 17.53

4 1 1.5 20 1.5 27.85 27.30


5 −1.41 1 8 1 15.70 11.07

6 1.41 1 22 1 22.34 22.49


7 0 0.3 15 0.3 12.29 10.93

8 0 1.7 15 1.7 19.75 20.40


9 0 1 15 1 18.12 18.85

10 0 1 15 1 18.60 18.65
11 0 1 15 1 19.36 19.04

TABLE4 Analysis of variance for the degree of hydrolysis (Y1 and Y2)

Degrees of
Model freedom Sum-of-squares Mean squares F value p Value R2 Adjacent R2
Y1 Regression 3 118.41 39.48 0.8607 0.7214
Linear 1 114.67 114.67 293.32 .0034
Square 1 3.60 3.60 9.20 .0936
Interaction 1 0.14 0.14 0.36 .6052
Residual 5 37.75 7.55
Lack of fit 3 27.80 9.27 23.71 .0407
Total 7 205.19
Y2 Regression 3 117.36 39.12 0.9544 0.9088
Linear 1 104.50 104.50 2,747.50 .0004
Square 1 1.44 1.44 37.92 .0254
Interaction 1 11.42 11.42 300.38 .0033
Residual 5 2029.29
Lack of fit 3 11.19 3.73 98.10 .0101

Total 7 2,247.18
6 of 12 MALUF ET AL.

F I G U R E 1 Response surface (a) and Pareto chart (b) for the porcine liver hydrolysate for the enzyme Alcalase 2.4L™

F I G U R E 2 Response surface (a) and Pareto chart (b) for the porcine liver hydrolysate for the enzyme Novo Pro-D™

different enzymes in meat raw materials. Thus, these results demon- hand, the interaction between the variables had no significant effect
strate that the hydrolytic process of this study was good, as it within a 95% confidence interval (p value <.05). The greatest
resulted in a higher DH compared to the literature. It also showed influence on the enzymatic hydrolysis process of porcine liver was
that the vol-ume of water and enzyme added in the process was provided by the variable X1, presenting better contour for the Novo
low, which, from an industrial point of view, is an important factor, as Pro-D™ enzyme.
it does not entail considerable costs regarding the acquisition of According to the response surface plots, the linearity of the
enzyme and energy for the product drying phase, thus contributing model was verified, as the higher X1 and X2, the higher Y1 and Y2. If
to the economic viability of the process. X1 is 22 and X2 is 1.7, the DH will be approximately 25% for the
Using the software Statistica, the contour plot regarding the X 1 enzyme Alcalase 2.4L™, whereas for Novo Pro-D™, if X 1 is 22 and
and X2 effects for Y1 and Y2 (Figures 1 and 2b) showed a positive X2 is 1.7, the DH is 27%. Thus, the directly proportional behavior
effect for Y1 and Y2, with X1 as the most significant. On the other between Y1 and Y2 reinforce the linearity of the model.
MALUF ET AL. 7 of 12

T A B L E 5 Overview of the hydrolytic ability of different enzymes and raw materials

Temperature Degree of
Raw material Enzyme pH ( C) T (hr) hydrolysis (%) References
Tilapia frame protein isolate Pepsin 2 37 4 23.46 Lin, Alashi, Aluko, Sun Pan, and Chang
(2017)

Tilapia Alcalase 2.4L™ 7.5 60 2 20.2 Roslan et al. (2015)


Sardine Alcalase 2.4L™ 8 50 2 13.7–14.9 García-Moreno et al. (2014)
Blue whiting (Micromesistius Alcalase 2.4L™ 7 50 4 29.13–32.58 Harnedy et al. (2018)
poutassou)
Porcine liver Papain 6.5 50 6 1.18–19.12 Verma et al. (2017)

Porcine liver Alcalase 2.4 L™ 8.0 50 6 1.22–23.56 Verma et al. (2017)


Porcine liver Alcalase 2.4 L™ 6.4 60 1 12.29–27.85 In this work

Porcine liver Novo Pro-D™ 6.4 64 1 10.93–27.30 In this work

F I G U R E 3 Maximum response of protein hydrolysis for the enzyme Alcalase 2.4L™ where (a) X 1 (%) constant and X2 (%) varying
and (b) X2 (%) constant and X1 (%) varying

enzyme/protein ratio was fundamental to obtain DH and conse-


quently the protein content
3.3 | Study in the regions of maximum protein Regarding the enzyme Alcalase 2.4L™, it was observed in Figure
hydrolysis response 3a that when the protein/water ratio is 20% (wt/wt), there is an increase

It was not possible to obtain an optimized condition by the CCRD,


but rather an indication of the best relations with the maximum
response, thus, the best conditions for the hydrolytic process
response of the DH was determined for each enzyme. These assays
were performed by fixing the protein/water ratio factor (20, 22, 24
and 26% [wt/wt]) and varying the enzyme/protein ratio factor (0.5,
1.0, 1.5, 2.0, 3.0, and 6.0). The tests regarding Alcalase 2.4L™
ranged from 23 to 46 and Novo Pro-D™ from 47 to 70.
Figure 3a shows that the enzyme/protein ratio of 1.5% (wt/wt)
was the one with the highest DH for Alcalase 2.4L™ and adopted as
reference for this analysis. Therefore, it is verified that the variable
in DH, which was adopted as reference for the kinetics. However, for
the variable enzyme/protein ratio, the best DH was 6.0% while the
vari-ation of the DH was not significant when compared to an
enzyme/pro-tein ratio of 1.5%, which justifies the choice of the
enzyme/protein ratio of 1.5% for the kinetic experiment, because the
enzyme repre-sents a material of high value for the process.
According to Sadana (1991), an increase in the velocity
decreases the size of the droplets formed, thereby increasing the
interfacial area, which leads to an extension of the hydrolysis. For
protein/water ratios higher than 24%, Figure 3b shows that for
Alcalase 2.4L™, the rheol-ogy of the product presented difficulties in
maintaining a homoge-neous velocity of the sample movement,
generating regions with lower speeds, which lead to less favorable
conditions for a hydrolyzed product uniformity.
Figure 4a shows that for the enzyme Novo Pro-D™, the protein/
water ratio of 22% produced the highest DH, while Figure 4b shows
that an enzyme/protein ratio of 3.0% provided the best response.
Thus, this condition was adopted for the kinetics study
8 of 12 MALUF ET AL.

F I G U R E 4 Maximum response of protein hydrolysis for the enzyme Novo Pro-D™ where (a) X1 (%) constant and X2 (%) varying and (b)
X2 (%) constant and X1 (%) varying

F I G U R E 5 Kinetics of the enzymatic hydrolysis varying X1 (%) for both enzymes, (a) Alcalase 2.4L™ (b) Novo Pro-D™

3.4 | Kinetics of the enzymatic hydrolysis for the enzyme would be more favorable to be applied in industrial
enzymes processes.

The kinetics was developed under the condition of maximum


response of protein hydrolysis. Figure 5a,b shows the kinetic curves 3.5 | Molecular weight distribution
for the hydrolysis condition X1 at 20% (wt/wt) and X2 at 1.5% (wt/wt)
at 62 C and pH of 6.4. It is observed that in the hydrolytic process Figure 6 shows the protein profile of the porcine liver both raw and
(Figure 5a,b), the reactions occurred with a high rate up to 60 min. hydrolyzed for the best condition of the CCRD after 1 and 4 hr for both
After that, there is a reduction in the reaction speed, reaching the enzymes. It was verified that there was hydrolysis in all the sam-ples
stationary phase in 550 min (Alcalase 2.4L™) and considering the formation of peptides with low molecular weight (14
300 min (Novo Pro-D™) with DH of 34 and 31%, respectively. How- kDa), therefore the cleavage of the peptide bonds happened. The
ever, a 9-hr (550 min) hydrolysis time would be economically unfeasible distribution of the peptides was similar and compatible with the DH
to the industry due to the process time as well as changes that could observed for the different enzymes. The formation of bioactive pep-tides,
arise in the hydrolysate, such as degradation of bioactive principles and which occur at low molecular weight, is likely to happen. Studies have
bitterness (Shen et al., 2012). Thus, the Novo Pro-D™ shown that the enzyme Alcalase 2.4L™ can produce low
MALUF ET AL. 9 of 12

ence of essential amino acids both balanced and similar to those of


muscle proteins. When comparing amino acid levels with those of

F I G U R E 6 Molecular weight distribution. Lane 1, control; Lane


2, Novo Pro-D™ in 1 hr; Lane 3, Novo Pro-D™ in 4 hr; Lane 4,
Alcalase 2.4L™ in 1 hr; and Lane 5, Alcalase 2.4L™ in 4 hr

molecular weight peptides when there is a high DH (Lalasidis, Boström,


& Sjöberg, 1978; Liaset, Lied, & Espe, 2000; Roslan et al., 2015). This
behavior was also observed for the enzyme Novo Pro-D™.

3.6 | Amino acid composition

Amino acid compositions are shown in Table 6. It is verified that


both porcine liver and protein hydrolysates (Y1 and Y2) showed high
levels of glutamic acid and aspartic acid. In general, there was no
significant difference for cysteine between the raw material and the
hydroly-sates. Protein hydrolysate with the Novo Pro-D™ enzyme
had higher tryptophan content than porcine liver and protein
hydrolysate with the enzyme Alcalase 2.4L™.
Protein hydrolysates showed a higher content for the amino
acids glutamic acid, aspartic acid, leucine, and lysine for both
enzymes, reaching values of 1.98, 1.51, 1.41, and 1.27%,
respectively, for Alcalase 2.4L™ and 2.18, 1.67, 1.53, and 1.37%,
respectively, for Novo Pro-D™.
All essential and nonessential amino acids were detected in bal-
anced amounts for both porcine liver and protein hydrolysate (Y1
and Y2), with percentages of essential amino acids of 50.22, 51.05,
and 50.33%, respectively. (Anderson, 1988), in his characterization
study of porcine by-products, found varied levels of essential amino
acids, but similar to those of the muscular tissues of the porcine.
Also, Aristoy (2011) found, in by-products of animal origin, the pres-
D™, hydrolysis rates were similar and relatively high. However, from
T A B L E 6 Total amino acid composition (%) of raw porcine liver the response surface and Pareto chart, it was found that, within the
and protein hydrolysate in 1 hr
conditions evaluated, the protein/water factor was more significative
Protein for the enzyme Alcalase 2.4L™, while for Novo Pro-D™, the
hydrolysate enzyme/protein ratio was the most significative fac-tor. Regarding
Raw porcine industrial application, the best condition for the porcine liver would
Total amino acid liver Y1 Y2 be a 1.5% (wt/wt) enzyme/protein and 20% (wt/wt) protein/water

Essential ratio, 62 C, pH 6.4, 5 hr, and enzyme Novo

Histidine 0.67 0.53 0.58


Arginine 0.97 0.70 0.77

Threonine 0.84 0.65 0.73


Valine 1.20 0.95 1.03

Methionine 0.44 0.35 0.38


Isoleucine 0.91 0.69 0.76

Leucine 1.82 1.41 1.53


Phenylalanine 0.98 0.76 0.82

Lysine 1.66 1.27 1.37


Tryptophan 0.30 0.25 0.39

Nonessential
Aspartic acid 2.10 1.51 1.67

Glutamic acid 2.57 1.98 2.18


Serina 0.83 0.63 0.75

Glycine 1.08 0.86 0.98


Alanine 1.17 0.92 0.99

Proline 0.92 0.66 0.81


Tyrosine 0.58 0.38 0.56

Cysteine 0.34 0.31 0.31


Total essential amino acid 9.79 7.56 8.36
content

Total nonessential amino acid 9.59 7.25 8.25


content
Total amino acid content 19.88 14.81 16.61

Percentage of essential amino 50.52 51.05 50.33


acids

FAO, it is found that it meets the demand of the daily recommenda-


tions established by FAO (2017).

4 | CONCLUSIONS

The use of the response surface methodology approach made it


possible to identify the factors relevant to the hydrolytic process. For
both enzymes used in the processes, Alcalase 2.4L™ and Novo Pro-
10 of 12 MALUF ET AL.

Pro-D™ The enzyme Novo Pro-D™ showed to be more efficient, Bhaskar, N., Benila, T., Radha, C., & Lalitha, R. G. (2008). Optimization
presenting a higher DH (28%), in less process time when compared of enzymatic hydrolysis of visceral waste proteins of Catla (Catla
catla) for preparing protein hydrolysate using a commercial protease.
to the commercially used for industrial processes, Alcalase 2.4L™.
Bio-resource Technology, 99(2), 335–343. https://doi.org/10.1016/j.
When it comes to molecular weight distribution, the values obtained biortech.2006.12.015
were low for both enzymes (<14 kDa), which may repre-sent the Chang, C. Y., Wu, K. C., & Chiang, S. H. (2007). Antioxidant properties
formation of bioactive compounds that show beneficial physiologic and protein compositions of porcine haemoglobin hydrolysates. Food
Chemistry, 100(4), 1537–1543. https://doi.org/10.1016/j.foodchem.
effects for the living organisms. Thus, we conclude that the porcine
2005.12.019
liver can be employed in the development of new products. Chapman, J., Ismail, A., & Dinu, C. (2018). Industrial applications of
enzymes: Recent advances, techniques, and outlooks. Catalysts,
8(6), 238. https://doi.org/10.3390/catal8060238
Choi, J.-M., Han, S.-S., & Kim, H.-S. (2015). Industrial applications of enzyme
ACKNOWLEDGMENTS
biocatalysis: Current status and future aspects. Biotechnology Advances,
The authors acknowledge the financial support from the company 33(7), 1443–1454. https://doi.org/10.1016/j.biotechadv.2015.02.014
Brasil Food (BRF-Food), Toledo Unit, Paraná. And the Coordination Corîci, L. N., Frissen, A. E., Van Zoelen, D. J., Eggen, I. F., Peter, F.,
of Improvement of Higher Level Personnel, Brazil (Capes), Finance Davidescu, C. M., & Boeriu, C. G. (2011). Sol-gel immobilization of
Alcalase from Bacillus licheniformis for application in the synthesis of C-
Code 001, Araucária Foundation of Support to the Scientific and
terminal peptide amides. Journal of Molecular Catalysis B: Enzymatic,
Technological Development of the State of Paraná. 73(1–4), 90–97. https://doi.org/10.1016/j.molcatb.2011.08.004
Damgaard, T., Lametsch, R., & Otte, J. (2015). Antioxidant capacity of
ORCID hydrolyzed animal by-products and relation to amino acid composition
and peptide size distribution. Journal of Food Science and Technology,
Mônica L. Fiorese https://orcid.org/0000-0001-5250-7178
52(10), 6511–6519. https://doi.org/10.1007/s13197-015-1745-z
Da Rosa, L. D. L., Santana, M. C., Avezedd, T. L., Brígoda, A. O. S.,
RE FE R ENC E S Gdddy, R., Pachecd, S., … Cabral, L. M. C. (2018). A comparison of
Adler-Nissen, J. (1986). Enzymatic hydrolysis of food proteins. New dual-functional whey hydrolysates by the use of commercial prote-
York, NY: Elsevier Science Publishing Co. ases. Food Science and Technology, 38, 31–36. https://doi.org/10.
Alavi, F., Jamshidian, M., & Rezaei, K. (2019). Applying native proteases 1590/fst.08417
from melon to hydrolyze kilka fish proteins (Clupeonella cultriventris De Castro Cardoso Pereira, P. M., & Dos Reis Baltazar Vicente, A. F.
caspia) compared to commercial enzyme Alcalase. Food Chemistry, (2013). Meat nutritional composition and nutritive role in the human
277, 314–322. https://doi.org/10.1016/j.foodchem.2018.10.122 diet. Meat Science, 93, 586–592. https://doi.org/10.1016/j.meatsci.
Amiza, M. A., Nurul Ashikin, S., & Faazaz, A. L. (2011). Optimization of 2012.09.018
enzymatic protein hydrolysis from silver catfish (Pangasius sp.) de Miranda, L. C. (2012). Obtaining and characterization of enzymatic
frame. International Food Research Journal, 18(2), 775–781. hydrolysates of soybean mealproteins. Retrieved from: https://
Anderson, B. A. (1988). Edible meat by-products, advance in meat research. repositorio.ufscar.br/handle/ufscar/4100
London: Elsevier Applied Science Publishing Ltd. de Oliveira, M. S. R., Franzen, F. d. L., Terra, N. N., & Kubota, E. H.
AOAC. (2016). Chapter 39, Official Methods of Analysis 20th. Rockville, (2015). Utilizaç~ao de enzimas proteolíticas para produç~ao de
Maryland: United States of America. hidrolisados proteicos a partir de carcaças de frango desossadas
Aspevik, T., Egede-Nissen, H., & Oterhals, Å. (2016). A systematic manualmente. Bra-zilian Journal of Food Technology, 18(3), 199–
approach to the comparison of cost efficiency of endopeptidases for 210. https://doi.org/10. 1590/1981-6723.4414
the hydrolysis of Atlantic salmon (Salmo salar) by-products. Food Di Bernardini, R., Harnedy, P., Bolton, D., Kerry, J., O’Neill, E., Mullen, A. M.,
Tech-nology and Biotechnology, 54(4), 421–431. & Hayes, M. (2011). Antioxidant and antimicrobial peptidic hydro-lysates from
https://doi.org/10.17113/ ftb.54.04.16.4553 muscle protein sources and by-products. Food Chemistry, 124(4), 1296–
Aspmo, S. I., Horn, S. J., & Eijsink, V. G. H. (2005). Enzymatic hydrolysis 1307. https://doi.org/10.1016/j.foodchem.2010.07.004 FAO. (2017). Protein
of Atlantic cod (Gadus morhua L.) viscera. Process Biochemistry, and amino acids requiremens in human nutrition. Roma,
40(5), 1957–1966. https://doi.org/10.1016/j.procbio.2004.07.011 Itália: Report a Joint.
Aristoy, M.-C. (2011). Essential Amino Acids. Handbook of Analysis of Feltes, M. M. C., Correia, J. F. G., Beir~ao, L. H., Block, J. M., Ninow, J.
Edible Animal By-Products, 123–135. https://doi.org/10.1201/ L., & Spiller, V. R. (2010). Alternativas para a agregaç~ao de valor
b10785-12 aos resíduos da industrializaç~ao de peixe. Revista Brasileira de
Baek, H. H., & Cadwallader, K. R. (1995). Enzymatic hydrolysis of Engenharia Agrícola e Ambiental, 14(6), 669–677.
crayfish processing by-products. Journal of Food Science, 60(5), https://doi.org/10.1590/s1415-43662010000600014
929–935. https://doi.org/10.1111/j.1365-2621.1995.tb06264.x Fernandes, P. (2010). Enzymes in food processing: A condensed
Bah, C. S. F., Carne, A., McConnell, M. A., Mros, S., & Bekhit, A. E.-D. overview on strategies for better biocatalysts. Enzyme Research,
A. (2016). Production of bioactive peptide hydrolysates from deer, 2010(May), 1–19. https://doi.org/10.4061/2010/862537
sheep, pig and cattle red blood cell fractions using plant and fungal García-Moreno, P. J., Batista, I., Pires, C., Bandarra, N. M., Espejo-
protease preparations. Food Chemistry, 202, 458–466. Carpio, F. J., Guadix, A., & Guadix, E. M. (2014). Antioxidant activity
https://doi.org/10.1016/ j.foodchem.2016.02.020 of protein hydrolysates obtained from discarded Mediterranean fish
Barros Neto, B., Bruns, R. R., & Scarminio, I. S. (2010). How to Experiment - spe-cies. Food Research International, 65(PC), 469–476.
Science and Industry Applications. S~ao Paulo: Bookman. https://doi.org/10. 1016/j.foodres.2014.03.061
Bernardi, D. M., de Paris, L. D., Dieterich, F., Silva, F. G. D. e., Boscolo, W. Guo, H., Kouzuma, Y., & Yonekura, M. (2009). Structures and properties of
R., Sary, C., … Sgarbieri, V. C. (2016). Production of hydro-lysate from antioxidative peptides derived from royal jelly protein. Food Chemistry,
processed Nile tilapia (Oreochromis niloticus) residues and assessment of 113(1), 238–245. https://doi.org/10.1016/j.foodchem.2008.06.081
its antioxidant activity. Food Science and Technology, 36 Hagen, S. R., Frost, B., & Augustin, J. (1989). Precolumn
(4), 709–716. https://doi.org/10.1590/1678-457x.15216 phenylisothiocyanate derivatization and liquid chromatography of
MALUF ET AL. 11 of 12

Martínez-Alvarez, O., Chamorro, S., & Brenes, A. (2015). Protein


amino acids in food. Journal of the Association of Official Analytical hydroly-sates from animal processing by-products as a source of
Chemists, 72(6), 912–916. bioactive mol-ecules with interest in animal feeding: A review. Food
Harnedy, P. A., Parthsarathy, V., McLaughlin, C. M., O'Keeffe, M. B., Research
Allsopp, P. J., McSorley, E. M., … FitzGerald, R. J. (2018). Blue
whiting (Micromesistius poutassou) muscle protein hydrolysate with
in vitro and in vivo antidiabetic properties. Journal of Functional
Foods, 40, 137–145. https://doi.org/10.1016/j.jff.2017.10.045
Hoyle, N. T., & Merrit, J. H. (1994). Quality of fish protein hydrolysates
from herring (Clupea harengus). Journal of Food Science, 59(1), 76–
79. https://doi.org/10.1111/j.1365-2621.1994.tb06901.x
Jang, A., Jo, C., Kang, K., & Lee, M. (2008). Food chemistry
antimicrobial and human cancer cell cytotoxic effect of synthetic
angiotensin-converting enzyme (ACE) inhibitory. Peptides, 107, 327–
336. https:// doi.org/10.1016/j.foodchem.2007.08.036
Honikel, K.-O. (2011). Composition and Calories. Handbook of Analysis
of Edible Animal By-Products, 105–121. https://doi.org/10.1201/
b10785-11
Kannan, A., Hettiarachchy, N. S., Marshall, M., Raghavan, S., &
Kristinsson, H. (2011). Shrimp shell peptide hydrolysates inhibit
human cancer cell proliferation. Journal of the Science of Food and
Agriculture, 91(10), 1920–1924. https://doi.org/10.1002/jsfa.4464
Kapoor, S., Rafiq, A., & Sharma, S. (2017). Protein engineering and its applica-tions
in food industry. Critical Reviews in Food Science and Nutrition, 57 (11), 2321–
2329. https://doi.org/10.1080/10408398.2014.1000481
Klomklao, S., & Benjakul, S. (2016). Utilization of Tuna Processing
Byproducts: Protein Hydrolysate from Skipjack Tuna (Katsuwonus
pelamis) Viscera. Journal of Food Processing and Preservation,
41(3), 1–8. https://doi.org/10.1111/jfpp.12970
Laemmli, U. (1970). Cleavage os structural proteins during the assembly
of the headbacteriophage T4. Nature, 680–685. https://doi.org/10.
1038/227680a0
Lalasidis, G., Boström, S., & Sjöberg, L. B. (1978). Low molecular weight
enzymatic fish protein hydrolysates: Chemical composition and nutri-
tive value. Journal of Agricultural and Food Chemistry, 26(3), 751–
756. https://doi.org/10.1021/jf60217a045
Liaset, B., Lied, E., & Espe, M. (2000). Enzymatic hydrolysis of by-products
from the fish-filleting industry; chemical characterisation and nutri-tional
evaluation. Journal of the Science of Food and Agriculture, 80(5),
581–589. https://doi.org/10.1002/(SICI)1097-0010(200004)80:
5<581::AID-JSFA578>3.0.CO;2-I
Lin, H.-C., Alashi, A. M., Aluko, R. E., Sun Pan, B., & Chang, Y.-W. (2017).
Antihypertensive properties of tilapia ( Oreochromis spp.) frame and skin
enzymatic protein hydrolysates. Food & Nutrition Research, 61(1),
1391666. https://doi.org/10.1080/16546628.2017.1391666
Liu, Q., Kong, B., Xiong, Y. L., & Xia, X. (2010). Antioxidant activity and
functional properties of porcine plasma protein hydrolysate as
influenced by the degree of hydrolysis. Food Chemistry, 118(2), 403–
410. https://doi.org/10.1016/j.foodchem.2009.05.013
Liu, R., Xing, L., Fu, Q., Zhou, G., & Zhang, W. (2016). A review of
antioxi-dant peptides derived from meat muscle and by-products.
Antioxi-dants, 5(3), 32. https://doi.org/10.3390/antiox5030032
Madhavan, A., Sindhu, R., Binod, P., Sukumaran, R. K., & Pandey, A.
(2017). Strategies for design of improved biocatalysts for industrial
applica-tions. Bioresource Technology, 245, 1304–1313.
https://doi.org/10. 1016/j.biortech.2017.05.031
MAPA (21 de june de 1999). Chapter V - Métodos Quantitativos,
Instruç~ao Normativa n 20. Brasília, Brazil.
Mari Silvia Rodrigues de, O., Felipe de Lima, F., Nelcindo Nascimento, T, &
Ernesto Hashime, K. (2015). Utilizaç~ao de enzimas proteolíticas para
produç~ao de hidrolisados proteicos a partir de carcaças de frango
desossadas manualmente. Brazilian Journal of Food Technology, 18(3),
199–210. https://doi.org/10.1590/1981-6723.4414
International, 73(1069, 204–212. https://doi.org/10.1016/j.foodres.
2015.04.005
Mora, L., Reig, M., & Toldrá, F. (2014). Bioactive peptides generated from
meat industry by-products. Food Research International, 65(PC), 344–
349. https://doi.org/10.1016/j.foodres.2014.09.014
Mullen, A. M., Alvarez, C., Zeugolis, D. I., Neill, E. O., & Drummond, L.
(2017). Alternative uses for co-products: Harnessing the potential of
valuable compounds from meat processing chains. Meat Science, 132,
90–98. https://doi.org/10.1016/j.meatsci.2017.04.243
Nollet, L. M. L., & Toldra, F. (Éd.). (2011). Handbook of Analysis of Edible
Animal By-Products. https://doi.org/10.1201/b10785
Normah, I., Jamilah, B., Saari, N., & Yaakob, B. C. M. (2005). Optimization of
hydrolysis conditions for the production of threadfin bream (Nemipterus
japonicus) hydrolysate by alcalase. Journal of Muscle Foods, 16(2), 87–102.
https://doi.org/10.1111/j.1745-4573.2005.07404.x
Ockerman, H. W., & Basu, L. (2004). BY-PRODUCTS j Edible, for Human
Consumption. Encyclopedia of Meat Sciences, 104–112. https://doi.
org/10.1016/b0-12-464970-x/00046-5
Rojas, M. J., Siqueira, P. F., Miranda, L. C., Tardioli, P. W., & Giordano, R. L.
C. (2014). Sequential proteolysis and cellulolytic hydro-lysis of soybean
hulls for oligopeptides and ethanol production. Indus-trial Crops and
Products, 61, 202–210. https://doi.org/10.1016/j. indcrop.2014.07.002
Roslan, J., Mustapa Kamal, S. M., Md. Yunos, K. F., & Abdullah, N. (2015).
Optimization of enzymatic hydrolysis of tilapia (Oreochromis niloticus)
byproduct using response surface methodology. International Food
Research Journal, 22(3), 1117–1123.
Roslan, J., Yunos, K. F. M., Abdullah, N., & Kamal, S. M. M. (2014). Charac-
terization of fish protein hydrolysate from tilapia (Oreochromis niloticus) by-
product. Agriculture and Agricultural Science Procedia, 2, 312–319.
https://doi.org/10.1016/j.aaspro.2014.11.044
Sadana, A. (1991). Biocatalysis Fundamentals of Enzyme Deactivation kinet-
ics. New Jersey: Prentice Hall.
Sánchez, A., & Vázquez, A. (2017). Bioactive peptides: A review. Food
Quality and Safety, 1(1), 29–46. https://doi.org/10.1093/fqsafe/ fyx006
Schmidt, C. G., & Salas-Mellado, M. (2009). Influência da aç~ao das enzimas
alcalase e flavourzyme no grau de hidrólise das proteínas de carne de
frango. Quimica Nova, 32(5), 1144–1150. https://doi.org/10.1590/ S0100-
40422009000500012
See, S. F., Hoo, L. L., & Babji, A. S. (2011). Skin by Alcalase. International
Food Research Journal, 18(4), 1359–1365.
Seong, P. N., Cho, S. H., Park, K. M., Kang, G. H., Park, B. Y., Moon, S. S.,
& Ba, H. V. (2015). Characterization of Chicken By-products by Mean of
Proximate and Nutritional Compositions. Korean Journal for Food Sci-
ence of Animal Resources, 35(2), 179–188. https://doi.org/
10.5851/kosfa.2015.35.2.179
Silva, J. F. X., Ribeiro, K., Silva, J. F., Cahú, T. B., & Bezerra, R. S. (2014).
Uti-lization of tilapia processing waste for the production of fish protein
hydrolysate. Animal Feed Science and Technology, 196, 96–106.
https://doi.org/10.1016/j.anifeedsci.2014.06.010
Shen, Q., Guo, R., Dai, Z., & Zhang, Y. (2012). Investigation of Enzymatic
Hydrolysis Conditions on the Properties of Protein Hydrolysate from Fish
Muscle (Collichthys niveatus) and Evaluation of Its Functional Properties.
Journal of Agricultural and Food Chemistry, 60(20), 5192– 5198.
https://doi.org/10.1021/jf205258f
Shimizu, M., Tanabe, S., Morimatsu, F., Nagao, K., Yanagita, T., Kato, N., &
Nishimura, T. (2006). Consumption of pork-liver protein Hydrolysate reduces
body fat in Otsuka long-Evans Tokushima fatty rats by suppressing hepatic
Lipogenesis. Bioscience, Biotechnology, and Bio-chemistry, 70(1), 112–118.
https://doi.org/10.1271/bbb.70.112
Slizyte, R., Rommi, K., Mozuraityte, R., Eck, P., Five, K., & Rustad, T. (2016).
Bioactivities of fish protein hydrolysates from defatted salmon back-
bones. Biotechnology Reports, 11, 99–109. https://doi.org/10.1016/j.
btre.2016.08.003
12 of 12 MALUF ET AL.

Srebernich, S. M., Gonçalves, G. M. S., & Domene, S. M. A. (2017). properties. Food Chemistry, 224, 160–171. https://doi.org/10.1016/j.
Fortifying pork liver mixture: Preparation and physicochemical foodchem.2016.12.057
characteristics - part 1. Food Science and Technology, 38(4), 647– White, J. A., Hart, R. J., & Fry, J. C. (1986). An evaluation of the Waters
652. https://doi.org/10. 1590/1678-457x.13517 Pico-Tag system for the amino-acid analysis of food materials.
Sun, H., Zhang, H., Ang, E. L., & Zhao, H. (2018). Biocatalysis for the Journal of Automatic Chemistry, 8(4), 170–177.
syn-thesis of pharmaceuticals and pharmaceutical intermediates. https://doi.org/10.1155/ s1463924686000330
Bioorganic and Medicinal Chemistry, 26(7), 1275–1284. Zheng, Z., Si, D., Ahmad, B., Li, Z., & Zhang, R. (2018). A novel
https://doi.org/ 10.1016/j.bmc.2017.06.043 antioxidative peptide derived from chicken blood corpuscle
Toldrá, F., Mora, L., & Reig, M. (2016). New insights into meat by- hydrolysate. Food Research International, 106, 410–419.
product utilization. Meat Science, 120, 54–59. https://doi.org/10.1016/j. foodres.2017.12.078
https://doi.org/10.1016/j. meatsci.2016.04.021 Zhou, C., Liu, H., Yuan, F., Chai, H., Wang, H., Liu, F., … Lu, F. (2019). Develop-
Venturin, A., Gonçalves, A. F. N., Oliveira, N. S., Skoronski, E., Pessatti, ment and application of a CRISPR/Cas9 system for Bacillus licheniformis
M. L., & Fabregat, T. E. H. P. (2017). Soluble fraction of sardine genome editing. International Journal of Biological Macromolecules, 122, 329–
protein hydrolysates in the feeding of the South American catfish. 337. https://doi.org/10.1016/j.ijbiomac.2018.10.170
Boletim Do Instituto de Pesca, 42(4), 878–888.
https://doi.org/10.20950/1678-2305.2016v42n4p878
Verma, A. K., Chatli, M. K., Kumar, P., & Mehta, N. (2017). Antioxidant
and antimicrobial activity of protein hydrolysate extracted from How to cite this article: Maluf JU, Fiorese ML, Maestre KL,
porcine liver. The Indian Journal of Animal Sciences, 87(6), 711–717. et al. Optimization of the porcine liver enzymatic hydrolysis
Retrieved from.
conditions. J Food Process Eng. 2020;e13370.
http://epubs.icar.org.in/ejournal/index.php/IJAnS/article/view/ 71070
Villamil, O., Váquiro, H., & Solanilla, J. F. (2017). Fish viscera protein hydro-
https://doi.org/ 10.1111/jfpe.13370
lysates: Production, potential applications and functional and bioactive

Vous aimerez peut-être aussi