You are on page 1of 7

International Journal of Food Science and Technology 2010, 45, 1807–1813 1807

Original article
Optimisation of extraction conditions and characteristics of skin
gelatin from Nile tilapia (Oreochromis niloticus)

Shaokui Zeng,1 Xiaoyan Yan,1 Wenhong Cao,1 Pengzhi Hong,1 Chaohua Zhang1* & Laihao Li2
1 Key Laboratory of Aquatic Product Advanced Processing of Guangdong Higher Education Institutes, College of Food Science & Technology,
Guangdong Ocean University, Zhanjiang, 524088, China
2 South China Sea Fishery Research Institute, Chinese Academy of Fishery Science, Guangzhou, 510300, China
(Received 21 December 2009; Accepted in revised form 4 June 2010)

Summary Response surface methodology was used to optimise gelatin extraction conditions from the skin of Nile
tilapia (Oreochromis niloticus), and characteristics of the gelatin were determined. Concentration of NaOH
(%, X1), alkaline treatment time (h, X2), concentration of HCl (&, X3) and acid treatment time (min, X4)
were chosen for independent variables. Dependent variable was yield of gelatin (%, Y). Optimal conditions
were X1 = 3.2 (%), X2 = 2.3 (h), X3 = 0.7 (&) and X4 = 84 (min), and predicted value of response
optimal conditions was Y = 20.4%. Actual value was 19.3% by verification experiments under optimal
conditions. Crude protein content of the tilapia skin gelatin was 88.5%. The content of imino acids including
proline and hydroxyproline in the gelatin was 185 residues per 1000 total amino acid residues. Its gel strength
was 260 g. The gelling and melting points were 18.0 and 22.4 C, respectively.
Keywords Amino acid, gels, response surface methodology, skin gelatin, tilapia.

China is by far the largest producer and exporter of


Introduction
tilapia in the world. According to the statistics of FAO,
Gelatin has been widely used in many industrial fields China’s annual production of tilapia in 2007 had risen to
such as food, material, pharmacy and photography, nearly 1.2 million ton, which accounts for about 50% of
especially in the food and pharmaceutical industries tilapia production of the world. Tilapia is usually
for its unique physico-chemical properties (Jamilah & processed as fresh and frozen fillets and exported to
Harvinder, 2002). Gelatin is a product of partial overseas markets. During tilapia fillet processing, large
hydrolysis of collagen and is usually extracted from quantities of by-products such as skin and scale are
porcine or bovine bones and skins. In the past produced. The skin is rich in collagen. If the skin can be
10 years, considerable effort has been expended on re-utilised to produce gelatin, it may increase the
studying gelatin extraction from different fish species economic value of the fish. According to previous
(Kołodziejska et al., 2004, 2008; Zhou & Regenstein, reports (Grossman & Bergman, 1992; Zhou et al.,
2005; Zhou et al., 2006; Cho et al., 2005; Liu et al., 2006), gelatins extracted from tilapia skin present good
2008; Muyonga et al., 2004; Aewsiri et al., 2008; gelling properties and high gel strength. Therefore, it can
Kwak et al., 2009). However, the production of fish serve as an additional source of extracting gelatin.
gelatin is still in its infancy, contributing only about Usually, the extraction procedures take a long time.
1% of the annual world gelatin production (Arnesen Jamilah & Harvinder (2002) previously reported a
& Gildberg, 2006). There are some limiting factors procedure for extracting tilapia skin gelatin, in which
such as insufficient availability of raw materials, the hot-water extracting process took about 12 h.
inferior rheological properties, prices. These problems However, Kołodziejska et al. (2008) used minced skins
have hampered the large-scale development of fish of cold-water fish instead of whole skins, the extraction
gelatin industry. time could be shortened from 12 h to 30 min. To
enhance industrial utilisation of tilapia fish gelatin, the
aim of this study is to optimise the pretreatment of the
*Correspondent: Fax: +86 0759 2382049; skin so as to shorten the extraction time and to determine
e-mail: chz2382@126.com the rheological properties of the gelatin.

doi:10.1111/j.1365-2621.2010.02332.x
 2010 The Authors. Journal compilation  2010 Institute of Food Science and Technology
1808 Optimisation of extraction conditions and characteristics of skin gelatin S. Zeng et al.

Table 1 Levels of independent variables used for tilapia skin gelatin


Materials and methods extraction and observations (Expt.) of gelatin yield (%) together with
predictions (Pred.) from the response surface model
Materials
Coded levels of variable Response (Y)
Oreochromis niloticus skin was provided by a local fish-
processing factory (Zhanjiang Global Aquatic Product Run no. X1 X2 X3 X4 Expt. Pred.
Co., Ltd., Zhanjiang, Guangdong, China). The skin was
1 )1 )1 )1 )1 12.91 14.89
mechanically separated from fresh tilapia while process-
2 )1 )1 )1 1 17.77 18.03
ing frozen fillets. The residue of adherent tissue such as
3 )1 )1 1 )1 11.47 14.89
scale was removed manually. The cleaned skin was 4 )1 )1 1 1 17.87 18.01
stored in )18 C for about 30 days until used. Gelatin 5 )1 1 )1 )1 15.62 14.89
extracted from the skin of porcine (9764) was purchased 6 )1 1 )1 1 17.41 18.03
from Amresco Co. (Solon, OH, USA). All reagents used 7 )1 1 1 )1 14.70 14.89
were of analytical grade. 8 )1 1 1 1 18.51 18.03
9 1 )1 )1 )1 10.74 14.89
10 1 )1 )1 1 15.62 18.03
Determination of proximate components 11 1 )1 1 )1 14.98 14.89
12 1 )1 1 1 21.33 18.03
Moisture (oven-drying procedure), crude protein and ash
13 1 1 )1 )1 13.30 14.89
contents were estimated by the AOAC official methods
14 1 1 )1 1 16.60 18.03
(Horwitz, 2000). The analyses were replicated three times. 15 1 1 1 )1 15.55 14.89
16 1 1 1 1 17.06 18.03
Extraction of gelatin 17 )2 0 0 0 18.29 14.93
18 2 0 0 0 12.69 14.93
The cleaned skin was treated with 10 volumes (v ⁄ w) of 19 0 )2 0 0 18.79 17.01
NaOH solution (1–5%, w ⁄ v) for 1–5 h at room 20 0 2 0 0 16.31 17.01
temperature (28–30 C) to remove the non-collagen 21 0 0 )2 0 18.99 20.13
protein and pigments with gentle stirring. The solution 22 0 0 2 0 18.58 20.13
23 0 0 0 )2 13.11 10.63
was changed when it became black. Alkaline-treated
24 0 0 0 2 15.50 16.91
skins were washed with distilled water until neutral.
25 0 0 0 0 18.82 20.13
Then the skin was soaked in HCl solution (0.1–0.9&) 26 0 0 0 0 19.20 20.13
for 30–120 min with gentle stirring and washed with 27 0 0 0 0 20.22 20.13
distilled water until neutral. For hot-water extraction, 6 28 0 0 0 0 22.88 20.13
volumes (v ⁄ w) of distilled water were added and heated 29 0 0 0 0 21.41 20.13
at temperature 60 C for 3 h. The extracted solution was 30 0 0 0 0 22.65 20.13
centrifuged at 900 g for 30 min. The upper phase was
X1, concentration of NaOH, %; X2, alkaline treatment time, h; X3,
vacuum-filtered with a Whatman No. 1 filter paper
concentration of HCl, &; X4, acid treatment time, min; Y, yield of
(Whatman International Ltd., Maidstone, UK). The
gelatin, %.
filtered solution was vacuum-concentrated to about
10% (w ⁄ v) at 60 C and dried at 1.4 m s)1 for 24 h in
a hot-air dryer (Shanghai Laboratory Instrument Works treatment time (h, X2), concentration of HCl (&, X3)
Co., Ltd, China) at 60 C. The powder obtained was and acid treatment time (min, X4) were chosen for
referred to as tilapia skin gelatin (TSG). independent variables. The range and centre point
values of four independent variables were based on the
results of preliminary experiments. After chemical
Experimental design
treatment, gelatin was extracted with water at 60 C
Central composite design (CCD) (Box & Wilson, 1951) for 3 h. The yield of gelatin (%, Y) was selected as the
was adopted in the optimisation of pretreatment the dependent variable. Experimental runs were randomised
tilapia skin. The samples were prepared according to a to minimise the effects of unexpected variability in the
4-factor, 5-level central composite design. CCD in the observed responses.
experimental design consists of twenty-four factorial
points, of which 8 are axial points (a = ±2) and six are
Analysis of data
replicates of the central point (Table 1). Processing fish
gelatin included two important processes – chemical The response surface regression (RSREG) procedure of
treatment and hot-water extraction. In general, a mild the Statistical Analysis System software (Version 8.01;
chemical treatment is used prior to gelatin extraction. In SAS Institute Inc., Cary, NC, USA) was used to fit the
our study, concentration of NaOH (%, X1), alkaline following quadratic polynomial equation

International Journal of Food Science and Technology 2010  2010 The Authors. Journal compilation  2010 Institute of Food Science and Technology
Optimisation of extraction conditions and characteristics of skin gelatin S. Zeng et al. 1809

X
4 X
4 3 X
X 4 Maries, VA, USA) with a load cell of 1000 N, cross-
Y ¼ b0 þ bi Xi þ bii X2i þ bij Xi Xj ð1Þ head speed 0.5 mm s)1 and equipped with a 1.27-cm-
i¼1 i¼1 i¼1 j¼iþ1 diameter flat-faced cylindrical plunger. The maximum
force (in grams), taken when the plunger had penetrated
where Y is dependent variable (yield of gelatin), b0 is
4 mm into the gelatin gels, was recorded. The determi-
intercept, bi, bii, bij are regression coefficients and Xi
nation was run in triplicate.
indicate the linear terms, Xi2 for the quadratic terms for
a single variable, Xi Xj for the interaction terms. For the
model calculated from the linear regression, analysis of Dynamic viscoelastic properties
variance (anova) was performed. The R2 value, the
Dynamic viscoelastic properties of TSG and PSG were
residual error, the pure error (calculated from the
measured by a concentric cylinder geometry of the
repeated measurements) and the lack of fit were calcu-
rheometer (Rheostress 1 RS30; HAAKE Co., Ltd,
lated. The lack of fit indicated whether the calculated
Karlsruhe, Germany). The gelatin solution (6.67%,
response surface represents the true surface. The sum of
w ⁄ v) was prepared in distilled water at 60 C. The
squares (SS) of the lack of fit was the SSresidual – SSpure
measurement was performed during cooling and heating
error, with its significance being tested with an F-test,
from 40 to 5 C and from 5 to 40 C, respectively. The
which was given by the relationship: F = MSlack of
temperature ramps were done at a scan rate of
fit ⁄ MSpure error, where MS was the mean sum of squares
0.5 C min)1, frequency 1 Hz, and oscillating applied
(Myers & Montgomery, 2002).
stress of 3.0 Pa. The elastic modulus (G¢; Pa), the
viscosity modulus (G¢¢; Pa) and the phase angle (d) were
Yield of extracted gelatin plotted as a function of temperature.
The yield was calculated as follows:
Determination of gelling and melting points
Yieldð%Þ ¼ dried gelatin (g)
 100=weight of wet skin used (g): Gelling and melting points were determined by the
method of Gudmundsson (2002). The gelling point was
evaluated from the intersection point, where the elastic
modulus (G¢; Pa) and the viscosity modulus (G¢¢; Pa)
Amino acid analysis
during the cooling process. The melting point was
Dry gelatin was dissolved in 6 m HCl solution and determined during the heating process in the same
hydrolysed at 110 C for 24 h. The solvent was analysed manner as for the gelling point.
by an amino acid analyzer (HITACHI 835-50 Amino
Acid Analyzer, Tokyo, Japan).
Results and discussion

SDS–polyacrylamide gel electrophoresis (SDS–PAGE) Diagnostic checking of the fitted model


SDS–PAGE was performed as described by Laemmli The RSREG procedure for SAS software was employed
(1970) using 7.5% separating gel and 4.0% stacking gel. to fit the quadratic polynomial equation to the exper-
TSG, porcine skin gelatin (PSG), molecular weight imental data. All the coefficients of linear (X1; X2; X3;
marker proteins (44.3–200 kDa; Biodee Biotechnology X4), quadratic (X12; X22; X32; X42) and interaction terms
Co., Ltd, Beijing, China) and acid-solubilised collagen were calculated for significance with t-statistic, and the
from tilapia skin (produced in our laboratory) were estimated coefficients of the model are presented in
applied and subjected to electrophoresis. After electro- Table 2. The coefficients of linear X4, quadratic X12 and
phoresis, the gel was stained with Coomassie brilliant X42 were highly significant (P < 0.01). The X22 was also
blue R-250. significant (P < 0.05). On the other hand, the coeffi-
cients of X32, all linear terms except X4 and all
interaction terms were not significant (P > 0.05). To
Determination of gel strength
develop the fitted response surface model equation, all
According to the method of Gómez-Guillén et al. (2002), insignificant terms (P > 0.05) were eliminated, and the
the gel strengths of TSG and PSG were determined on a final regression equation was formed:
6.67% gel (w ⁄ v), formed by dissolving the dry gelatin in
distilled water at 60 C and cooling the solution in a Y ¼ 20:13þ1:57X4 1:30X1 2 0:78X2 2 1:59X4 2 ; ð2Þ
refrigerator at 7 C (maturation temperature) for
16–18 h. The gel strength was determined at 7 C using where Y is the predicted yield of gelatin real value, X1,
samples with 3.3 cm diameter and 6 cm height on a Tms- X2 and X4 are the coded values of variable concentration
pro Testing Machine (Food Technology Corporation, of NaOH, alkaline and acid treatment time, respectively.

 2010 The Authors. Journal compilation  2010 Institute of Food Science and Technology International Journal of Food Science and Technology 2010
1810 Optimisation of extraction conditions and characteristics of skin gelatin S. Zeng et al.

Table 2 Coefficients of the response surface model of tilapia skin Response surface
gelatin yield
Figure 1 shows the estimated response function and the
Term Coefficient P value effects of the independent variables X1, X2, X4 on
the dependent variable Y. Fig. 1a, b and c suggest that
Intercept 20.1341 <0.0001
the yield of tilapia skin gelatin increases with the
X1 )0.5117 0.1775
X2 0.0458 0.9017
increase in concentration of NaOH, alkaline and acid
X3 0.4450 0.2384
treatment time to optimum point. When concentration
X4 1.5700 0.0003 of NaOH (X1), alkaline and acid treatment time (X2, X4)
X12 )1.2973 0.0005 further increase, the yield of gelatin decreased. Over
X2X1 )0.3988 0.3847 treatment, skin with alkaline and acid could derive
X22 )0.7823 0.0225 protein degradation to other low-molecular weight
X3X1 0.8638 0.0681 fragments. This resulted in the loss of extracted collagen
X3X2 )0.3575 0.4349 through leaching during the series of washing steps
X32 )0.4735 0.1508 (Jamilah & Harvinder, 2002). For gelatin manufacture,
X4X1 )0.0513 0.9102
the degree of collagen cross-linking is a key factor. Acid
X4X2 )0.7550 0.1076
X4X3 0.2025 0.6567
treatment would destroy the covalent and non-covalent
X42 )1.5935 <0.0001
crosslink and would result in collagen swelling
(Stainsby, 1987). Subsequent heating cleaves hydrogen
X1, concentration of NaOH, %; X2, alkaline treatment time, h; X3, and covalent bonds, leading to the conversion into
concentration of HCl, &; X4, acid treatment time, min. gelatin by a helix-to-coil transition (Djabourov et al.,
1993). Many previous studies demonstrated that acid
Table 3 Analysis of variance (anova) for the response surface model of concentration is very important for skin gelatin extrac-
tilapia skin gelatin yield tion (Gómez-Guillén & Montero, 2001; Giménez et al.,
2005; Kasankala et al., 2007). However, in our results,
Sources DF SS MS F value P value acid concentration (X3) is not significant. This may be
Regression 14 258.4624 18.4616 5.72 0.0002 related to the method used for pretreatment of the skin.
Linear 4 70.2439 17.5610 5.44 0.0036 The pretreatment by alkaline solution followed by an
Quadratic 4 161.8740 40.4685 12.54 <0.0001 acid solution in our study, not only removed the non-
Cross-product 6 26.3445 4.3908 1.36 0.2758 collagenous proteins but also provided the proper pH
Residual 21 67.7713 3.2272 – – condition for extraction (Zhou & Regenstein, 2005).
Lack of fit 10 40.4304 4.0430 1.63 0.2183
Pure error 11 27.3409 2.4856 – –
Total 35 326.2337 – – – Verification of predicted gelatin yield under optimal
conditions
Statistical testing of the regression model has been Skin gelatin production was conducted under optimal
done by the Fisher’s statistical test for anova (Table 3). conditions (concentration of NaOH 3.2%, alkaline
The probability (P) value of the regression model was treatment time 2.3 h, concentration of HCl 0.7& and
<0.01, with no significant lack of fit (P > 0.05), acid treatment time 84 min) to verify the predicted
indicating that the model is well adapted to the gelatin yield in the response surface model. The
response. The determination coefficient (R2 = 0.7923) extraction was carried out in distilled water at temper-
was satisfactory. The predicted value agreed well with ature 60 C for 3 h. Then, the solution was separated
the experimental one. and dried. The results of the three time repeated
experiments indicated that the actual value of gelatin
yield was 19.3 ± 2.3%, which is close to predicted
Conditions for optimum responses
value (20.4%).
Four independent variables, concentration of NaOH,
alkaline treatment time, concentration of HCl and acid
Chemical composition
treatment time were chosen as the critical factors of the
CCD model. The coded values of X1, X2, X3 and X4 for The contents of moisture, crude protein and ash in TSG
the optimised model calculated by response optimisers were 10.7, 88.5 and 0.6%, respectively. Amino acid
were 0.1202, )0.3080, 0.5360 and 0.3514, respectively. composition of the gelatin from tilapia skin is shown in
The actual values of X1, X2, X3 and X4 were 3.2%, 2.3 h, Table 4. The results indicated that the amino acid
0.7& and 84 min, respectively. The predicted value of composition of TSG was quite similar to that of warm-
extraction yield of gelatin was 20.4%, which is a water fish gelatins such as megrim and Dover sole
maximum. (Gómez-Guillén et al., 2002; Giménez et al., 2005). The

International Journal of Food Science and Technology 2010  2010 The Authors. Journal compilation  2010 Institute of Food Science and Technology
Optimisation of extraction conditions and characteristics of skin gelatin S. Zeng et al. 1811

(a)
2008). Ledward (1986) has reported that the stability of
the triple helical structures in gelatins is dependent on
the content of imino acids, as regions rich in proline and
hydroxyproline are likely to be involved in the forma-
tion of nucleation zones. A gelatin with high contents of
proline, hydroxyproline and alanine shows better visco-
elastic properties than others with low contents in these
amino acids (Gómez-Guillén et al., 2002; Sarabia et al.,
2000). The content of alanine in TSG was 12.6%. Thus,
TSG could be expected to have good rheological
properties.

Molecular weight distribution


The protein patterns of TSG and PSG are presented in
Fig. 2. In the case of PSG, the distinct band corre-
sponding to b component (200 kDa) in the electro-
(b) phoretic pattern was not visible. Over the whole length
of the gel, only a smudged band was visible. For TSG,
the distinct bands corresponding to b component of
tilapia skin collagen and to the one a chain (100 kDa)
in the electrophoretic pattern were visible. Additional
smudged bands with molecular weight lower than
100 kDa appeared below a chain. They could be the
products of hydrolysis of elementary chains of collagen.
According to Gómez-Guillén et al. (2002), damage or
partial loss of a1 chains can occur during the extraction
procedure. The difference of molecular weight distribu-
tions of TSG and PSG may be because of the different
content of covalent cross-links in the native collagen.
The results of Normand et al. (2000) suggested that
neither high nor low molecular weight chains contribute
to the elasticity. Besides, large amounts of b- and c-
chains have been shown to negatively affect some of the
(c) functional properties of fish gelatins, such as lowering
melting and setting points (Muyonga et al., 2004).
Therefore, TSG may have a lower gel strength, gelling
and melting points than PSG. This is most probably
because of a different molecular weight distribution as
well as the presence of different hydrolytic fragments
(Eysturskarð et al., 2009).

Gel strength
The gel strength of TSG was 260 ± 2.1 g, which was
lower than that of PSG (367 ± 1.9 g). The result was
similar to the reports of Grossman & Bergman (1992)
and Zhou et al. (2006). Songchotikunpan et al. (2008)
also reported th at the gelatin from Nile tilapia skin
showed high gel strength (328 g). Their differences could
Figure 1 The response surface for optimisation of gelatin extraction
from tilapia skin. X1 (concentration of NaOH, %); X2 (alkaline
be explained by differences in manufacturing process
treatment time, h); X4 (acid treatment time, min); Y (yield of gelatin, %). used. The study of Kwak et al. (2009) suggested that gel
strength of shark (Isurus oxyrinchus) cartilage gelatins
content of amino acids including proline and hydroxy- were different with drying methods. In addition, the
proline in TSG was 18.5%, which was lower than that in source and type of collagen will influence the properties
PSG (23%) and tuna fin gelatin (36%) (Aewsiri et al., of the resulting gelatins. According to the report of Zeng

 2010 The Authors. Journal compilation  2010 Institute of Food Science and Technology International Journal of Food Science and Technology 2010
1812 Optimisation of extraction conditions and characteristics of skin gelatin S. Zeng et al.

Table 4 Amino acid composition of gelatin from tilapia skin (residues


per 1000 total amino acid residues)
200.0 kDa
Amino acids

Aspartic acid 48
Threonine 27 116.0 kDa
Serine 39
Glutamic acid 71 97.2 kDa
Glycine 339
Alanine 126
Valine 19
66.4 kDa
Methionine 9
Isoleucine 9
Leucine 25
Tyrosine 4
Phenylalanine 13
44.3 kDa
Lysine 26
Histidine 5 1 2 3 4
Arginine 55
Proline 101 Figure 2 Protein patterns of tilapia and porcine skin gelatin. (1)
Hydroxyproline 84 molecular weight markers; (2) acid-solubilised collagen from tilapia
skin; (3) porcine skin gelatin (PSG); (4) tilapia skin gelatin (TSG).

et al. (2009), tilapia skin collagen molecule composed of


three a-chains and its denaturation temperature was and showed a lower melting temperature in comparison
high up to 32.0 C. Badii & Howell (2006) confirmed with porcine skin gelatin. Low gelling temperature
that hydrophobic amino acids (Ala, Val, Leu, Ile, Pro, (22.4 C) of TSG may offer new potential applications
Phe, and Met) could also contribute to the high bloom for dry products such as for micro-encapsulation.
value of tilapia gelatin. Montero & Gómez-Guillén
(2000) suggested that the hydrophobic amino acid
composition and distribution influence the physical Acknowledgments
properties of gelatin even more than the imino acids Project supported by the Foundation for Agriculture
content. The contents of Ala, Val and Leu in TGS were Project of Guangdong (Grant No. 2008A020100006,
higher than that in the reports of Gómez-Guillén et al. 2009B020201003) and the Important Projects in Key
(2002) and Sarabia et al. (2000). These may be the Fields in Guangdong and Hongkong (Grant No.
reasons for such high gel strength of TSG. Besides, the 2009A020700004)*. Authors also thank Engineer Wang
high bloom values of both gelatins in our study were Zhongming with Rousselot (Guangdong) Gelatin Co.,
related to the low temperature during curing. Ltd. for gelling and melting points determination.
*Correction added on 16 August 2010, after first Online
Publication. Grant No. 2009498D14 was changed to Grant
Gelling and melting points No. 2009A020700004.
The gelling and melting points can be judged from the
sharp decrease in phase angle (Fernández-Dı́az et al.,
References
2003). The gelling and melting points of TSG were
18.0 ± 0.3 and 22.4 ± 0.2 C, respectively. They were Aewsiri, T., Benjakul, S., Visessanguan, W. & Tanaka, M. (2008).
lower than that of PSG (25.8 ± 0.1 and 32.4 ± 0.1 C, Chemical compositions and functional properties of gelatin from
pre-cooked tuna fin. International Journal of Food Science &
respectively). The results were similar to the volumes of Technology, 43, 685–693.
previous literatures (Gudmundsson, 2002; Jamilah & Arnesen, J.A. & Gildberg, A. (2006). Extraction of muscle proteins
Harvinder, 2002). and gelatin from cod head. Process Biochemistry, 41, 697–700.
Badii, F. & Howell, N.K. (2006). Fish gelatin: structure, gelling
properties and interaction with egg albumen proteins. Food Hydro-
Conclusion colloids, 20, 630–640.
Box, G.E.P. & Wilson, K.B. (1951). On the experimental attainment of
The yield of gelatin extracted from Oreochromis niloticus optimum conditions. Journal of the Royal Statistical Society
skin was high up to 19.3 ± 2.3%. SDS–PAGE analysis (Series B), 13, 1–45.
Cho, S.M., Gu, Y.S. & Kim, S.B. (2005). Extraction optimization
showed that the gelatin was much less degraded than and physical properties of yellowfin tuna (Thunnus albacares) skin
gelatin from porcine skin. The gelatin showed high gelatin compared to mammalian gelatins. Food Hydrocolloids, 19,
strength (260 ± 2.1 g). It could not form a gel at 20 C 221–229.

International Journal of Food Science and Technology 2010  2010 The Authors. Journal compilation  2010 Institute of Food Science and Technology
Optimisation of extraction conditions and characteristics of skin gelatin S. Zeng et al. 1813

Djabourov, M., Lechaire, J. & Gaill, F. (1993). Structure and rheology cartilage gelatin produced by different drying methods. International
of gelatin and collagen gels. Biorheology, 30, 191–205. Journal of Food Science & Technology, 44, 1480–1484.
Eysturskarð, J., Haug, I.J., Ulset, A.-S. & Draget, K.I. (2009). Laemmli, U.K. (1970). Cleavage of structural proteins during the assembly
Mechanical properties of mammalian and fish gelatins based on of the head of bacteriophage T4. Nature (London), 227, 680–685.
their weight average molecular weight and molecular weight Ledward, D.A. (1986). Gelation of gelatin. In: Functional Properties of
distribution. Food Hydrocolloids, 23, 2315–2321. Food Macromolecules (edited by J.R. Mitchell & D.A. Ledward). Pp.
Fernández-Dı́az, M.D., Montero, P. & Gómez-Guillén, M.C. (2003). 171–201. London: Elsevier.
Effect of freezing fish skins on molecular and rheological properties Liu, H., Li, D. & Guo, S. (2008). Extraction and properties of gelatin
of extracted gelatin. Food Hydrocolloids, 17, 281–286. from channel catfish (Ietalurus punetaus) skin. LWT – Food Science
Giménez, B., Turnay, J., Lizarbe, M.A., Montero, P. & Gómez- and Technology, 41, 414–419.
Guillén, M.C. (2005). Use of lactic acid for extraction of fish skin Montero, P. & Gómez-Guillén, M.C. (2000). Extracting conditions for
gelatin. Food Hydrocolloids, 19, 941–950. megrim (Lepidorhombus boscii) skin collagen affect functional
Gómez-Guillén, M.C. & Montero, P. (2001). Extraction of gelatin properties of the resulting gelatin. Journal of Food Science, 65,
from megrim (Lepidorhombus boscii) skins with several organic 434–438.
acids. Journal of Food Science, 66, 213–216. Muyonga, J.H., Cole, C.G.B. & Duodu, K.G. (2004). Extraction and
Gómez-Guillén, M.C., Turnay, J., Fernández-Dı́az, M.D., Olmo, N., physico-chemical characterisation of Nile perch (Lates niloticus) skin
Lizarbe, M.A. & Montero, P. (2002). Structural and physical and bone gelatin. Food Hydrocolloids, 18, 581–592.
properties of gelatin extracted from different marine species: a Myers, R.H. & Montgomery, D.C. (2002). Experimental Designs for
comparative study. Food Hydrocolloids, 16, 25–34. Fitting Response Surfaces-I. Response Surface Methodology: Process
Grossman, S. & Bergman, M. (1992). Process for the production of and Product Optimisation Using Designed Experiments, 2nd edn.
gelatin from fish skins. US Patent, 5, 093, 474. John Wiley and Sons, Inc., New York, Pp. 303–376.
Gudmundsson, M. (2002). Rheological properties of fish gelatins. Normand, V., Muller, S., Ravey, J.C. & Paker, A. (2000). Gelation
Journal of Food Science, 67, 2172–2176. kinetics of gelatin: a master curve and network modeling. Macro-
Horwitz, W. (2000). Official Methods of Analysis of the Association of molecules, 33, 1063–1071.
Official Analytical Chemists, 17th edn. Washington DC, USA: Sarabia, A.I., Gómez-Guillén, M.C. & Montero, P. (2000). The effect
AOAC International. of added salts on the viscoelastic properties of fish skin gelatin. Food
Jamilah, B. & Harvinder, K.G. (2002). Properties of gelatins from Chemistry, 70, 71–76.
skins of fish-black tilapia (Oreochromis mossambicus) and red tilapia Songchotikunpan, P., Tattiyakul, J. & Supaphol, P. (2008). Extraction
(Oreochromis nilotica). Food Chemistry, 77, 81–84. and electrospinning of gelatin from fish skin. International Journal of
Kasankala, L.M., Xue, Y., Weilong, Y., Hong, S.D. & He, Q. (2007). Biological Macromolecules, 42, 247–255.
Optimization of gelatine extraction from grass carp (Catenophar- Stainsby, G. (1987). Gelatin gels. In: Collagen as a Food. Vol. 4 (edited
yngodon idella) fish skin by response surface methodology. Biore- by A.M. Pearson, T.R. Dutson & A.J. Bailey) Advances in meat
source Technology, 98, 3338–3343. research. Pp. 209–222. New York: Van Nostrand Reinhold.
Kołodziejska, I., Kaczorowski, K., Piotrowska, B. & Sadowska, M. Zeng, S.K., Zhang, C.H., Lin, H., Yang, P., Hong, P.Z. & Jiang, Z.H.
(2004). Modification of the properties of gelatin from skins of Baltic (2009). Isolation and characterisation of acid-solubilised collagen
cod (Gadus morhua) with transglutaminase. Food Chemistry, 86, from the skin of Nile tilapia (Oreochromis niloticus). Food Chem-
203–209. istry, 116, 879–883.
Kołodziejska, I., Skierka, E., Sadowska, M., Kolodziejski, W. & Zhou, P. & Regenstein, J.M. (2005). Effects of alkaline and acid
Niecikowska, C. (2008). Effect of extracting time and temperature pretreatments on Alaska pollock skin gelatin extraction. Journal of
on yield of gelatin from different fish offal. Food Chemistry, 107, Food Science, 70, C392–C396.
700–706. Zhou, P., Mulvaney, S.J. & Regenstein, J.M. (2006). Properties of
Kwak, K.-S., Cho, S.-M., Ji, C.-I., Lee, Y.-B. & Kim, S.-B. (2009). alaska pollock skin gelatin: a comparison with tilapia and pork skin
Changes in functional properties of shark (Isurus oxyrinchus) gelatins. Journal of Food Science, 71, C313–C321.

 2010 The Authors. Journal compilation  2010 Institute of Food Science and Technology International Journal of Food Science and Technology 2010

You might also like