Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Food Hydrocolloids 39 (2014) 68e76

Contents lists available at ScienceDirect

Food Hydrocolloids
journal homepage: www.elsevier.com/locate/foodhyd

Isolation and characterization of gelatin from the skins of skipjack


tuna (Katsuwonus pelamis), dog shark (Scoliodon sorrakowah), and
rohu (Labeo rohita)
K. Shyni*, G.S. Hema, G. Ninan, S. Mathew, C.G. Joshy, P.T. Lakshmanan
Biochemistry & Nutrition Division, Central Institute of Fisheries Technology, Cochin, India

a r t i c l e i n f o a b s t r a c t

Article history: Gelatin was extracted from the skins of dog shark (Scoliodon sorrakowah), skipjack tuna (Katsuwonus
Received 27 January 2013 pelamis) and rohu (Labeo rohita) and their physico-chemical properties were measured. The skins of
Accepted 4 December 2013 shark, tuna and rohu yielded 19.7, 17.2 and 11.3% gelatin, respectively. The gel strength of dog shark
gelatin (6.67%, 10  C) was found to be higher (206 g) than tuna and rohu skin gelatins (177 g and 124 g,
Keywords: respectively). Similarly, molecular weight, viscosity, melting point, foaming properties, water holding
Gelatin
capacity, odour, colour and clarity of dog shark gelatin were in general better than the tuna and rohu skin
Skipjack tuna
gelatins. The amino acid analysis showed that hydroxyproline content in dog shark skin gelatin was the
Dog shark
Rohu
highest when compared to tuna and rohu skin gelatins.
Gel strength Ó 2013 Elsevier Ltd. All rights reserved.
Fish gelatin

1. Introduction importance (Gomez-Guillen et al., 2009). This may be due to the


shortage of the primary raw materials mostly cattle hides, bones
Gelatin is a denatured fibrous protein derived from collagen by and pigskins (Spend Matters, 2012).
partial thermal hydrolysis. It is an important functional biopolymer Gelatins from land animal sources are preferred over marine
that has broad applications in the food, pharmacy and photography sources due to their superior gel strength, melting point and vis-
industries (Hao et al., 2009). The source, type of collagen and the cosity (Cho, Gu, & Kim, 2005). However, fish gelatin received
processing conditions will influence the properties of the resulting increasing attention as an alternative to land animal gelatin due to
gelatin. Different types of gelatins have varying thermal and religious constraints and health issues associated with the latter.
rheological properties such as Bloom strength, melting and gelling Both Judaism and Islam forbid the consumption of any pork-related
temperatures. These properties are governed by factors such as products and non-religiously slaughtered beef, while Hindus
chain length or molecular weight distribution, amino acid refrain from consuming cow-related products (Karim & Bhat,
composition and hydrophobicity, etc. (Gomez-Guillen et al., 2002; 2009). In addition, gelatin from aquatic sources has been shown
Norziah, Al-Hassan, Khairulnizam, Mordi, & Norita, 2009). to be free of infectious materials such as bovine spongiform en-
The global demand for gelatin has shown an increasing trend in cephalopathy (Sadowska, Kolodziejska, & Niecikowska, 2003).
recent years. Recent reports indicate that the annual world pro- The fish skins and bones contribute almost 30% of the total
duction of gelatin is nearly 326,000 tonnes, with pig skin derived weight of the fish (Gomez-Guillen et al., 2002). Fish skins are a
gelatin accounting for the highest (44%) output, followed by bovine major by-product of the fisheries and aquaculture industry. The
hides (28%), bovine bones (27%), and other sources (1%) (Ahmad & skin yield is highly variable according to species, fish size and
Benjakul, 2011). Other sources, which include fish gelatin, processing styles. Conversion of these wastes into value-added
accounted for around 1.5% of total gelatin production in 2007, but products such as gelatin to yield additional income has both eco-
this percentage was double that in 2002, indicating that gelatin nomic and waste management benefits for the fish industry (Choi &
production from alternative non-mammalian species had grown in Regenstein, 2000).
A number of studies have addressed properties of fish skin
gelatins (Arnesen & Gildberg, 2007; Choi & Regenstein, 2000;
* Corresponding author. Tel.: þ91 4842666845; fax: þ91 4842668212.
Fernandez-Diaz, Montero, & Gomez-Guillen, 2001; Gomez-
E-mail address: shynicof@yahoo.co.in (K. Shyni). Guillen & Montero, 2001; Grossman & Bergman, 1992;

0268-005X/$ e see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.foodhyd.2013.12.008
K. Shyni et al. / Food Hydrocolloids 39 (2014) 68e76 69

Gudmundsson, 2002; Gudmundsson & Hafsteinsson, 1997; Holzer, skin/solution ratio of 1:10 (w/v) for 2 h. The alkaline solution was
1996; Jamilah & Harvinder, 2002; Jongjareonrak, Benjakul, Vises- changed every 1 h to remove non-collagenous proteins and pig-
sanguan, & Tanaka, 2006; Jongjareonrak, Benjakul, Visessanguan, ments. Alkaline-treated skins were washed with tap water until
Prodpran, & Tanaka, 2006; Yang et al., 2007; Zhou & Regenstein, the wash water was neutral or faintly basic. The pH of wash water
2005) showing that their properties differ from those of mamma- was monitored using Cyberscan 510 pH meter (Eutech In-
lian gelatins. The gelatin properties were found to vary with the struments Pte Ltd, Singapore). The skins were then soaked in
method of production and between species. 0.2 M acetic acid with a skin/solution ratio of 1:10 (w/v) for 24 h
Around the world, sharks are mainly used for producing shark with gentle stirring at 4  C. The acidic solution was changed every
fins and fillets. As a result, skins are discarded along with the rest of 12 h to swell the collagenous material in the fish skin matrix. Acid-
the carcass or treated as a by-product with low market value. pretreated skins were washed thoroughly with tap water until the
Studies have shown that cartilaginous fishes including sharks have wash water became neutral. The skins were then subjected to a
a higher content of collagen, the precursor of gelatin, when final wash with distilled water to remove any residual matter. The
compared to bony fishes (Nair, 2002). Cho et al. (2005) reported final extraction was carried out in distilled water at 45  C for 12 h
that gelatin extracted from yellowfin tuna skin showed better with a skin/water ratio of 1:10 (w/v). The clear extract obtained
functional properties than those from other fish sources. Some was filtered through a Buchner funnel with Whatman filter paper
studies have ascertained that freshwater fish can have a high No. 4 (Sunshine Instruments, Coimbatore, Tamil Nadu, India). Fat
gelatin yield (Grossman & Bergman, 1992; Jamilah & Harvinder, separation was done using a simple fat separating funnel (India-
2002; Muyonga, Cole, & Duodu, 2004a). Only a few studies have MART, Maharashtra, India). The extract was poured in to the
been conducted on warm-water fish gelatin and these showed that funnel and was allowed to settle for few seconds. Within few
gelatin from these species had better functional properties than seconds, suspended fat formed a layer over the gelatin extract.
those from cold-water fish species (Gilsenan & Ross-Murphy, 2000; Then the gelatin was allowed to flow slowly through the outlet
Grossman & Bergman, 1992; Leuenberger, 1991). valve of the funnel. The valve was carefully closed on reaching the
Dog shark and skipjack tuna are available along the south west level of suspended fat layer. The clear extract obtained was
coast of India. Rohu is one of the major carp species, a natural concentrated by evaporation under vacuum at 5  C, with a flash
inhabitant of the freshwater sections of the rivers of India and evaporator (Buchi rotavapor R215, Buchi Labortechnik, Flawil,
contributes to the inland catch. At present, the fishery is sustainable Switzerland). The concentrated viscous solution was frozen in an
for all the above species. air blast freezer (Icematic T10, CastelMAC SpA, Castelfranco Veneto
The present study was undertaken to extract and characterize (TV), Italy) at 40  C and then freeze-dried with freeze drier
gelatin from the skins of skipjack tuna (Katsuwonus pelamis) from (Gamma 1e16 LSC, Osterode am Harz, Germany) in two steps. The
the family scombridae, dog shark (Scoliodon sorrakowah) from the solution was subjected to main drying for 8 h at set shelf tem-
family of carcharhinidae, a cartilaginous fish and rohu (Labeo perature of 20  C and set pressure of 0.01 mbar. The final drying
rohita) from the family of cyprinidae, a freshwater fish. was done for 2 h at set shelf temperature of 25  C and set pressure
of 0.01 mbar at a condenser temperature of 55  C.
2. Materials and methods
2.3. Yield of gelatin
2.1. Raw material
The yields of the gelatins obtained were calculated as:
The species used for the study were skipjack tuna (K. pelamis),
dog shark (S. sorrakowah), and rohu (L. rohita). The marine fishes Dry wt: of gelatin
% Yield ðwet wt: basisÞ ¼  100
were landed from long-line day boats that iced the fish on board, at Wet wt: of skins
the local fish landing centre in Cochin during May 2011, from the
Arabian sea. The length of the fishes were measured and recorded,
i.e., tuna (92  2.4 cm, 17 in number) shark (73  2.9 cm, 23 in
number). The skinning of the fish was carried out by the fisherman 2.4. Determination of proximate composition
at the market under supervision. The iced skin in prime quality was
brought to the laboratory. Rohu (39  6.6 cm, 55 in number) freshly Moisture, lipid, ash and protein were determined by AOAC
caught and dead, in iced condition, was procured from a local fish (1995) methods 950.46, 960.39, 900.2A and 928.08, respec-
farm and skinning was carried out at laboratory in the iced tively. Protein digestion was done as described by Eastoe and
condition. Eastoe (1952) to ensure complete hydrolysis of collagen. A con-
The skins were cut into 2e3 cm2 pieces using a scalpel and version factor of 5.4 was used for calculating the protein content
washed with ice cold tap water. They were then placed in poly- from the Kjeldahl nitrogen content since collagen, the main
ethylene bag with added glaze water (10%) and stored at 20  C protein in skin, contains approximately 18.7% nitrogen (Eastoe &
until use. The storage time was less than 2 months. All chemicals Eastoe, 1952). This is an estimate and is based on mammalian
used unless otherwise noted were of analytical grade (Merck KGaA gelatin.
Chemical & Pharmaceutical Company, Darmstadt, Germany;
SigmaeAldrich Corporation, MO, USA). 2.5. Determination of pH

2.2. Gelatin extraction The pH values of raw fish skins and gelatin solutions were
measured using the British Standard Institution method (BSI, 1975).
Based on preliminary extraction trials, it was decided to follow For determining the pH of the skins, samples were chopped and
the gelatin extraction method of Gudmundsson and Hafsteinsson blended for 5 min at ambient temperature (37  C) by vigorous
(1997). Thawed skins were thoroughly washed and treated with shaking (ICS-BLENDER, Hyderabad, Andhra Pradesh, India) in
warm water (38e40  C) for 10 min to remove superfluous material distilled water to form a 1% (w/v) skin suspension. For the gelatin
and reduce the fat content. Before gelatin extraction, skins were solution, a 1.0% (w/v) gelatin solution was prepared by adding 1 g of
soaked in 0.1 M NaOH at ambient temperature (w27  C) with a gelatin in 99 ml of distilled water. The mixture was heated to 45  C
70 K. Shyni et al. / Food Hydrocolloids 39 (2014) 68e76

for 5 min, for dissolving gelatin. The solution was allowed cool 5 kg and crosshead speed of 0.5 mm/s. The maximum force (g) was
down to room temperature before pH measurement. The pH was determined when the probe penetrated to a depth of 4 mm into the
measured using a Cyberscan 510 pH meter (Eutech Instruments Pte gel.
Ltd, Singapore). The test determines the weight (in grams) needed by a probe
(normally with a diameter of 0.5 inches) to deflect the surface of the
2.6. Gelatin colour and gel clarity gel 4 mm without breaking it. The result is expressed as Bloom
(grams). Gelatin, if it gels, is usually between 30 and 300 Bloom.
Instrumental colour measurements of the dry gelatin samples
were made using a HunterLab MiniScan XE Plus Spectrocolorimeter 2.10. Determination of amino acid composition
(Hunter Associates Laboratory Inc., Reston, VA, USA) based on three
colour co-ordinates, namely L* (lightness), a* (redness/greenness) A dry sample (100 mg) of gelatin was weighed and taken in a
and b* (yellow ness/blueness). The equipment was standardized test tube. 10 ml 6 N HCl was added to in to the test tube. Then the
using a white tile and black glass. The gelatin sample was filled into tubes were filled with nitrogen, sealed and hydrolyzed at 110  C for
the glass sample cup and the cup was fed into the instrument. 24 h. After the hydrolysis, the contents were quantitatively trans-
Instrumental colour measurement readings were taken from the ferred into a round bottom flask through Whatman filter paper. No.
instrument monitor. Visual observations for colour were also noted. 42. The contents were flash evaporated to dryness. Distilled water
Clarity was determined by measuring transmittance (%T) at was added and flash evaporated again to remove traces of HCl. The
620 nm in spectrophotometer (Spectronic Instruments Inc., process was repeated thrice. The residue was then dissolved and
Rochester, N.Y., USA) through 6.67% (w/v) gelatin solutions which made up to 5e10 ml (known volume) with 0.05 M HCl. The sample
were heated at 60  C (Julabo TW20 water bath, Allentown, PA, USA) was then filtered through a membrane filter (Thermo Scientific,
for 1 h (Avena-Bustillos et al., 2006). MK, Buckinghamshire, England) with a nominal pore size of
0.45 mm. The sample (20 ml) was injected into the amino acid
2.7. Determination of odour analyzer (HPLCeLC 10 AS, Hitachi L-2130 Elite La Chrome, Tokyo,
Japan) and the amino acid composition was determined as per the
Sensory evaluation was conducted using a seven member expert method of Ishida, Fugita, and Asai (1981). The system consists of a
panel according to the method of Muyonga et al. (2004a). The binary pump (L-7100), autosampler (L-2200), column oven (L-
samples were prepared in test tubes with screw caps, by dissolving 2350), fluorescent (FL) detector (L-2485) and a post-column deri-
0.5 g of gelatin in 7 ml of distilled water, thus obtaining a solution vatisation unit fitted with a peristaltic pump. The column used is a
containing approximately 6.67% gelatin. The tubes were then held cation exchange column (Shodex, CX Pak, 4.6  15 mm) connected
in a water bath at 50  C for dissolving, with the screw caps closed. with a guard column. The eluents used were buffer A (13.3 g tri-
Panelists were instructed to remove the screw caps, sniff the con- sodium citrate, 70 ml ethanol, 12.8 ml citric acid monohydrate,
tents and identify the odour they perceived as well as to indicate 3.74 g NaCl, 4 ml Brij (SigmaeAldrich Corporation, MO, USA), pH 3.2
the odour intensity, using a six point scale. i.e., 0 ¼ no odour, made up to 1 l with distilled water) and buffer B (117.6 g trisodium
1 ¼ very mild and only perceivable on careful assessment, 2 ¼ mild citrate and 24.8 g boric acid, 500 ml distilled water, 45 ml 4 N
but easily perceivable, 3 ¼ strong but not offensive, 4 ¼ strong and NaOH, pH to 10 and make up to 2 l with distilled water). The sep-
offensive, 5 ¼ very strong and very offensive. aration is effected using a gradient programme (0 min-100% A,
30 min-50% B). The eluted amino acids are subjected to post-
2.8. Determination of viscosity column derivatisation with ortho-phthalaldehyde (OPA) reagent
delivered using a peristaltic pump at the rate of 0.5 ml/min at 30  C.
The viscosity of the gelatin (6.67% concentration at 60  C) was Detection was done using a FL detector with absorption wave
measured using a Brookfield digital viscometer (model DV-E, length of 340 nm and emission wave length of 450 nm. Amino acid
Brookfield Engineering, Middleboro, MA, USA) equipped with a standards for collagen hydrolyzate (Sigma A-9531) were run to
No. 1 spindle at 30  0.5  C. The measured values were obtained calculate the concentration of amino acids in the sample. Base line
directly in centipoises (cP) from the instrument. auto-zero was made through software D-2000 Elite, Hitachi. The
method showed a linear response (R2 0.99) in this range. The
2.9. Determination of gel strength identification of amino acids in the sample is done using D-2000
Chromatography Data Station Software. Tryptophan analysis was
The gel strength (Bloom) was determined according to British carried out according to colorimetric method (Sastry & Tammuru,
Standard 757:1975 method (BSI, 1975), by using a texture analyzer 1985).
(Lloyd Instruments, Model LRX Plus, Sussex, U.K.). A solution con-
taining 6.67% (w/v) gelatin was prepared by mixing 7.50 g of gelatin 2.11. Determination of molecular weight distribution
and 105 ml of distilled water in a Bloom bottle (Schott Duran,
Mainz, Rhineland-Palatinate, Germany). The mixture was swirled Electrophoretic patterns of the different gelatins were analyzed
and left to stand at room temperature for 1 hr, for allowing the according to the method of Laemmli (1970). The samples were
gelatin to absorb water and swell. The Bloom bottles were then dissolved in 50 g/l SDS solution. The mixtures were then heated at
transferred to a water bath (Julabo TW20, Allentown, PA, USA) 85  C for 1 h, followed by centrifugation at 8500  g for 5 min at
maintained at 65  C and held for 25 min with occasional swirling to room temperature using a microcentrifuge (MIK-RO20, Hettich
dissolve the gelatin. The bottles were taken out of the water bath, Zentrifugen, Germany) to remove undissolved debris if any. Solu-
allowed to cool for 15 min at room temperature and then placed in bilized samples were mixed (1:1 v/v) with the sample buffer
a cold-water bath (circulating bath, Haake D3 Model, Berlin, Ger- (0.5 mol/l TriseHCl, pH 6.8 containing 40 g/l SDS, 200 ml/l glycerol
many) maintained at 10  0.1  C and held at this temperature for in the presence or absence of 100 ml/l b mercaptoethanol). The
16e18 h before the determination of the gel strength. The Bloom mixtures were loaded onto a polyacrylamide gel having a 7.5%
bottle was placed centrally under the plunger (Delrin probe, which resolving gel (10% SDS: 100 ml, acrylamide: 2.5 ml, ammonium
is an acetal homopolymer, 12.7 mm diameter, black in colour) of the persulphate (APS) 10%: 50 ml, distilled water: 4.85 ml, TEMED: 5 ml,
instrument. The Bloom strength was determined with a load cell of TriseHCl, 1.5 M: 2.5 ml) and a 4% stacking gel (10% SDS: 100 ml,
K. Shyni et al. / Food Hydrocolloids 39 (2014) 68e76 71

acrylamide: 1.33 ml, APS 10%: 50 ml, distilled water: 6.1 ml, tetra- at intervals of 15 s. The time at which the rod could not detach from
methylethylenediamine (TEMED): 10 ml, TriseHCl, 1.5 M: 2.5 ml). the gelatin sample was recorded as the setting time.
The electrophoresis was done till the base line reached bottom of
the gel, at a constant current of 20 mA per gel using a Bio-Rad Tetra
Mini Protean II unit (Bio-Rad Laboratories, Inc., Richmond, CA, USA) 2.14. Determination of foaming properties
with gel documentation system. After electrophoresis, gels were
fixed with a mixture of 500 ml/l methanol and 100 ml/l acetic acid Foam formation ability (FA) and foam stability (FS) of gelatin
for 30 min, followed by staining with 0.5 ml/l Coomassie blue R- were determined by the procedure of Cho et al. (2004). Gelatin
250 (ultra-pure grade) in 150 ml/l methanol and 50 ml/l acetic acid solution, 1 g/100 ml was put in a beaker and swollen at 60  C. The
for 1 h. Finally, they were destained with a mixture of 300 ml/l foam was prepared by homogenizing at 10,000 rpm for 5 min in a
methanol and 100 ml/l acetic acid for 1 hr and destained again with homogenizer (Euro Turrax T20b, IKA Labortechnik, Staufen, Ger-
the same solution for 30 min. High molecular weight protein many). The homogenized solution was then poured into a 250 ml
marker (Sigma Chemical Co., St. Louis, MO, USA) with a molecular measuring cylinder. The FA was calculated as the ratio of volume of
weight range of 36,000e205,000 Da was used to estimate the foam to the initial volume of liquid. The foam stability was calcu-
molecular weight of proteins. The molecular weight distribution of lated as the ratio of the initial volume of foam to the final volume of
sigma marker (Da) used in the study is foam after 30 min.

 Myosin, rabbit muscle e 205,000


 b-Galactosidase, Escherichia coli e 116,000 2.15. Determination of water holding capacity and fat binding
 Phosphorylase b, rabbit muscle e 97,000 capacity
 Fructose-6-phosphate Kinase, rabbit muscle e 84,000
 Albumin, bovine serum e 66,000 Fat binding capacity (FBC) and water holding capacity (WHC) of
 Glutamic dehydrogenase, bovine liver e 55,000 gelatin were determined as per the procedure of Cho et al. (2004).
 Ovalbumin, chicken egg e 45,000 For measuring FBC, 1 g of gelatin powder was placed in a
 Glyceraldehyde-3-phosphate dehydrogenase, rabbit muscle e centrifuge tube and weighed (tube with gelatin). Then, 10 ml sun-
36,000 flower oil was added, and held at room temperature for 1 h. During
this period, the gelatin solutions were mixed with a Vortex mixer
(CM-101 Plus, REMI Instruments, Maharashtra, India) for 5 s every
15 min. The gelatin solutions were then centrifuged at 450  g
2.12. Determination of melting point of gelatin
(Model CPR 24, REMI Instruments, Maharashtra, India) for 20 min
with cylinder bottom centrifuge of 20 ml capacity (REMI In-
The melting point measurement was done according to the
struments, Maharashtra, India). The upper phase was removed by
method of Wainewright (1977). Gelatin solutions, 10% (w/w), were
tilting the centrifuge tube to 45 angle and draining on to a filter
prepared and a 5 ml aliquot of each sample was transferred to a
paper for 30 min. The FBC was calculated as the weight of the
small culture tube (12  75 mm). The samples were degassed in
contents of the tube after draining divided by the weight of the
vacuum desiccators (Heraeus vacutherm, Hohenpriessnitz, Sach-
dried gelatin, and expressed as the weight % of dried gelatin.
sen, Germany) for 5 min. The tubes were then covered with Par-
For measuring WHC, 1 g of gelatin powder was placed in a
afilm (Pechiney Plastic Packaging, Inc., Chicago, IL, USA) and heated
centrifuge tube and weighed (tube with gelatin). Distilled water
in a water bath at 60  C for 15 min. The tubes were immediately
(50 ml) was added, and held at room temperature for 1 h. During
cooled in ice-chilled water and matured at 10  C for 18 h. Five drops
this period, the gelatin solutions were mixed with a Vortex mixer
of a mixture of 75% chloroform and 25% red food colour (Magnil
(CM-101 Plus, REMI Instruments, Maharashtra, India) for 5 s every
Dye Chem, Mumbai, Maharashtra, India) was placed on the surface
15 min. The gelatin solutions were then centrifuged at 450  g
of the gel. The gels were put in a water bath (Haake D3, Lab
(Heraeus Multifuge 3SR Plus, Thermo Scientific, MK, Buck-
Extreme, Inc., Kent City, MI, USA) at 10  C and the bath was heated
inghamshire, England) for 20 min with 600 ml bucket centrifuge
at the rate of 0.2  C/min. The temperature at which the dye drops
(Thermo Scientific, MK, Buckinghamshire, England). The upper
began to move freely down the gel was taken as the melting point.
phase was removed and the centrifuge tube was drained for 30 min
on a filter paper after tilting to 45 angle. WHC was calculated as
2.13. Determination of setting point and setting time the weight of the contents of the tube after draining divided by the
weight of the dried gelatin, and expressed as the weight % of dried
The method used for the determination of setting point (SP) and gelatin.
setting time (ST) of gelatin was that described by Muyonga et al.
(2004a). Gelatin solutions of 10% (w/v) dissolved in thin wall
(12  75 mm) test tubes were prepared in the same way as 2.16. Statistical data analysis
described for the Bloom samples. For setting point determination,
the dissolved samples from the warm-water bath were transferred One way analysis of variance was used to find the effect of
to a circulating water bath held at 40  C (Haake D3). The bath was different species on yield and functional properties of gelatin. Once
then cooled at the rate of 2  C/min. A thermometer was inserted ANOVA was found significant, Tukey’s test was used as a post-hoc
into the sample and lifted out at 15 s intervals. The temperature of test to compare the means of different species. The statistical
the mixture at which the gelatin solution no longer dripped from analysis was carried out using SAS. 9.3 software (SAS Institute Inc,
the tip of the thermometer was recorded as the setting Cary, NC, USA). The mean value of three replications and its stan-
temperature. dard deviations are shown in the tables. Letters are placed as a
Setting time was determined on samples prepared in the same superscript along with the mean to indicate significant difference
way as those for the determination of the setting temperature. between the means in all the tables except the amino acid table, i.e.,
Samples were transferred to the water bath maintained at 10  C means with different superscripts are statistically significant at 1%
(Haake D3). A rod was inserted into the gelatin solution and raised level of significance (p < 0.01).
72 K. Shyni et al. / Food Hydrocolloids 39 (2014) 68e76

3. Result and discussion 88.6% in gelatin extracted from the skins of bigeye snapper and
brown eye snapper, respectively. The gelatin from the skins of adult
3.1. Proximate composition of fish skins Nile perch also contained 88% protein when extracted at 50  C
(Muyonga et al., 2004a). Moisture content in all the samples was
The protein, moisture and ash content values of three selected well below the prescribed limit of 15% (GME, 2005) for edible
fish skins after cleaning, are tabulated in Table 2. All three fish skins gelatin. At 6e8% moisture, gelatin is very hygroscopic and it be-
had considerably higher ash contents (2.03e4.39%) compared to comes difficult to determine the physico-chemical attributes with
that of the Red tilapia (0.51%), Walking catfish (0.52%) and Striped accuracy (Cole, 2000). The gelatin samples extracted were almost
catfish (0.46%) (Jamilah, Tan, Umi Hartina, & Azizah, 2011). free of fat. This showed that the simple de-fattening method fol-
Compared to the other two species, ash content in rohu skins was lowed here had eliminated the fat content as desired. The gelatins
lower, probably due to perfect de-scaling. De-scaling was not car- were found to be low in ash content, well below the recommended
ried out for shark and tuna skins since the scales were too tiny for maximum of 2.6% (Jones, 1977).
tuna and was deeply embedded in the shark skin. Shark skins
contained comparatively higher amount of protein (27.7%) than
other two skins. The skins of rohu had highest moisture content, 3.4. The pH values of gelatin samples
followed by shark and tuna skins. Subcutaneous fat was highest
(18.3%) for tuna skins when compared with shark and rohu skins. The pH of extracted gelatins varied between 4.17 and 4.34 as
This can be explained as the site of lipid storage in shark is liver, shown in Table 1, indicating their category as Type B. It has been
whereas it is skin and viscera in tuna. The results of lipid can be reported that alkali pretreatment results in Type B gelatin with pH
correlated with the moisture content in reverse order in the above in the range of 4e5 (Baziwane & He, 2003) and in the present study
species. an alkali pretreatment was employed during the extraction of
gelatin. Wide variations in the pH of skins gelatin of cod (2.7e3.9)
(Gudmundsson & Hafsteinsson, 1997), red tilapia (3.05) and black
3.2. Yield of gelatin
tilapia (3.91) (Jamilah & Harvinder, 2002), freshwater carps (4.01e
4.88) (Ninan, Abubacker, & Jose, 2011; Ninan, Jose, & Abubacker,
Gelatin yield is calculated as g of gelatin per 100 g of clean skins.
2011) and bigeye snapper (6.44) (Binsi, Shamasundar, Dileep,
The yield of shark skin gelatin (SSG), tuna skin gelatin (TSG) and
Badii, & Howell, 2009) have been reported. This was, most likely
rohu skin gelatin (RSG) were significantly different (p < 0.01) and
due to the different pretreatments employed during the extraction
were 19.7  0.04%, 11.3  0.03% and 17.2  0.03%, respectively. It
involving both alkaline and acid treatments.
was observed that shark skins tend to swell more in the alkaline
It has also been reported that for Type B gelatin, the viscosity is
and acidic solutions compared to the skins of rohu and tuna.
minimum and gel strength is maximum at pH 5.0 (Cole, 2000)
Therefore, shark skins gave a better yield; possibly due to increased
signifying the importance of pH for gelatin’s rheological properties.
opening of cross-links during swelling. As a cartilaginous fish, shark
Hence from the manufacturer’s point of view it is advantageous to
has comparatively more connective tissue and hence more collag-
manufacture gelatin with a pH of 5.0. However, due to the strong
enous material than bony fishes. This can be correlated with the
buffering capacity of gelatin, this pH may not be the most advan-
high yield of gelatin from shark skins.
tageous for the customer.
Gelatin processing has three steps, alkali pretreatment, acid
pretreatment and hot water extraction. The alkali & acid treatment
removes non-collagenous protein and the sample swells in the acid 3.5. Gelatin colour and gel clarity
solution. The hot water extraction uses thermohydrolysis to solu-
bilize gelatin which is then separated. The lower yield observed in Both colour and clarity of a gelatin gel are important aesthetic
rohu could be due to the leaching of collagen during the washing properties, depending on the application for which the gelatin is
treatment or insufficient denaturation of collagen during extraction intended. Instrumental colour measurements of gelatins are tabu-
(Jamilah & Harvinder, 2002). The crude protein content and hy- lated in Table 2. The colour of gelatin depends on the raw materials
droxyproline content were also lower in rohu skins. used and method of extraction (Ockerman & Hansen, 1999). In
general, colour does not influence the functional properties. How-
3.3. Proximate composition of gelatin ever light colour is preferred because it is easier to incorporate
gelatins into any food systems without imparting any strong colour
The proximate compositions of extracted gelatins are tabulated attribute to the product. According to the instrumental colour
in Table 1. The gelatin extracted from skins of the three species measurement, shark skins gelatin was significantly lighter/whiter
showed protein as the major component and protein content varied than the other two samples, Rohu skins gelatin was the greenest
between 88.4 and 90.1%. Jongjareonrak, Benjakul, Visessanguan, and it was significantly (p < 0.01) less yellowish when compared to
Prodpran, et al. (2006) reported a protein content of 87.9% and shark and tuna gelatins. By visual observation, shark and tuna

Table 1
Proximate composition and pH of skins and gelatins from shark, rohu and tuna.

Sample Components and pH

Moisture Protein Fat Ash pH

Skins Shark 68.4  0.43 27.7  0.36 0.16  0.02 4.19  0.03 6.23  0.01
Rohu 76.5  0.45 18.8  0.06 2.93  0.05 2.03  0.04 6.37  0.02
Tuna 56.5  0.09 20.5  0.26 18.3  0.11 4.39  0.03 6.31  0.01
Gelatin Shark 8.7  0.25b 90.1  0.30a 0.42  0.05b 0.72  0.06a 4.34  0.01a
Rohu 9.3  0.36b 89.2  0.49b 0.48  0.08b 0.73  0.07a 4.17  0.01c
Tuna 10.9  0.24a 88.4  0.12b 0.82  0.05a 0.68  0.06a 4.29  0.01b

Values are given as mean  standard deviation of triplicate. Values with the same superscript letters within a column are not significantly different (p < 0.01).
K. Shyni et al. / Food Hydrocolloids 39 (2014) 68e76 73

Table 2 gelatin has been observed in the pH range of 6e8 (Stainsby, 1987).
Instrumental colour, visual observation and transmittance of shark skin gelatin The above results thus indicate that natural variations in the vis-
(SSG), rohu skin gelatin (RSG) and tuna skin gelatin (TSG).
cosity can be expected from different fish species, although the
Sample L* a* b* Colour Transmittance extraction methods do play a role.
SSG 82.8  0.40a 1.65  0.04b 18.3  0.21a Pearly 44.8  0.72a
white
RSG 78.1  0.62b 0.75  0.01c 5.36  0.04b Shiny 44.4  0.59a
3.8. Gel strength values of gelatin samples
whitish
TSG 75.3  0.60c 9.13  0.04a 18.4  0.17a Pearly 35.5  0.87b Gel strength (Bloom) is the most important physical property of
white a gelatin. The gel strengths of shark, rohu and tuna gelatins were
Values are given as mean  standard deviation of triplicate. Values with the same 206,124 and 177 g, respectively. All three samples were significantly
superscript letters within a column are not significantly different (p < 0.01). (p < 0.01) different (Table 3). The gel strength of fish gelatin has
been reported in a wide range of 124e426 Bloom, compared to
200e300 Bloom for bovine or porcine gelatin (Karim & Bhat, 2009).
gelatins appeared pearly white while rohu gelatin was shiny
This may be due to many factors such as pH, molecular weight
whitish.
distribution and amino acid content. Hydroxyproline contents of
While shark and rohu gelatins showed good transmittance (%T),
shark, rohu and tuna gelatins were about 9.85%, 6.78% and 7.86%, of
tuna gelatin showed significantly (p < 0.01) poor %T values
the total amino acids (Table 4), while that of bovine gelatin was 14%
(Table 3). The turbidity and dark colour of gelatin is commonly
(Nalinanon, Benjakul, Visessanguan, & Kishimura, 2008). The gel-
caused by inorganic, proteinaceous and mucosubstance contami-
ling properties of gelatin are influenced by the source of raw ma-
nants introduced or not removed during its extraction (Eastoe &
terials, which vary in proline and hydroxyproline contents
Leach, 1977). Muyonga et al. (2004a) stated that the efficiency of
(Jongjareonrak, Benjakul, Visessanguan, Prodpran, et al., 2006). The
the filtration process during gelatin extraction affected the degree
main difference between fish and mammalian gelatins is the imino
of clarity of gelatin solution.
acid content, where the mammalian gelatins have higher amounts
(Gudmundsson, 2002). Imino acids played a role in gel formation.
3.6. Odour of gelatin samples The hydroxyl groups of hydroxyproline play a part in the stability of
the helix by inter-chain hydrogen bonding via a bridging water
Sensory evaluation showed a difference in odour between gela- molecule as well as direct hydrogen bonding to the carbonyl group
tins. The shark and rohu gelatins were found to be free of fishy odour (Wong, 1989). Gelatin, with the shorter chain lengths cannot form
and had a mild putrid odour (hedonic score of 1.83  0.29 and as strong a gel due to the lower inter-junction zones
2.17  0.29, respectively), whereas the tuna gelatin had a detectable (Intarasirisawat et al., 2007). The quality of gelatin is generally
fishy odour with a hedonic score of 3 which was significantly higher determined by the gel strength or Bloom value, including low
(p < 0.01) than shark and rohu gelatins. Muyonga et al. (2004a) re- (<150), medium (150e220) and high Bloom (220e300) (Johnston-
ported that the gelatins prepared from the skins and bones of Nile Bank, 1983). Gudmundsson and Hafsteinsson (1997) suggested that
perch by activated carbon treatment were found to have a mild the gel strength may depend on the isoelectric point and may be
putrid odour. Strong fishy odour was reported for freeze-dried controlled, to certain extent, by adjusting the pH.
gelatin prepared from the skins of black tilapia (Jamilah & Harvinder,
2002). Choi and Regenstein (2000) observed that activated carbon
treatment at the final stage of extraction can further reduce the 3.9. Amino acid composition
odour and improve the acceptability of gelatin.
The amino acid composition of gelatin extracted from the skins
of shark, rohu and tuna are shown in Table 4. It is expressed as g of
3.7. Viscosity values of gelatin samples

Viscosity is the second most important commercial property of Table 4


gelatin after gel strength (Ward & Courts, 1977). The viscosity for Amino acid composition of shark skin gelatin (SSG), rohu skin gelatin (RSG) and tuna
the samples was in the range of 2.5e5.6 cP and all the samples were skin gelatin (TSG).

significantly (p < 0.01) different. The viscosity was lowest in rohu Amino acid Residues/100 residues
gelatin compared to shark and tuna gelatins (Table 3). Viscosity is SSG RSG TSG
partially controlled by molecular weight and molecular size dis-
Aspartic 3.64  0.09 4.39  0.10 4.08  0.17
tribution (Sperling, 1985). The viscosities of most of the commercial
Threonine 2.08  0.19 2.38  0.11 2.33  0.13
gelatins have been reported to be up to 13.0 cP (Johnston-Banks, Glutamic 7.69  0.27 6.30  0.36 7.26  0.23
1990). Jamilah and Harvinder (2002) reported the viscosity values Glycine 32.8  062 33.4  0.74 33.2  0.55
of 3.2 cP and 7.12 cP for red and black tilapia, respectively, whereas Alanine 10.9  0.30 11.7  0.20 12.2  0.36
Cystine 0 0 0
for channel catfish, it was 3.23 cP (Yang et al., 2007). The changes in
Valine 2.52  0.24 3.06  0.25 2.95  0.22
pH are known to influence the viscosity and minimum viscosity for Methionine 1.07  0.15 1.27  0.12 1.24  0.07
Isoleucine 1.58  0.05 0.77  0.03 1.09  0.07
Leucine 2.17  0.17 2.14  0.18 2.04  0.09
Table 3 Tyrosine 0.08  0.03 0.25  0.07 0.26  0.03
Viscosity and Bloom values of shark skin gelatin (SSG), rohu skin gelatin (RSG) and Phenylalanine 1.58  0.10 1.93  0.09 1.37  0.09
tuna skin gelatin (TSG). Lysine 2.77  0.13 2.48  0.25 2.57  0.12
Hydroxylysine 0.73  0.10 0.70  0.08 0.77  0.08
Sample Viscosity Bloom
Histidine 0.86  0.07 0.67  0.12 0.88  0.08
SSG 5.60  0.10a 206  3.51a Arginine 5.29  0.09 5.47  0.13 4.76  0.15
RSG 2.50  0.00c 124  3.61c Proline 9.90  0.11 11.6  0.17 10.1  0.12
TSG 4.37  0.06b 177  3.51b Hydroxyproline 9.85  0.33 6.78  0.17 7.86  0.27
Serine 3.61  0.11 4.11  011 4.36  0.11
Values are given as mean  standard deviation of triplicate. Values with the same
Total 99.1 ± 0.74 99.4 ± 0.63 99.4 ± 0.50
superscript letters within a column are not significantly different (p < 0.01).
74 K. Shyni et al. / Food Hydrocolloids 39 (2014) 68e76

amino acids per 100 g of solids. All gelatins had glycine as their
major amino acid (32.8e33.4) and had relatively high content of
alanine (10.9e12.2). For imino acids, all samples had proline and
hydroxyproline content of 9.90e11.6 and 6.78e9.85, respectively. It
was observed that the level of tryptophan was below the detectable
limit in all three samples. The preparation of amino acid monomers
from gelatin through the acid hydrolysis process leads to the
destruction of indole ring of tryptophan. Hence the detection of
tryptophan is not possible in HPLC. The absence can also be resulted
from acid hydrolysis during the gelatin extraction process
(Chapman & Hall, 1997; Jamilah & Harvinder 2002). Cysteine and
tryptophan are not commonly present in collagen and gelatin
(Foegeding, Lanier, & Hultin, 1996; Jongjareonrak, Benjakul,
Visessanguan, & Tanaka, 2006; Muyonga, Cole, & Duodu, 2004b;
Yata, Yoshida, Fujisawa, Mizuta, & Yoshinaka, 2001). Regenstein
and Zhou (2007) reported that glycine, alanine, proline and hy-
droxyproline are four of the most abundant amino acids in gelatin.
The samples had an imino acid content of 17.96e19.75% of
protein. Kasankala, Xue, Weilong, Hong, and He (2007) reported an
imino acid content of 19.47% for gelatin prepared from the skin of
carp. Imino acid content of Nile perch gelatin was 21.5% (Muyonga
et al., 2004a). Grossman and Bergman (1992) reported an imino
acid content of 17% and 5%, for gelatin from cod and tilapia,
respectively. Mammalian gelatins generally contain 30% imino
acids (Poppe, 1992). Johnston-Banks (1990), reported that the
imino acids, especially hydroxyproline, impart considerable rigidity
to the collagen structure and a relatively limited imino acid content
might affect the dynamic properties of gelatin. Maximum gel
strength was observed for shark skin gelatin which shows that
Plate I. Protein pattern of gelatin gels from of shark skin gelatin (SS), rohu skin gelatin
hydroxyproline is the major determinant of stability due to its
(RS) and tuna skin gelatin (TS) run alongside of the high molecular weight marker
hydrogen bonding ability through its hydroxyl group, although (HM).
proline is also important (Burjandze, 1979; Ledward, 1986).

3.10. Molecular weight distribution of black tilapia (28.9  C) (Jamilah & Harvinder, 2002) which is a
warm-water fish. Fish gelatins with lower melting temperatures
The gelatin preparations from the skins of three species were had a better release of aroma and offered stronger flavour and
analyzed using SDS-PAGE, to characterize differences in molecular useful in product development to control the texture and flavour
weight distribution (Plate I). For gelatin gels a1 and a2 chains were release during mastication (Zhou, Steven, & Regenstein, 2006).
found as the major components for all the samples. However, b
component (covalently linked a chain dimer) with molecular
3.12. Setting point and setting time of gelatin
weight w200 kDa was observed in shark gelatin and it was absent
in the other two samples. The presences of peptides with molecular
Setting point denotes the temperature at which setting/gelling
weight below 100 kDa were slightly noticeable in all three samples.
process begins, which involves the transition from random coil to
That may be the results of heat induced cleavages of protein chains
triple helical structure of gelatin (Haug, Draget, & Smidsrod, 2004).
that occurred during the extraction process (Muyonga et al.,
Setting points of shark and tuna gelatins were significantly higher
2004a). The result suggested that b chains of shark skin gelatin
(p < 0.01) than that of rohu gelatin (Table 5). Bovine and porcine
were more tolerant to thermal hydrolysis than those of tuna and
gelatins have considerably higher setting and melting points than
rohu skin gelatin.
most fish gelatins (Choi & Regenstein, 2000; Gilsenan & Ross-
Murphy, 2000; Gudmundsson, 2002; Leuenberger, 1991). Setting
3.11. Melting point of gelatin
points of cold-water fish gelatin was in general lower than warm-
water fish/mammalian gelatins. This difference is related to the
The melting point of shark and tuna gelatins was significantly
imino acid composition. Setting and melting temperatures are also
higher (p < 0.01) than that of rohu gelatin (Table 5). This can be
influenced by the change in ionic strength and pH of gelatin. They
correlated with the higher hydroxyproline content in these gela-
decreased with the increase in ionic strength of >0.5 mol/L, which
tins. Melting point increases with the maturation time and it has
been observed that the levels of hydroxyproline contribute to the
melting point characteristics (Choi & Regenstein, 2000; Table 5
Gudmundsson, 2002). It is known that fish gelatins have lower Melting point (MP), setting point (SP) and setting time (ST) parameters for shark
skin gelatin (SSG), rohu skin gelatin (RSG) and tuna skin gelatin (TSG).
melting points than mammalian gelatins. Norland (1990) and
Gudmundsson (2002) had observed melting points of 29.7  C and Sample MP ( C) SP ( C) ST (s)
32.3  C, for bovine and porcine gelatin, respectively. The melting SSG 25.8  0.29a 20.8  0.29a 93.3  1.15c
points observed in the present study are far higher than those re- RSG 18.2  0.29c 13.8  0.29c 114  1.53a
ported for cold-water fishes such as cod (13.8  C), hake (14  C) TSG 24.2  0.29b 18.7  0.29b 103  1.15b
(Gomez-Guillen et al., 2002) and hoki (16.6  C) (Mohtar, Perera, & Values are given as mean  standard deviation of triplicate. Values with the same
Quek, 2010). However, these melting points were lower than that superscript letters within a column are not significantly different (p < 0.01).
K. Shyni et al. / Food Hydrocolloids 39 (2014) 68e76 75

was probably due to the reduced electrostatic interaction pre- 4. Conclusion


venting ionic inter-chain bridging and gelation of fish gelatin (Haug
et al., 2004). It may be concluded that there are considerable differences
Setting time required for shark skin gelatin was significantly between yield and functional properties of gelatin from the skins of
(p < 0.01) lower than rohu and tuna gelatins (Table 5). Gelling dog shark, skipjack tuna and rohu. The characteristics of the gela-
temperatures reported for Nile perch bone gelatin were 18.5e tins extracted through this process indicate that they have good
19.0  C and gelling time of 90 s (Muyonga et al., 2004a). Similarly, yield and functional properties. There is, therefore, a potential for
Ninan, Abubacker, et al. (2011) and Ninan, Jose, et al. (2011) exploitation of processing waste for gelatin extraction from these
observed gelling temperatures of 18.5  C and 18.0  C and gelling species. The potential is higher for dog shark skins than the other
times of 106 and 103 s in rohu and common carp skin gelatin which two species because dog shark skins give higher gelatin yield and
differed from the result of the present study. That may be attributed the skin gelatin showed better functional properties.
to the difference in age and size of the fish or because of the dif- The ability to form weak gels may find new applications for fish
ference in extraction conditions. gelatin and it could possibly be used in refrigerated products and in
products where low gelling temperatures are required
(Gudmundsson, 2002).
3.13. Foaming properties

Foaming ability (FA) and foam stability (FS) are another impor- Acknowledgements
tant property of gelatin for commonly used foods such as marsh-
mallows (Zuniga & Aguilera, 2009). Foaming ability and foam The authors are thankful to the Director, Central Institute of
stability of the gelatins are shown in Table 6. The foaming capacity Fisheries Technology (Kochi), Kerala, India for according permission
of shark gelatin was found to be significantly higher (p < 0.01) than to publish the paper. The authors also wish to express their grati-
both tuna and rohu gelatins. Foam formation is generally controlled tude to the Department of Biotechnology, Ministry of Science and
by transportation, penetration and reorganization of protein mol- Technology, Government of India, for the financial support given for
ecules at the airewater interface. A protein must be capable of this research work.
migrating rapidly to the airewater interface, unfolding and rear-
ranging at the interface to show good foaming properties (Halling,
1981). Foam stability depends on the nature of the film and in- References
dicates the extent of protein interaction within the matrix
Ahmad, M., & Benjakul, S. (2011). Characteristics of gelatin from the skin of unicorn
(Mutilangi, Panyam, & Kilara, 1996). Foam stability was significantly leatherjacket (Aluterus monoceros) as influenced by acid pretreatment and
lower (p < 0.01) in rohu gelatin when compared to shark and tuna extraction time. Food Hydrocolloids, 25(3), 25e34.
AOAC. (1995). Official methods of analysis of AOAC international (16th ed.). Wash-
gelatins. Foams with higher concentration of proteins were denser
ington, DC: Association of Official and Analytical Chemists International.
and more stable because of an increase in the thickness of inter- Arnesen, J. A., & Gildberg, A. (2007). Extraction and characterisation of gelatin from
facial films (Zayas, 1997). It has been suggested that reduced foam Atlantic salmon (Salmo salar) skin. Bioresource Technology, 98, 53e57.
formation and stability may be due to aggregation of proteins Avena-Bustillos, R. J., Olsen, C. W., Chiou, B., Yee, E., Bechtel, P. J., & McHugh, T. H.
(2006). Water vapor permeability of mammalian and fish gelatin films. Journal
which interfere with interactions between the protein and water of Food Science, 71, E202eE207.
needed for foam formation (Kinsella, 1977). Baziwane, D., & He, Q. (2003). Gelatin: the paramount food additive. Food Reviews
International, 19(4), 423e435.
Binsi, P. K., Shamasundar, B. A., Dileep, A. O., Badii, F., & Howell, N. K. (2009).
Rheological and functional properties of gelatin from the skin of bigeye snapper
3.14. Water holding capacity and fat binding capacity (Priacanthus hamrur) fish: influence of gelatin on the gel-forming ability of fish
mince. Food Hydrocolloids, 23(1), 132e145.
Water holding and fat binding capacities are functional prop- BSI (British Standards Institution). (1975). Methods for sampling and testing gelatin
(physical and chemical methods). London: BSI.
erties that are closely related to texture by the interaction between Burjandze, T. V. (1979). Hydroxyproline content and location in relation to collagen
components such as water, oil and other components (Cho et al., thermal stability. Biopoly, 18, 931e936.
2004). Water holding capacity and fat binding capacity of gelatins Chapman & Hall. (1997). Thickening and gelling agents for food (2nd ed.) (pp. 150e
153). London: Blackie Academic & Professional.
are shown in Table 6. Water holding capacity is significantly higher Cho, S. M., Gu, Y. S., & Kim, S. B. (2005). Extraction optimization and physical
(p < 0.01) in shark gelatin. Higher water holding capacity is mainly properties of yellowfin tuna (Thunnus albacares) skin gelatine compared to
related to the higher amounts of hydrophilic amino acids and mammalian gelatins. Food Hydrocolloids, 19, 221e229.
Cho, S. M., Kwak, K. S., Park, D. C., Gu, Y. S., Ji, C. L., & Jang, D. H. (2004). Processing
higher hydroxyproline content (Ninan, Abubacker, et al., 2011;
optimization and functional properties of gelatin from shark (Isurus oxyrinchus).
Ninan, Jose, et al., 2011). Fat binding capacity is significantly Food Hydrocolloids, 18, 573e579.
higher (p < 0.01) in tuna gelatin. The degree of exposure of the Choi, S. S., & Regenstein, J. M. (2000). Physicochemical and sensory characteristics of
fish gelatin. Journal of Food Science, 65, 194e199.
hydrophobic residues and the high amount of tyrosine were found
Cole, C. G. B. (2000). Gelatin. In J. F. Frederick (Ed.), Encyclopedia of food science and
responsible for the high fat binding capacity (Ninan, Abubacker, technology (2nd ed.) (pp. 1183e1188). New York: John Wiley & Sons.
et al., 2011; Ninan, Jose, et al., 2011). Eastoe, J. E., & Eastoe, B. (1952). A method for the determination of total nitrogen in
proteins. The British Gelatine and Glue Research Association Research Report, Se-
ries B, 5, 1e17.
Eastoe, J. E., & Leach, A. A. (1977). The chemical constitution of gelatin. In A. G. Ward,
Table 6
& A. Courts (Eds.), The science and technology of gelatin (pp. 73e107). London:
Foaming ability (FA), foam stability (FS), water holding capacity (WHC) and fat
Academic Press.
binding capacity (FBC) of shark skin gelatin (SSG), rohu skin gelatin (RSG) and tuna Fernandez-Diaz, M. D., Montero, P., & Gomez-Guillen, M. C. (2001). Gel properties of
skin gelatin (TSG). collagens from skins of cod (Gadus morhua) and hake (Merluccius merluccius)
and their modification by the coenhancers magnesium sulphate, glycerol and
Sample FA (%) FS (%) WHC FBC
transglutaminase. Food Chemistry, 74, 161e167.
SSG 21.5  0.45a 17.6  0.35a 256  4.51a 347  5.51b Foegeding, E., Lanier, T. C., & Hultin, H. O. (1996). Characteristics of edible muscle
RSG 17.4  0.15c 10.5  0.25c 163  4.00c 360  29.19b tissue. In O. R. Fennema (Ed.), Food chemistry (pp. 879e942). New York: Marcel
TSG 19.2  0.36b 14.4  0.25b 214  3.61b 452  6.51a Dekker.
Gilsenan, P. M., & Ross-Murphy, S. B. (2000). Viscoelasticity of thermoreversible
Values are given as mean  standard deviation of triplicate. Values with the same gelatin gels from mammalian and piscine collagens. Journal of Rheology, 44(4),
superscript letters within a column are not significantly different (p < 0.01). 871e883.
76 K. Shyni et al. / Food Hydrocolloids 39 (2014) 68e76

GME. (2005). Standardised methods for the testing of edible gelatine. In Gelatine Mutilangi, W. A. M., Panyam, D., & Kilara, A. (1996). Functional properties of hy-
monograph (p. 104). 4, Avenue E. Van Nieuwenhuyse, B e 1160 Bruxelles: drolysates from proteolysis of heat-denatured whey protein isolate. Journal of
Gelatin Manufacturers of Europe. Version e Jun 2005. Food Science, 61(2), 270e274.
Gomez-Guillen, M. C., & Montero, P. (2001). Method for the production of gelatin of Muyonga, J. H., Cole, C. G. B., & Duodu, K. G. (2004a). Extraction and physico-
marine origin and product thus obtained. International Patent PCT/S01/00275. chemical characterization of nile perch (Lates niloticus) skin and bone gelatin.
Gomez-Guillen, M. C., Perez-Mateos, M., Gomez-Estaca, J., Lopez-Caballero, E., Food Hydrocolloids, 18, 581e592.
Gimenez, B., & Montero, P. (2009). Fish gelatin: a renewable material for Muyonga, J. H., Cole, C. G. B., & Duodu, K. G. (2004b). Characterisation of acid soluble
developing active biodegradable films. Trends in Food Science & Technology, collagen from skins of young and adult nile perch (Lates niloticus). Food
20(1), 3e16. Chemistry, 85(1), 81e89.
Gomez-Guillen, M. C., Turnay, J., Fernandez-Diaz, M. D., Ulmo, N., Lizarbe, M. A., & Nair, V. (2002). Biochemical composition of fish. In K. Gopakumar (Ed.), Text book of
Montero, P. (2002). Structural and physical properties of gelatin extracted from fish processing technology (p. 22). New Delhi, India: ICAR.
different marine species: a comparative study. Food Hydrocolloids, 16, 25e34. Nalinanon, S., Benjakul, S., Visessanguan, W., & Kishimura, H. (2008). Tuna pepsin:
Grossman, S., & Bergman, M. (1992). Process for the production of gelatin from fish characteristics and its use for collagen extraction from the skin of threadfin
skin. United States Patent, 5,093,474. bream (Nemipterus spp.). Journal of Food Science, 73(5), 413e419.
Gudmundsson. (2002). Rheological properties of fish gelatin. Journal of Food Science, Ninan, G., Abubacker, Z., & Jose, J. (2011). Physico-chemical and texture properties of
67, 2172e2176. gelatins and water gel desserts prepared from the skin of freshwater carps. Fish
Gudmundsson, M., & Hafsteinsson, H. (1997). Gelatin from cod skins as affected by Technology, 48(1), 67e74.
chemical treatments. Journal of Food Science, 62, 37e47. Ninan, G., Jose, J., & Abubacker, Z. (2011). Preparation and characterization of gelatin
Halling, P. J. (1981). Protein stabilized foams and emulsions. CRC Critical Reviews in extracted from the skins of rohu (Labeo rohita) and common carp (Cyprinus
Food Science and Nutrition, 12, 155e203. carpio). Journal of Food Processing and Preservation, 35(2), 143e162.
Hao, S., Li, L., Yang, X., Cen, J., Shi, H., Bo, Q., et al. (2009). The characteristics of Norland, R. E. (1990). Advances in fisheries and biotechnology for increased profitability.
gelatin extracted from sturgeon (Acipenser baeri) skin using various pre- In M. N. Voigt, & J. K. Botta (Eds.), Fish gelatin (pp. 325e333). Lancaster: Technomic.
treatments. Food Chemistry, 115, 124e128. Norziah, M. H., Al-Hassan, A., Khairulnizam, A. B., Mordi, M. M., & Norita, M. (2009).
Haug, I. J., Draget, K. I., & Smidsrod, O. (2004). Physical and rheological properties of Characterization of fish gelatin from surimi processing wastes: thermal analysis and
fish gelatin compared to mammalian gelatin. Food Hydrocolloids, 18, 203e213. effect of transglutaminase on gel properties. Food Hydrocolloids, 23(6), 1610e1616.
Holzer, D. (1996). Gelatin production. United States Patent, 484,888. Ockerman, H. W., & Hansen, C. L. (1999). Glue and gelatine. In Animal by-product
Intarasirisawat, R., Benjakul, S., Visessanguan, W., Prodpran, T., Tanaka, M., & processing and utilization (pp. 183e216). Baca Raton, FL: CRC Press.
Howell, N. K. (2007). Autolysis study of bigeye snapper (Priacanthus macra- Poppe, J. (1992). Gelatin. In A. Imeson (Ed.), Thickening and gelling agents for food
canthus) skin and its effect on gelatin. Food Hydrocolloids, 21, 537e544. (pp. 98e123). Glasgow, UK: Blackie Academic & Professional.
Ishida, Y., Fugita, T., & Asai, K. (1981). Journal of Chromatography, 204, 143e148. Regenstein, J. M., & Zhou, P. (2007). Collagen and gelatin from marine by-products.
Jamilah, B., & Harvinder, K. G. (2002). Properties of gelatins from skins of fish-black In F. Shahidi (Ed.), Maximising the value of marine by-products (pp. 279e303).
tilapia (Oreochromis mossambicus) and red tilapia (Oreochromis nilotica). Food Cambridge: Woodhead Publishing Limited.
Chemistry, 77, 81e84. Sadowska, M., Kolodziejska, I., & Niecikowska, C. (2003). Isolation of collagen from
Jamilah, B., Tan, K. W., Umi Hartina, M. R., & Azizah, A. (2011). Gelatins from three the skins of Baltic cod (Gadus morhua). Food Chemistry, 81, 257e262.
cultured freshwater fish skins obtained by liming process. Food Hydrocolloids, Sastry, C. S. P., & Tammuru, M. K. (1985). Spectrophotometric determination of
25, 1256e1260. tryptophan in protein. Journal of Food Science and Technology, 22, 146e147.
Johnston-Bank, F. A. (1983). From tannery to table: an account of gelatin production. Spend Matters. (2012). http://spendmatters.com/2012/11/12/gelatin-rising-prices-
Journal of the Society of Leather Technologists and Chemists, 68, 141e145. growing-asiansouth-american-demand/ Accessed on 29.04.13.
Johnston-Banks, F. A. (1990). Gelatin. In P. Harris (Ed.), Food gels (pp. 233e289). Sperling, L. H. (1985). Introduction of physical polymer science. New York: John Wiley
London: Elsevier Applied Science Publishers. & Sons.
Jones, N. R. (1977). Uses of gelatin in edible products. In A. G. Ward, & A. Courts (Eds.), Stainsby, G. (1987). Gelatin gels. In A. M. Pearson, T. R. Dutson, & A. J. Bailey (Eds.),
The science and technology of gelatin (pp. 366e395). New York: Academic Press. Advances in meat research: Vol. 4. Collagen as food (pp. 209e222). New York: Van
Jongjareonrak, A., Benjakul, S., Visessanguan, W., Prodpran, T., & Tanaka, M. (2006). Nostrand Reinhold Company, Inc.
Characterization of edible films from skin gelatin of brownstripe red snapper Wainewright, F. W. (1977). Physical tests for gelatin and gelatin products. In
and bigeye snapper. Food Hydrocolloids, 20, 492e501. A. G. Ward, & A. Courts (Eds.), The science and technology of gelatin (pp. 507e
Jongjareonrak, A., Benjakul, S., Visessanguan, W., & Tanaka, M. (2006). Fatty acids 534). London, UK: Academic Press.
and their sucrose esters affect the properties of fish skin gelatin based film. Ward, A. G., & Courts, A. (1977). The science and technology of gelatin. London: Ac-
European Food Research and Technology, 222, 650e657. ademic Press Inc.
Karim, A. A., & Bhat, R. (2009). Fish gelatin: properties, challenges and prospects as Wong, D. W. S. (1989). Mechanism and theory in food chemistry. New York: Van
an alternative to mammalian gelatins. Food Hydrocolloids, 23, 563e576. Nostrand Reinhold Company Inc.
Kasankala, L. M., Xue, Y., Weilong, Y., Hong, S. D., & He, Q. (2007). Optimization of Yang, H., Wang, Y., Jiang, M., Oh, J. H., Herring, J., & Zhou, P. (2007). 2-Step opti-
gelatin extraction from grass carp (Ctenopharyngodon idella) fish skin by mization of the extraction and subsequent physical properties of channel cat-
response surface methodology. Bioresource Technology, 98(17), 338e343. fish (Ictalurus punctatus) skin gelatin. Journal of Food Science, 72(4), 188e195.
Kinsella, J. E. (1977). Functional properties of protein: possible relationships be- Yata, M., Yoshida, C., Fujisawa, S., Mizuta, S., & Yoshinaka, R. (2001). Identification
tween structure and function in foams. Journal Food Chemistry, 7, 273e288. and characterization of molecular species of collagen in fish skin. Journal of Food
Laemmli, U. K. (1970). Cleavage of structural proteins during the assembly of the Science, 66(2), 247e251.
head of bacteriophage T4. Nature, 227, 680e685. Zayas, J. F. (1997). Functionality of proteins in food (pp. 6e22). Berlin: Springer-
Ledward, D. A. (1986). Gelation of gelatin. In J. R. Mitchel, & D. A. Ledward (Eds.), Verlag.
Functional properties of food macromolecules (pp. 171e201). Elsevier Applied Zhou, P., & Regenstein, J. M. (2005). Effects of alkaline and acid pretreatments on
Science Publishers Ltd. Alaska pollock skin gelatin extraction. Journal of Food Science, 70, 392e396.
Leuenberger, B. H. (1991). Investigation of viscosity and gelation properties of Zhou, P., Steven, J. M., & Regenstein, J. M. (2006). Properties of Alaska pollock skin
different mammalian and fish gelatins. Food Hydrocolloids, 5, 353e361. gelatin: a comparison with tilapia and pork skin gelatins. Journal of Food Science,
Mohtar, N. F., Perera, C., & Quek, S.-Y. (2010). Optimisation of gelatine extraction 71(6), 313e321.
from hoki (Macruronus novaezelandiae) skins and measurement of gel strength Zuniga, R. N., & Aguilera, J. M. (2009). Structureefracture relationships in gas-filled
and SDSePAGE. Food Chemistry, 122, 307e313. gelatin gels. Food Hydrocolloids, 23(5), 1351e1357.

You might also like