You are on page 1of 9

Meat Science 121 (2016) 333–341

Contents lists available at ScienceDirect

Meat Science

journal homepage: www.elsevier.com/locate/meatsci

Physicochemical and structural properties of composite gels prepared


with myofibrillar protein and lard diacylglycerols
Xiaoqin Diao a,b, Haining Guan a,b, Xinxin Zhao a, Xinping Diao c, Baohua Kong a,⁎
a
College of Food Science, Northeast Agricultural University, Harbin, Heilongjiang 15000, China
b
College of Food and Pharmaceutical Engineering, Suihua University, Suihua, Heilongjiang 152061, China
c
College of Animal Science, Northeast Agricultural University, Harbin, Heilongjiang 15000, China

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

Article history: The objective of this study was to investigate the physicochemical and structural properties of composite gels
Received 28 April 2016 prepared with porcine myofibrillar protein (MP) and lard, glycerolized lard (GL) or purified glycerolized lard
Received in revised form 2 July 2016 (PGL). The gels prepared with MP and GL or PGL had significantly higher penetration force and water-holding ca-
Accepted 7 July 2016
pacity (WHC) than the gel with lard (P b 0.05) and formed a more compact and orderly microstructure. Com-
Available online 08 July 2016
pared with the distributions of T2 relaxation times of the pure MP gel, T21 and T22 of the gels that were
Keywords:
prepared with GL or PGL moved in the direction of slower relaxation time, which suggests that the water mobility
Lard diacylglycerols in the gel system was restricted. The presence of lard, GL and PGL did not affect the participating proteins in com-
Myofibrillar protein posite gels. The presence of GL and PGL altered the secondary and tertiary structures of MP in composite gels,
Gel properties which changed the gel properties. In general, the composite gels that were prepared with MP and GL or PGL
Structural changes showed improved gel quality.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction (2011) reported that DAGs could improve the food texture and water
retention. In addition, DAGs have been reported to significantly sup-
Myofibrillar protein (MP) has notably important biological func- press abdominal and visceral fat accumulation and reduce body weight
tions, and its gel-forming ability plays a key role in processed meat (Maki et al., 2002; Meng, Zou, Shi, Duan, & Mao, 2004). Meanwhile, the
products and significantly affects the texture and sensory characteristics safety of DAGs has been confirmed through several animal and human
of the final products (Sun & Holley, 2011). Ziegler and Acton (1984) studies (Morita & Soni, 2009). Therefore, DAGs may totally or partially
noted the MP gel formation is responsible for the formation of a three- replace animal fat in the processing of meat products. Some reports
dimensional gel matrix because of the association of protein during show that lard-based diacylglycerols can be applied as a fat replacer in
heat processing. The formed protein gels are notably important for meat emulsions and fermented sausages (Miklos et al., 2011, 2014;
their contribution to meat binding, fat immobilization and water en- Mora-Gallego et al., 2013). Our previous study investigated the emulsi-
trapment in meat products (Wu, Xiong, Chen, Tang, & Zhou, 2009). In fying properties and oxidative stability of emulsions of MP and different
addition, Mendoza, García, Casas, and Selgas (2001) have reported lipids (lard, GL and PGL). The results revealed that lard diacylglycerols
that animal fats play an important role in providing good mouthfeel enhanced the emulsifying abilities and had no adverse effects on the ox-
and juiciness in meat products and are stabilized by a proteinaceous idation stability of the emulsions that were prepared with MP (Diao,
membrane (Gordon & Barbut, 1992). However, overconsumption of Guan, Zhao, Chen, & Kong, 2016).
fat causes health problems such as obesity, hypertension and cardiovas- The interaction between fat and the MP gel matrix plays a determi-
cular heart diseases (Nejat, Polotsky, & Pal, 2010). To improve human nant role in the stability of cooked meat products. Xiong and Kinsella
health, low-fat meat foods have been developed in the meat industry (1991) revealed that fat globules of various sizes and concentrations
without compromising on the texture and mouthfeel. could reinforce the milk protein-based gel matrix. Wu et al. (2009) re-
Triacylglycerols (TAGs) are the main component of pork fat. Previ- ported that lipid types and concentrations could alter the rheological
ous studies have found that TAGs can be converted into diacylglycerols and microstructural properties of MP-lipid composite gels. However,
(DAGs) through the enzymatic glycerolysis of fat with glycerol (Cheong, to our knowledge, no literature has been reported about the interaction
Zhang, Xu, & Xu, 2009; Miklos, Xu, & Lametsch, 2011). Miklos et al. between myofibrillar proteins and DAGs in composite-gel formation
during the heating process. Therefore, in this study, we attempt to de-
⁎ Corresponding author. scribe the effect of lard DAGs on the gel strength, water-holding capac-
E-mail address: kongbh63@hotmail.com (B. Kong). ity, gel-forming proteins and microstructures of composite gels that

http://dx.doi.org/10.1016/j.meatsci.2016.07.002
0309-1740/© 2016 Elsevier Ltd. All rights reserved.
334 X. Diao et al. / Meat Science 121 (2016) 333–341

were prepared with MP and lard DAGs. The secondary and tertiary investigate the secondary and tertiary structural changes of MP in com-
structures of MP, mobility of water molecules and molecular forces in posite gels using Fourier transform infrared spectra and intrinsic fluo-
composite gels were elucidated. rescence measurement.

2. Materials and methods 2.5. Gel strength

2.1. Materials The MP and fat composite gels were penetrated with a flat-surface
cylindrical probe (P/0.5, 12 mm in diameter), which was attached to a
Fresh pork back fat and pork loin muscle were obtained from the Model TA-XT2 texture analyser (Stable Micro Systems Ltd., England,
Beidahuang Meat Corporation (Harbin, Heilongjiang, China). Sodium U.K.) at a test speed of 1 mm/s over a 10 mm displacement. The required
dodecyl sulfate (SDS), piperazine-1 and 4 bisethanesulfonic acid penetration force (N) to rupture the gels was expressed as the gel
(PIPES) were purchased from Sigma Chemical Co. (St. Louis, MO, USA). strength (Xiong & Brekke, 1991).
All other chemicals and reagents in this work were obtained from com-
mercial sources and were of analytical grade. 2.6. Gel water-holding capacity

2.2. Preparation of lard diacylglycerols (DAGs) The water-holding capacity (WHC) of the gels was determined using
a centrifugal method. Briefly, the gel samples (5 g) were placed into a
Lard DAGs were prepared according to Diao et al. (2016). Lard was centrifuge tube and centrifuged at 10,000 ×g for 15 min at 4 °C. The sur-
extracted by heating the backfat at 120 °C. The reaction mixture, face of the gels was soaked using filter paper. WHC (%) was expressed as
which included glycerol, melted lard and Lipozyme RMIM, was used the ratio between the gel weights after centrifugation and before centri-
in the glycerolysis reaction under the following conditions: 1:1 M fugation, which was multiplied by 100. The supernatant was collected
ratio of lard to glycerol, 14:100 (W/W) of enzyme-to-lard substrate and used to analyse the gel-forming proteins in section 2.8.
ratio and 500 rpm magnetic stirring speed. First, the reaction mixture
was incubated at 65 °C for 2 h and transferred to 45 °C for 8 h. After 2.7. Low-field nuclear magnetic resonance analysis
the reaction, the enzyme was filtered to yield the glycerolized lard
(GL), whose DAG content was 61.8%. The GL was purified using the The nuclear magnetic resonance relaxation of the gel samples was
two-step wiped film molecular distillation (SPE10, manufactured in measured using an LF-NMR analyser minispec mq 20 (Bruker Optik
Haiyuan biochemical equipment Co. Ltd., Wuxi, China). The purified GmbH, Germany) to determine the mobility and proportion of different
glycerolized lard (PGL), which had a higher DAG content (82.0%), was fractions of water molecules in the gel system without destroying the
obtained in the second purification step. gel structure. Approximately 2 g of composite gel was placed in an
NMR glass tube (1.8 cm in diameter and 18 cm in height). The analyser
2.3. Preparation of myofibrillar protein (MP) was operated at a magnetic field strength of 0.47 T and a proton reso-
nance frequency of 20 MHz. The transverse relaxation time (T2) was
MP was extracted from the pork loin muscle according to the de- measured using the Carr-Purcell-Meiboom-Gill pulse sequence
scribed procedure (Xia, Kong, Liu, & Liu, 2009). The final protein concen- (CPMG). For each sample, 16 scans were obtained at a 2 s interval
tration was measured using the Biuret method with bovine serum with 3000 echoes in total. The relaxation data were treated using the
albumin (Sigma Chemical Co., St. Louis, MO) as a standard. The MP CONTIN software that was provided with the equipment which resulted
was maintained at 2–4 °C and used within 48 h. in the corresponding distributions of relaxation times from the decay
curve. The mean apparent relaxation time (T2i) and amplitude (A2i)
2.4. Preparation of MP and fat composite gel for each detected population in CONTIN were recorded.

Predetermined amounts of melted fats (lard, GL and PGL at 45 °C) 2.8. Identification of gel-forming proteins
were separately mixed with a myofibril solution (1% protein in 0.6 M
NaCl, 50 mM PIPES, pH 6.0). These mixtures were placed in a 35 °C The solutions that gathered from the WHC measurement were used
water bath for 5 min to guarantee that the fats maintained liquid and to perform the sodium-dodecyl sulfate-polyacrylamide gel electropho-
were subsequently homogenized at 10,000 rpm for 1 min with an IKA resis (SDS-PAGE) to monitor the composition of uncoagulated proteins
T18 Ultra-Turrax (IKA-Werke GmbH & Co., Staufen, Germany). The in the complex gels according to the method of Laemmli (1970). For
pre-emulsified fats were immediately used after preparation. SDS-PAGE, a resolving gel of 12% acrylamide and a stacking gel of 5% ac-
A predetermined amount of MP was dissolved in 50 mM PIPES rylamide were used. The amount of loaded samples per lane was 12 μL.
(pH 6.0), which contained 0.6 M NaCl, to form a protein solution. The following proteins were used as the molecular weight standards:
Then, each specific amount of pre-emulsified fat was added into the myosin (200.0 kDa), β-galactosidase (116.0 kDa), phosphorylase b
protein solution by gently stirring with a glass rod to produce compos- (97.2 kDa), serum albumin (66.4 kDa), ovalbumin (44.3 kDa), carbonic
ites with fat contents of 4%, 8% and 12% (w/w). Simultaneously, the total anhydrase (29.0 kDa) and trypsin inhibitor (20.1 kDa).
amount of MP was maintained constant (4%). Then, aliquots of 15 g of
MP and lard, GL and PGL composites were separately poured into 2.9. Molecular forces in composite gels
25 mm (inner diameter) × 40 mm (length) glass vials and covered
with aluminum foil. These composites were stored at 2–4 °C for one The major molecular forces that were involved in the composite gels
night to reach maximum protein solubility (Ramírez-Suárez, Xiong, & were evaluated according to the method of Jiang and Xiong (2013). Dif-
Wang, 2001), subsequently equilibrated at room temperature (25 ± ferent dissolving solutions were used: 8 M urea + 50 mM sodium phos-
1 °C) for 30 min and heated in a water bath at 72 °C for 10 min. After phate (pH 7.0) to analyse the hydrogen bond; 0.5% (w/v) SDS + 50 mM
heating, the formed gels were cooled and stored in crushed ice for 2 h sodium phosphate (pH 7.0) to analyse the total non-covalent forces;
prior to further analysis. Some prepared gel samples were allowed to 0.25% (v/v) β-mercaptoethanol +50 mM sodium phosphate (pH 7.0)
equilibrate at ambient temperature (approximately 23 °C) for 30 min to analyse the disulfide bands. Samples (1 g) of composite gels were ho-
to measure these gel properties: gel strength, water-holding capacity, mogenized in 9 mL of various solvents at a speed of 13,500 rpm for 20 s.
molecular forces, gel-forming proteins, low-field NMR analysis and gel The homogenates were heated at 80 °C for 1 h to dissolve the gel-
microstructure. Other aliquots of gel samples were lyophilized to forming protein, chilled to room temperature and subsequently
X. Diao et al. / Meat Science 121 (2016) 333–341 335

centrifuged at 10,000 ×g for 15 min. The protein content in the superna- in gels were analysed using a mixed model. In this model, each triplicate
tant was determined using the biuret method. The relevant molecular was included as a random term, and the fat contents (0%, 4%, 8% and
forces in the protein gels can be expressed based on the gel solubility 12%) and fat types (lard, GL or PGL) were included as fixed terms. The
after different dissolving buffer treatments. The gel solubility was calcu- data were analysed using the General Linear Models procedure of the
lated using the relative protein content in the supernatants in compar- Statistix 8.1 software package (Analytical Software, St. Paul, MN, USA).
ison to that in the original suspension, which was taken to prepare the
protein gel. 3. Results and discussion

2.10. Microstructure of composite gels 3.1. Gel strength and water-holding capacity

The microstructure of the composite gels was examined using a The gel strength is an important quality property of muscle protein,
scanning electron microscope (SEM). The gel samples which is notably related to the texture and sensory quality of muscle
(5 × 5 × 5 mm3) were fixed with 2.5% (v/v) glutaraldehyde in 0.1 M food. To elucidate the effect of lard DAGs on MP gel properties, gels
phosphate buffer (pH 7.2) at 4 °C overnight and subsequently washed that were prepared with different fat contents were assessed using a
three times with the above phosphate to remove the glutaraldehyde penetration test. As displayed in Fig. 1A, the addition of lard, GL and
fluid. The composite gels were sequentially dehydrated in ethanol PGL significantly increased the gel strength (P b 0.05), and the extent
with a series of concentrations of 50, 70, 80, 90 and 100% (v/v) for of increase was proportional to the fat addition. MP gels with GL and
10 min each. To remove the ethanol, the dehydrated samples were suc- PGL tended to be more rigid than those with lard, and the difference
cessively immersed in tertiary butanol-ethanol (100%) (1:1) and tertia- was significant (P b 0.05). The gel strength was improved because the
ry butanol for 15 min and freeze-dried. The dried samples were fat globules might behave as “fillers” in the voids of the gels to make
mounted on a bronze stub and sputter-coated with gold (Sputter coater the gels have a more compact structure. As proven in our previous find-
SPI-Module, West Chester, PA, USA). Then, the specimens were ob- ing, PGL and GL have smaller fat globules (Diao et al., 2016), which can
served and photographed with an SEM (S-3400N, Hitachi, Tokyo, adequately fill the gel network structure. Increased the gel strength by
Japan) at an accelerating voltage of 5 kV. addition of lard DAGs can improve the texture and sensory quality of
cooked meat products, so lard DAGs have a good application prospect
2.11. Fourier transforms infrared spectra in the meat products.
A well-structured protein gel can entrap and immobilize a large
The Fourier transform infrared spectra (FT-IR) of the freeze-dried amount of water, fat and other food components through the capillary
samples were recorded using a Perkin-Elmer Spectrum 100 (Perkin- effects of its matrices. WHC is one of the most important functional
Elmer Corp., Norwalk, CT) at 4 cm−1 resolution in the wave number
range of 4000 to 400 cm−1 with KBr pellets. The peak fitting procedure
was used to quantitatively analyse the secondary structural compo-
nents of MP. The characteristic peak of the amide I band was analysed
in the region of 1700–1600 cm− 1 using the Peak Fit v 4.12 software.
In the study, a baseline was first corrected to accurately measure the
band areas of the second-derivative spectra in amide I. Then, the Fourier
self-deconvolution, which affects the number, position and intensity of
the bands, was further performed using a Gaussian curve fit (GCF). Fi-
nally, the GCF was adjusted to give the best least-squares fit of the indi-
vidual bands to each deconvoluted spectrum. The relative amounts of
α-helix, β-sheet, β-turn and random coil structures of MP in the com-
posite gels were determined from the second derivative of amide I by
manually computing the areas under the bands that were assigned to
a particular substructure.

2.12. Intrinsic fluorescence measurement

The intrinsic fluorescence emission spectra of different samples


were determined using an F-4500 fluorescence spectrofluorometer
(Tokyo, Japan). The concentration of MP in the composite gels was
fixed at 0.2 mg/mL in 10 mM phosphate buffer (pH 7.0). The emission
spectra of tryptophan were recorded from 300 to 400 nm with an exci-
tation wavelength of 295 nm. The excitation and emission slit widths
were set at 5 nm. A solution of 10 mM phosphate buffer (pH 7.0) was
used as negative control.

2.13. Statistical analysis

Three batches of composite gels were independently prepared to


study the effects of MP and different pork fats (lard, GL or PGL) on the
physicochemical and structural properties of the gels. For each batch
of gel samples, all specific experiments were conducted in triplicate
(triplicate observations). The data are presented as the mean ± stan-
Fig. 1. Penetration force (A) and water-holding capacity (B) of composite gels that were
dard errors. The analysis of variance (ANOVA) with Tukey's multiple prepared with myofibrillar proteins (MP) and lard, glycerolized lard (GL) or purified
comparison was used to measure the significance of the main effects glycerolized lard (PGL). The lowercase letters (a–c) indicate significant differences
(P b 0.05). Data regarding physicochemical and structural properties (P b 0.05) among different types of fat at identical fat contents.
336 X. Diao et al. / Meat Science 121 (2016) 333–341

properties of protein gel. The WHC of the gels that were prepared with that were prepared with MP and lard, GL or PGL. Three relaxation com-
lard, GL or PGL increased with the increase in fat concentration (Fig. 1B). ponents were detected in all gels: a minor component at 1–5 ms (T2b), a
The gel that was prepared with lard had lower WHC than those pre- main component at 6–100 ms (T21), and a component at 100–600 ms
pared with GL or PGL (P b 0.05). The gels that were prepared with GL (T22). Compared with the distributions of T2 relaxation times of the
and PGL retained more water possibly because the hydrophilic polar pure MP gel, T21 and T22 of the gels that were prepared with MP and
group in the molecular structure of GL and PGL enhanced the interaction lard, GL or PGL moved in the direction of slower relaxation time, and
between water and protein (Nakajima, 2004). Xu, Han, Fei, and Zhou T21 (major component) became wider with a decrease in intensity,
(2011) and Sun, Wu, Xu, and Li (2012) revealed that the interactions be- which suggests that the water mobility in the gel system was restricted.
tween proteins and water could bind more water molecules in the cap- Noronha, Duggan, Ziegler, O'Riordan, and O'Sullivan (2008) showed
illaries of the insoluble protein matrix. Wu et al. (2009) also reported that shorter relaxation time corresponded to less mobile water fraction.
that those gels prepared with fat and oil could have improved WHC T21 of the gels that were prepared with MP and GL or PGL had broader
due to their compact structure because of the space-filling effect of the distribution probably because of a greater variation in the chemical
lipid droplets. WHC is generally used to objectively evaluate the quality and physical properties of the gels (Salomonsen, Sejersen, Viereck,
of meat products (Rosenvold & Andersen, 2003). A higher amount of Ipsen, & Engelsen, 2007).
water in cooked meat products can improve their juiciness. The change in T21 relaxation times of the gels that were prepared
with different fat contents is shown in Fig. 2D, and the corresponding
3.2. Low-field NMR analysis area of these T21 distributions is shown in Fig. 2E. The relaxation time
T21 of the gels that were prepared with different fats decreased with in-
The low-field pulsed NMR T2 relaxation times are sensitive to the creased fat contents (P b 0.05). The observation is consistent with the
mobility of water molecules in the gel system (Zhang, Yang, Tang, result of Andersen, Frøst, and Viereck (2010), who noted that higher
Chen, & You, 2015). In the gel system, T2b, T21 and T22 are suggested as fat content in cream cheeses decreased the relaxation time. There is
the relaxation components. The T2b component represents bound no significant difference for composite gels at 8% and 12% fat levels
water that is closely associated with macromolecules, the T21 compo- (P N 0.05). The gel prepared with PGL had significantly lower relaxation
nent reflects trapped water in the gel structure, i.e., the immobilized time T21 (P b 0.05) than those prepared with GL and lard, which indi-
water, and the T22 component corresponds to free water outside the cates that PGL significantly increased the ability of protein-binding hy-
gel structure (Shaarani, Nott, & Hall, 2006). Fig. 2A, B and C show the drogen protons. The result can be attributed to the development of a
measured distributions of T2 relaxation times of the composite gels three-dimensional network structure, which holds water in a less

Fig. 2. Distributions of LF NMR T2 relaxation times of composite gels that were prepared with myofibrillar proteins (MP) and lard (A), glycerolized lard (GL) (B) or purified glycerolized lard
(PGL) (C) and relaxation time T21 (D) and the corresponding peak area A21 (E). The lowercase letters (a-c) indicate significant differences (P b 0.05) among different types of fat at identical
fat contents.
X. Diao et al. / Meat Science 121 (2016) 333–341 337

mobilized state (Ishioroshi, Samejima, & Yasui, 1979). The correspond- presence of disulfide interactions, hydrogen bonds and hydrophobic
ing areas A21 of composite gels increased with increasing fat content contacts in MP gel. Jiang and Xiong (2013) suggested that with the ad-
(P b 0.05), and composite gels with PGL had significantly higher A21 dition of certain chemical reagents into the protein gel, the constituent
than that with lard at 4%, 8% and 12% fat levels and that with GL at 8% protein that contributed to gel formation would dissolve. When MP
and 12% fat levels (P b 0.05). This result indicates an increase of water and fat are mixed, the balance of these forces in the gels is destroyed.
content in the intramyofibrillar space and suggests that the higher Therefore, the protein solubility of composite gels that were prepared
PGL can combine more water in the gel matrix. with MP and lard, GL or PGL after the treatments with different force-
The T2b components were almost invisible in the MP gel without fat disruption chemicals was measured to clarify possible physicochemical
and appeared in the gel prepared with different types of fats (Fig. 2A, B bonds and interactions that stabilized the composite gels. The changes
and C), particularly that with higher fat concentrations, which suggests of three molecular forces in MP gels with lard, GL or PGL are described
that more water molecules were constrained in the gel structure as in Fig. 4. The gels that were prepared with different fats were solubilized
immobilized water because of the fat addition. The T22 component of in 8 M urea, which destroys intermolecular hydrogen bonds, and the
the gels that were prepared with different types of fat decreased with protein solubility significantly increased compared to the gel without
increased fat contents, particularly in PGL samples (P b 0.05). The gels fat (P b 0.05). Moreover, Fig. 4A shows that the protein solubility in-
without fat had obviously higher T22 intensity than the other groups creased when the fat additive amount increased from 4% to 12%
(P b 0.05), which shows that there was more free water in the gels with- (P b 0.05), and the gel with GL or PGL had significantly higher protein
out fat. The gel with fats, particularly PGL, obviously reduced the inten- solubility than the gel with lard (P b 0.05) possibly because GL and
sity of T22 (P b 0.05), which shows that the PGL addition can trap more PGL have more hydrophilic polar groups in their molecular structure.
free water in the gel matrix and transform it into combined or However, no significant difference (P N 0.05) was observed between
immobilized water. the gels with identical GL and PGL contents. The result is consistent

3.3. Identification of gel-forming proteins

Weakly bound and noncontributing proteins in the gel matrix can be


easily removed by centrifugation (Jiang & Xiong, 2013). SDS-PAGE of
the supernatants of centrifuged composite gels was performed to ob-
serve the proteins that remained extractable (Fig. 3). The myofibrillar
protein band mainly contained myosin heavy chains (MHC), actin, tro-
ponin-T and tropomyosin. Notably, MHC and actin were undetectable
in the supernatants of all gels, which indicated that they were a key
component of the heat-induced protein gel network, and the presence
of lard, GL and PGL did not affect the participating proteins in composite
gels. In addition, compared with the gel without fat, the electrophoretic
patterns of all gels with fat were consistent, which shows that troponin
T and tropomyosin were present in the supernatant after centrifugation,
and they were not the active gel-building substance. This result is simi-
lar to that observed by Wu et al. (2009). Xiong and Brekke (1991) also
reported that troponin T and tropomyosin could not coagulate and did
not contribute to the muscle protein gel network formation.

3.4. Molecular forces in composite gels

Lefevre, Fauconneau, Ouali, and Culioli (2002) attributed the forma-


tion of a three-dimensional network of protein gel to the equilibrium of
intermolecular forces. Wu, Xiong, and Chen (2011) reported the

Fig. 4. Solubility of composite gels that were prepared with myofibrillar proteins (MP) and
Fig. 3. SDS-PAGE of the supernatants of centrifuged (10,000 ×g) composite gels that were lard, glycerolized lard (GL) or purified glycerolized lard (PGL) in 50 mM phosphate buffer
prepared with myofibrillar proteins (MP) and lard, glycerolized lard (GL) or purified (pH 7.0), which contained different chemicals: A, 8 M urea to analyse the intermolecular
glycerolized lard (PGL). MW: protein standard with the indicated marker protein hydrogen bonds; B, 0.5% sodium dodecyl sulfate (SDS) to analyse the total non-covalent
molecular weights. Lane 1: gel with 0% fat; lanes 2, 3 and 4: gels with lard at 4%, 8% and forces; and C, 0.25% β-mercaptoethanol (βME) to analyse the disulfide bands. The
12% fat levels; lanes 5, 6 and 7: gels with GL at 4%, 8% and 12% fat levels; lanes 8, 9 and means of the samples with different lowercase letters (a-f) significantly differed from
10: gels with PGL at 4%, 8% and 12% fat levels. one another (P b 0.05).
338 X. Diao et al. / Meat Science 121 (2016) 333–341

with the WHC measurement. As a consequence, the hydrogen bond (2004) reported that hydrophobic and hydrogen bonds were the main
plays an important role in the composite-gel formation. forces in the network formation of legumin proteins, and disulfide
When the gels were dissolved in the 0.5% SDS solution (Fig. 4B), a bonds had minimum involvements.
significantly higher amount (P b 0.05) of protein was extracted by SDS
from the composite gels with different concentrations of lard, GL or 3.5. Microstructures of composite gels
PGL than that from the gel without fat, which indicates that hydropho-
bic interactions also play an active role in the composite gels. During the The microstructures of the composite gels that were prepared with
gel formation process, when the fat concentration increases, the hydro- MP and lard, GL or PGL are shown using SEM in Fig. 5. Obvious variations
phobic groups become more exposed, which contributes to the increase were observed in the microstructure of composite gels with different
in hydrophobicity, and the unfolded protein molecules aggregate to types and concentrations of fat. Compared with the MP gel without
form a gel. The observation is consistent with the results of the penetra- fat, the gels with GL and PGL formed more compact and homogeneous
tion force. Wu et al. (2011) reported that an interactive protein film sur- three-dimensional network structures. Moreover, when the fat additive
rounding a fat droplet was conductive to form a composite gel. amount increased from 4% to 12%, the composite gels showed a more
To clarify the role of disulfide bond cross-linking, which results from compacted gel network, whereas the pure MP gel exhibited a large
sulfhydryl group explosion (Sano, Ohno, Otsuka-Fuchino, Matsumoto, & amount of void spaces or pores. Youssef and Barbut (2010) showed
Tsuchiya, 1994), the gels were treated with βME (Fig. 4C). There were that oil could function as fillers in the protein network in composite
some differences in protein solubility between the MP gel without fat gels and reduce the empty spaces. It was noted that the gels with GL
and those with GL or PGL (P b 0.05). The results suggest that the disul- or PGL had a more compacted and smoother gel network than the
fide bonds in the fat globule membrane and between the membrane gels with lard, which indicates that the addition of GL or PGL was
and the continuous protein gel matrix are also significant for the stabil- more beneficial to the gel network formation. Our previous experiment
ity of the composite gels. However, the gels in the βME solution had reveals that GL and PGL are smaller than lard in emulsion. Hence, PGL
lower protein solubility than those treated with urea and SDS can evenly disperse in the MP network and result in a compacted and
(P b 0.05), which shows that disulfide bonds are not a significant force homogeneous microstructure. These structural features explain why
in composite gels. O'Kane, Happe, Vereijken, Gruppen, and van Boekel MP gels with GL and PGL had stronger penetration forces and water

Fig. 5. Scanning electron micrographs (magnification: 2000×) of the composite gel that was prepared with myofibrillar proteins (MP) and lard, glycerolized lard (GL) or purified
glycerolized lard (PGL).
X. Diao et al. / Meat Science 121 (2016) 333–341 339

binding. Chen and Dickinson (1998) attributed the reinforcement of increased, whereas the random-coil content reduced (P b 0.05), which
composite milk protein gels to the interaction of the protein membrane indicates that the types and concentrations of fat have an essential effect
of active filler particles with protein in the gel matrix. on the secondary structure of MP. In addition, the composite gels with
GL and PGL had slightly higher α-helix and β-sheet fractions than
3.6. Changes in secondary structures those with lard (P b 0.05). This difference may be attributed to the em-
bedment of protein-coated GL and PGL in a continuous, three-dimen-
To gain insight into the gelation mechanisms of composite gels, the sional myofibrillar protein gel matrix, which transformed the
molecular conformation of MP was measured using FTIR to discuss the unordered structure into an ordered structure during the gel formation
changes in secondary structures of protein. The percentages of α- (Gordon & Barbut, 1990). Several reports have shown that the WHC of
helix, β-sheet, β-turn and random coil structures of MP in composite myosin gel is positively correlated with the α-helix fraction, a large
gels with lard, GL or PGL are shown in Fig. 6. Liu, Zhao, Xiong, Xie, and amount of α-helices prior to heating is beneficial for the WHC of myosin
Qin (2008) indicated that the unfolding of α-helix and formation of β- gel, and a large amount of β-sheet can improve the gel strength (Choi &
sheet favoured the gelation of porcine myosin. Sano et al. (1994) re- Ma, 2007; W. Liu, Z. Q. Zhang et al., 2010; R. Liu, S. M. Zhao et al., 2010).
vealed that the α-helix structure was an important conformation to These observations are consistent with our results of the WHC and pen-
maintain the secondary structure of native protein and was mainly sta- etration force of composite gels. In conclusion, lard, GL and PGL in com-
bilized by hydrogen bonds between carbonyl oxygen (\\CO\\) and posite gels can lead to more α-helix and β-sheet at the expense of
amino hydrogen (\\NH\\) of a polypeptide chain. Meng, Ma, and random coil structures. Furthermore, the higher additive fat content
Phillips (2003) reported that β-sheet was an important conformational can induce the more obvious secondary structural changes of MP in
component in the aggregated globular protein. Fig. 6 shows that α-helix composite gels.
and β-sheet was the predominant secondary structure of MP in com-
posite gels. With the addition of lard, GL and PGL, the contents of α- 3.7. Intrinsic tryptophan fluorescence
helix, β-sheet and β-turn of MP in the composite gels gradually
The fluorescence spectra were measured to investigate the interac-
tion between MP and fats in composite gels. Tryptophan is an essential
amino acid with relevant biological functions in muscle protein (Lu, Li,
Yin, Zhang, & Wang, 2008), and its location in protein affects the fluores-
cence energy (Burstein, Vedenkina, & Ivkova, 1973). At 280 nm, the
tryptophan residues of MP can be excited; therefore, the changes of ter-
tiary structure of the protein can be determined from the fluorescence
fluctuations (W. Liu, Z. Q. Zhang et al., 2010; R. Liu, S. M. Zhao et al.,
2010; Shen & Tang, 2012). When a protein is partially or completely un-
folded, the tryptophan residues become exposed to the hydrophilic en-
vironment, which decreases the fluorescence intensity (Pallarès,
Vendrell, Avilés, & Ventura, 2004). In the folded state, tryptophan resi-
dues are generally located in the hydrophobic environment of the pro-
tein with high fluorescence intensity. Puscasu and Birlouez-Aragon
(2002) also reported that tryptophan fluorescence had good correla-
tions with the tryptophan concentration. As shown in Fig. 7, the trypto-
phan fluorescence intensity of MP in the composite gels with lard, GL or
PGL remarkably decreased compared with that of the pure MP, which
indicates that adding lard, GL and PGL can make the protein unfold. In
addition, MP in the composite gel with PGL had significantly
(P b 0.05) lower fluorescence intensity values than that with lard at
identical fat concentrations. The higher fat content causes a further de-
crease in fluorescence intensity, which indicates further unfolding and
possible interactions between fat and tryptophan residues. The results
show that lard, GL and PGL can alter the tertiary structure of MP in com-
posite gels.

4. Conclusions

The types and concentrations of fat significantly affect the physico-


chemical and structural properties of MP composite gels. The gels that
were prepared with MP and GL or PGL had significantly higher penetra-
tion force and WHC as well as more compact and orderly microstructure
than other treatments. This effect was largely attributed to the stronger
interaction of MP and PGL through hydrogen bonds and hydrophobic
association. Furthermore, the additive content of GL and PGL had no ob-
vious effect on the participating proteins in composite gels but changed
the secondary and tertiary structures of MP in the gelation of composite
gels, which changed the gel properties. Such physicochemical interac-
tions between the protein and fat are responsible for the juiciness, ten-
derness and mouthfeel of low-fat meat products. In general, the
Fig. 6. Secondary structure fractions of composite gels that were prepared with
composite gels that were prepared with MP and GL or PGL improve
myofibrillar proteins (MP) and lard (A), glycerolized lard (GL) (B) or purified the gel quality, and lard DAGs have the potential for meat product
glycerolized lard (PGL) (C). applications.
340 X. Diao et al. / Meat Science 121 (2016) 333–341

References

Andersen, C. M., Frøst, M. B., & Viereck, N. (2010). Spectroscopic characterization of low-
and non-fat cream cheeses. International Dairy Journal, 20, 32–39.
Burstein, E. A., Vedenkina, N. S., & Ivkova, M. N. (1973). Fluorescence and the location of
tryptophan residues in protein molecules. Photochemistry and Photobiology, 18,
263–279.
Chen, J. S., & Dickinson, E. (1998). Viscoelastic properties of protein-stabilized emulsions:
Effect of protein-surfactant interactions. Journal of Agricultural and Food Chemistry, 46,
91–97.
Cheong, L. Z., Zhang, H., Xu, Y., & Xu, X. B. (2009). Physical characterization of lard partial
acylglycerols and their effects on melting and crystallization properties of blends
with rapeseed oil. Journal of Agricultural and Food Chemistry, 57, 5020–5027.
Choi, S. M., & Ma, C. Y. (2007). Structural characterization of globulin from common buck-
wheat (Fagopyrum esculentum Moench) using circular dichroism and Raman spec-
troscopy. Food Chemistry, 102, 150–160.
Diao, X. Q., Guan, H. N., Zhao, X., Chen, Q., & Kong, B. H. (2016). Properties and oxidative
stability of emulsions prepared with myofibrillar protein and lard diacylglycerols.
Meat Science, 115, 16–23.
Gordon, A., & Barbut, S. (1990). Cold stage scanning electron microscopy study of meat
batters. Journal of Food Science, 55, 1196–1198.
Gordon, A., & Barbut, S. (1992). Mechanisms of meat batter stabilization: A review. Critical
Reviews in Food Science and Nutrition, 32, 299–332.
Ishioroshi, M., Samejima, K., & Yasui, T. (1979). Heat-induced gelation of myosin: Factors
of pH and salt concentrations. Journal of Food Science, 44, 1280–1284.
Jiang, J., & Xiong, Y. L. (2013). Extreme pH treatments enhance the structure reinforce-
ment role of soy protein isolate and its emulsions in pork myofibrillar protein gels
in the presence of microbial transglutaminase. Meat Science, 93, 469–476.
Laemmli, U. K. (1970). Cleavage of structural proteins during the assembly of the head of
bacteriophage T4. Nature, 227, 680–685.
Lefevre, F., Fauconneau, B., Ouali, A., & Culioli, J. (2002). Thermal gelation of brown trout
myofibrils from white and red muscles: Effect of pH and ionic strength. Journal of the
Science of Food and Agriculture, 82, 452–463.
Liu, R., Zhao, S. M., Liu, Y. M., Yang, H., Xiong, S. B., Xie, B. J., et al. (2010b). Effect of pH on
the gel properties and secondary structure of fish myosin. Food Chemistry, 121,
196–202.
Liu, R., Zhao, S. M., Xiong, S. B., Xie, B. J., & Qin, L. H. (2008). Role of secondary structures in
the gelation of porcine myosin at different pH values. Meat Science, 80, 632–639.
Liu, W., Zhang, Z. Q., Liu, C. M., Xie, M. Y., Tu, Z. C., Liu, J. H., et al. (2010a). The effect of
dynamic high-pressure microfluidization on the activity, stability and conformation
of trypsin. Food Chemistry, 123, 616–621.
Lu, P., Li, D. F., Yin, J. D., Zhang, L. Y., & Wang, Z. Y. (2008). Flavour differences of cooked
longissimus muscle from Chinese indigenous pig breeds and hybrid pig breed
(Duroc × landrace × large white). Food Chemistry, 107, 1529–1537.
Maki, K. C., Davidson, M. H., Tsushima, R., Matsuo, N., Tokimitsu, I., Umporowicz, D. M.,
et al. (2002). Consumption of diacylglycerol oil as part of a reduced energy diet en-
hances loss of body weight and fat in comparison with consumption of a triacylglyc-
erol control oil. The American Journal of Clinical Nutrition, 76, 1230–1236.
Mendoza, E., García, M. L., Casas, C., & Selgas, M. D. (2001). Insulin as fat substitute in low
fat, dry fermented sausages. Meat Science, 57, 387–393.
Meng, G. T., Ma, C. Y., & Phillips, D. L. (2003). Raman spectroscopic study of globulin from
Phaseolus angularis (red bean). Food Chemistry, 81, 411–420.
Meng, X. H., Zou, D. Y., Shi, Z. P., Duan, Z. Y., & Mao, Z. G. (2004). Dietary diacylglycerol
prevents high-fat diet-induced lipid accumulation in rat liver and abdominal adipose
tissue. Lipids, 39, 37–41.
Miklos, R., Mora-Gallego, H., Larsena, F. H., Serra, X., Cheong, L. Z., Xu, X., et al. (2014). In-
fluence of lipid type on water and fat mobility in fermented sausages studied by low-
field NMR. Meat Science, 96, 617–622.
Miklos, R., Xu, X. B., & Lametsch, R. (2011). Application of pork fat diacylglycerols in meat
emulsions. Meat Science, 87, 202–205.
Mora-Gallego, H., Serra, X., Guàrdia, M. D., Miklos, R., Lametsch, R., & Arnau, J. (2013). Ef-
fect of the type of fat on the physicochemical, instrumental and sensory characteris-
tics of reduced fat non-acid fermented sausages. Meat Science, 93, 668–674.
Morita, O., & Soni, M. G. (2009). Safety assessment of diacylglycerol oil as an edible oil: A
review of the published literature. Food and Chemical Toxicology, 47, 9–21.
Nakajima, Y. (2004). Water-retaining ability of diacylglycerol. Journal of the American Oil
Chemists Society, 81, 907–912.
Nejat, E. J., Polotsky, A. J., & Pal, L. (2010). Predictors of chronic disease at midlife and be-
yond-the health risks of obesity. Maturitas, 65, 106–111.
Noronha, N., Duggan, E., Ziegler, G. R., O'Riordan, E. D., & O'Sullivan, M. (2008). Investiga-
tion of imitation cheese matrix development using light microscopy and NMR
relaxometry. International Dairy Journal, 18, 641–648.
O'Kane, F. E., Happe, R. P., Vereijken, J. M., Gruppen, H., & van Boekel, M. A. J. S. (2004).
Heat-induced gelation of pea legumin: Comparison with soybean glycinin. Journal
Fig. 7. Tryptophan fluorescence of MP in composite gels that were prepared with of Agricultural and Food Chemistry, 52, 5071–5078.
myofibrillar proteins (MP) and lard (A), glycerolized lard (GL) (B) or purified Pallarès, I., Vendrell, J., Avilés, F. X., & Ventura, S. (2004). Amyloid fibril formation by a par-
glycerolized lard (PGL) (C). tially structured intermediate state of α-chymotrypsin. Journal of Molecular Biology,
342, 321–331.
Puscasu, C., & Birlouez-Aragon, I. (2002). Intermediary and/or advanced Maillard prod-
ucts exhibit prooxidant activity on Trp: In vitro study on alpha-lactalbumin. Food
Acknowledgements Chemistry, 78, 399–406.
Ramírez-Suárez, J. C., Xiong, Y. L., & Wang, B. (2001). Transglutaminase cross-linking of
This study was funded by the Science and Technology Major Projects bovine cardiac myofibrillar proteins and its effect on protein gelation. Journal of
Muscle Foods, 12, 85–96.
in Heilongjiang (grant no. GA15B302) and National Natural Science Rosenvold, K., & Andersen, H. J. (2003). Factors of significance for pork quality-A review.
Foundation of China (grant no. 31471599). Meat Science, 64, 219–237.
X. Diao et al. / Meat Science 121 (2016) 333–341 341

Salomonsen, T., Sejersen, M. T., Viereck, N., Ipsen, R., & Engelsen, S. B. (2007). Water mo- Xia, X. F., Kong, B. H., Liu, Q., & Liu, J. (2009). Physicochemical change and protein oxida-
bility in acidified milk drinks studied by low-field H-1 NMR. International Dairy tion in porcine longissimus dorsi as influenced by different freeze-thaw cycles. Meat
Journal, 17, 294–301. Science, 83, 239–245.
Sano, T., Ohno, T., Otsuka-Fuchino, H., Matsumoto, J. J., & Tsuchiya, T. (1994). Carp natural Xiong, Y. L., & Brekke, C. J. (1991). Gelation properties of chicken myofibrils treated with
actomyosin: Thermal denaturation mechanism. Journal of Food Science, 59, calcium and magnesium chlorides. Journal of Muscle Foods, 2, 21–36.
1002–1008. Xiong, Y. L., & Kinsella, J. E. (1991). Influence of fat globule-membrane composition and
Shaarani, S. M., Nott, K. P., & Hall, L. D. (2006). Combination of NMR and MRI quantitation fat type on the rheological properties of milk based composite gels. II. Results.
of moisture and structure changes for convection cooking of fresh chicken meat. Meat Milchwissenschaft, 46, 207–212.
Science, 72, 398–403. Xu, X. L., Han, M. Y., Fei, Y., & Zhou, G. H. (2011). Raman spectroscopic study of heat-in-
Shen, L., & Tang, C. H. (2012). Microfluidization as a potential technique to modify surface duced gelation of pork myofibrillar proteins and its relationship with textural charac-
properties of soy protein isolate. Food Research International, 48, 108–118. teristic. Meat Science, 87, 159–164.
Sun, J. X., Wu, Z., Xu, X. L., & Li, P. (2012). Effect of peanut protein isolate on functional Youssef, M. K., & Barbut, S. (2010). Physiochemical effects of the lipid phase and protein
properties of chicken salt-soluble proteins from breast and thigh muscles during level on meat emulsion stability, texture and microstructure. Journal of Food
heat-induced gelation. Meat Science, 91, 88–92. Science, 75, S108–S114.
Sun, X. D., & Holley, R. A. (2011). Factors influencing gel formation by myofibrillar pro- Zhang, Z. Y., Yang, Y. L., Tang, X. Z., Chen, Y. J., & You, Y. (2015). Chemical forces and water
teins in muscle foods. Comprehensive Reviews in Food Science and Food Safety, 10, holding capacity study of heat-induced myofibrillar protein gel as affected by high
33–51. pressure. Food Chemistry, 188, 111–118.
Wu, M. G., Xiong, Y. L., & Chen, J. (2011). Rheology and microstructure of myofibrillar pro- Ziegler, G. R., & Acton, J. C. (1984). Mechanisms of gel formation by proteins of muscle tis-
tein-plant lipid composite gels: Effect of emulsion droplet size and membrane type. sue. Food Technology, 38, 77–81.
Journal of Food Engineering, 106(4), 318–324.
Wu, M. G., Xiong, Y. L., Chen, J., Tang, X. Y., & Zhou, G. H. (2009). Rheological and micro-
structural properties of porcine myofibrillar protein–lipid emulsion composite gels.
Journal of Food Science, 74, 207–217.

You might also like