You are on page 1of 9

International Journal of Biological Macromolecules 115 (2018) 1174–1182

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules

journal homepage: http://www.elsevier.com/locate/ijbiomac

Physicochemical, molecular, emulsifying and rheological


characterizations of sage (Salvia splendens) seed gum
Soo Young Seo, Yu-Ra Kang, Yun-Kyung Lee, Jin Hye Lee, Yoon Hyuk Chang ⁎
Department of Food and Nutrition, Kyung Hee University, Seoul 02447, Republic of Korea

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

Article history: The objective was to investigate the physicochemical, molecular, rheological, and emulsifying properties of water
Received 9 February 2018 soluble-sage seed gum (WSG). WSG mainly comprised galacturonic acid and xylose. FTIR and NMR analyses con-
Received in revised form 27 March 2018 firmed the presence of pectic polysaccharides in WSG. Additionally, the molecular weight of WSG was higher
Accepted 30 April 2018
than that of pectin standard. Compared to pectin standard solutions, WSG solutions exhibited higher shear thin-
Available online 03 May 2018
ning behavior and higher values of apparent viscosity (ηa,100) and consistency index (K) in steady shear measure-
Keywords:
ments. According to the results of frequency sweep test, the dynamic moduli (G′ and G″) for WSG solutions were
Sage (Salvia splendens) seed gum increased with increasing frequency and concentration. The changes in dynamic moduli of WSG solutions as a
Chemical structure function of aging time at 4 °C indicated that WSG could form a more rigid network than pectin standard. Accord-
Rheological properties ing to the results of temperature sweep test, the dynamic moduli of WSG solutions were higher than those of pec-
Emulsifying properties tin standard solutions. Emulsion capacity and stability analyses indicated that WSG (58 and 56%, respectively)
had better emulsifying properties than pectin standard (46 and 37%). In conclusion, compared to pectin standard,
superior rheological and emulsifying properties observed in WSG might be related to higher molecular weight
and protein content, respectively.
© 2018 Published by Elsevier B.V.

1. Introduction Dynamic rheological properties can be used along with steady shear
rheological properties to provide insight on the structure of the gum so-
Salvia splendens is the plants of belongs to Labiatae family, a large lutions [12]. Dynamic shear rheological tests have been used to charac-
number of annual and perennial plant species and is one of several gen- terize or classify the viscoelastic properties of macromolecular
era commonly referred to as sage [1]. In addition to Salvia splendens, dif- dispersions [13]. In a dynamic test, the storage modulus (G′), loss mod-
ferent Salvia species, such as Salvia macrosiphon, Salvia officinalis, Salvia ulus (G″) and complex viscosity (η*) have been determined using a si-
sclarea, Salvia miltiorrhiza and Salvia grandiflora, have been used for food nusoidal strain cycle. The dynamic test is a non-destructive technique
and pharmaceutical purposes [2–4]. Sage seeds are small rounded seed, enabling measurements to be made without structural breakdown to
which quickly swells in water to give mucilaginous gum [4]. The sage the sample and it allows researchers to relate dynamic rheological pa-
seed gum has viscous suspension properties which are comparable rameters with the molecular structure of the sample. Thus, dynamic
with commercial food hydrocolloids, and therefore it is regarded as a rheology is one of the most extensively used methods to assess the vis-
new source of hydrocolloid [5–7]. Gum is hydrophilic high molecular coelastic behavior of gum solutions or gels [14].
weight biopolymers which have the ability to greatly hydrate in contact At present, the food industry has placed a significant amount of at-
with water, thus producing colloidal systems of different structures [8]. tention on gums for their roles as stabilizers in oil in water (O/W) emul-
Rheological and emulsifying characterizations of gum depend on con- sions. Most of gums can form stable O/W emulsions because they are
centration, temperature, dispersion, dissolution, electrolyte, mechanical adsorbed at the oil-water interfaces [15]. The adsorption properties
treatment and structure that influence food process conditions [9]. are seemed to be attributed to the hydrophobic groups or proteinaceous
Gums, such as pectin, guar gum, locust bean gum and tara gum, have moieties related to gums [16]. Therefore, it is important to extract gums
been widely used as a thickener, emulsifier, stabilizer and gel strength- from natural sources and investigate relation between their structure
ener in the food industry due to their ability to form viscous solutions at and emulsifying capacities.
relatively low concentrations [10,11]. Until now, a few researchers have focused on the structural and
physicochemical properties of sage gums extracted from Salvia
macrosiphon seed [4]. It has been reported that the gum extracted
⁎ Corresponding author. from Salvia macrosiphon seed was a galactomannan (mannose to galac-
E-mail address: yhchang@khu.ac.kr (Y.H. Chang). tose ratio of 1.78–1.93:1) containing 28.2–32.2% uronic acids and could

https://doi.org/10.1016/j.ijbiomac.2018.04.173
0141-8130/© 2018 Published by Elsevier B.V.
S.Y. Seo et al. / International Journal of Biological Macromolecules 115 (2018) 1174–1182 1175

be used as stabilizer, thickener, and emulsifier in food products because 2.4. Fourier transform-infrared spectroscopy
of its characteristic physicochemical and rheological properties [4,17].
In general, the physicochemical, molecular, emulsifying, and rheological Fourier transform infrared (FTIR) spectrum of WSG was determined
properties of gums can be remarkably affected by the species, growth using a Fourier transform infrared spectrophotometer (Spectrum GX,
conditions, and origin of Salvia. However, no detailed studies on sage Perkin Elmer, Massachusetts, USA). The sample was mixed with 100 mg
seed gums extracted from Salvia splendens seed have been found in pre- of KBr under the anhydrous condition, and then pressed into a KBr pellet.
vious literatures. Therefore, the present study was carried out to inves- The pellet was scanned in the spectra range of 400–4000 cm−1.
tigate the chemical composition, molecular weight distribution, partial
structure, and rheological and emulsion properties of water soluble- 2.5. 1H and 13C NMR spectroscopy
sage seed gum (WSG) to explore its potential application as a thickening
agent and emulsion stabilizer in the food industry. WSG (20 mg) was dissolved in D2O (99.9 atom %) and kept at room
temperature. 1H and 13C NMR spectra were performed by 300 NMR
2. Materials and methods spectrometer (JNM-AL300, JEOL, Tokyo, Japan), with a frequency of
75 MHz at 25 °C. The chemical shifts (δ) were reported in ppm units
2.1. Materials using tetramethylsilane (TMS) as an internal standard; coupling
constants (J) were given in Hertz and referred to apparent peak
The sage seed (Salvia splendens) was obtained from Danong Ltd. multiplicities.
(Namyangju, Korea). The sage seed was grinded into fine powder
using a laboratory blender (7011S, Waring Blender, Torrington, CT, 2.6. Molecular weight distribution
USA), sieved and stored at −20 °C until further use. Commercial pectin
standard was provided from CP Kelco [citrus pectin, high methoxyl (DE The molecular weight distribution of WSG was analyzed using a
N 50%), Lille Skensved, Denmark]. Monosaccharide standards of high-performance lipid chromatography system (Agilent 1100, Palo
rhamnose, arabinose, glucose, xylose, galactose and mannose were Alto, CA, USA) equipped with a refractive index detector (RID). WSG
purchased from Sigma Chemical Co. (St. Louis, MO, USA). All other (10 mg) was dissolved in distilled water (10 mL) and passed through
chemicals and solvents used were of analytical grade unless otherwise a gel filtration chromatographic column of PL aquagel-OH MIXED-H
specified. column (300 × 7.5 mm, Agilent, Amherst, MA, USA). The column was
operated at 30 °C and eluted by triple distilled water at a flow rate of
1 mL/min with an injection volume of 20 μL. The molecular weight
2.2. Extraction of water soluble-sage seed gum
was estimated by the standard curve which was calibrated with dextran
standards (Mw 12,000, 50,000, 150,000, and 410,000 Da, Sigma Chem-
The extraction procedure of WSG followed the method of Razavi
ical Co., St. Louis, MO, USA).
et al. [4] with slight modifications. First, the dried powder of sage seed
was mixed with acetone, hexane and 95% ethanol to remove lipids.
2.7. Rheological properties
The residue was placed in a beaker and distilled water was added
in 1:20 (w/v). The temperature during extraction was maintained at
2.7.1. Sample preparation
70 °C for 24 h using a shaking water bath (BS-11, Jeio tech Co., Ltd., Dae-
WSG was moderately stirred for 30 min at room temperature in a
jeon, Korea) and the extract was centrifuged (VS-5000N, Vision scien-
magnetic stirrer for hydration and then heated at 70 °C for 1 h in a
tific Co., Ltd., Daejeon, Korea) at 3500 rpm for 10 min. The collected
water bath with mild agitation provided by a magnetic stirrer. During
supernatants were mixed with 95% ethanol and kept at 4 °C for over-
heating, a screw-cap Erlenmeyer flask was used to prevent water evap-
night. The ethanol-insoluble precipitate was collected by centrifugation
oration. At the end of the heating period, the sample mixture was cooled
at 3500 rpm for 10 min and dried in a dry oven (Isotemp 500 Series,
at room temperature and placed on a rheometer plate for measurement
Fisher Scientific Inc., Pittsburgh, PA, USA) at 60 °C for 24 h.
of its rheological properties.

2.3. Chemical composition and monosaccharide composition analysis 2.7.2. Steady shear rheological analysis
The steady shear rheological properties of pectin standard and WSG
The chemical composition of WSG including moisture content, ash solutions (2, 4, 6 and 8%, w/w) at 25 °C on a strain controlled Phyusica
content, protein content and fat content was determined according to MCR 102 rheometer (Anton Paar, Graz, Austria) using a plate/plate
A.O.A.C. methods [18] by 925.09B, 923.03, 979.09 and 920.39C, respec- geometry (50 mm diameter, 0.5 mm gap). The steady shear (shear
tively. The content of total sugar was determined by the phenol- stress and shear rate) data were obtained over a shear rate rage of
sulfuric acid method [19]. The uronic acid content was measured by 0.1–1000 s−1. After loading, the gum solution was held for 5 min at
m-hydroxyphenyl colorimetric method [20] with D-galacturonic acid the initial test temperature before testing to allow stress relaxation
(Sigma Chemical Co., St. Louis, MO, USA) as the standard. and temperature equilibration. In order to describe the variation in the
The neutral monosaccharide composition of WSG was determined rheological properties of gum solutions under steady shear, the data
by the procedure of Cui and Chang [21] with slight modifications. were fitted to the well-known power law (Eq. (1)) model, which is
WSG (10 mg) was hydrolyzed with 2 M TFA (2 mL) at 121 °C for 1 h. used extensively to describe the flow properties of non-Newtonian liq-
After TFA was removed by a rotary vacuum evaporator (N-N1, Eyela, uids in theoretical analysis as well as in practical engineering applica-
Japan), distilled water (1 mL) was added to dissolve the residue. The tions [22]:
sample was passed through a 0.45 μm filter and injected to a high-
performance anion exchange chromatography with pulsed ampero-
σ ¼ Kγ_
metric detection (HPAEC-PAD) (Dionex, Sunnyvale, CA, USA). Separa- n
ð1Þ
tion of each neutral sugar was performed on a CarboPac PA1 column
(250 mm × 4 mm, Dionex, Sunnyvale, CA, USA). The temperature was
kept at 30 °C and the injection volume was 15 μL. The solvents were where σ is the shear stress (Pa), γ_ is the shear rate (s−1), K is the consis-
A: 18 mM NaOH and B: 200 mM NaOH. The solvent flow rate was tency index (Pa·sn), n is the flow behavior index (dimensionless). The
1 mL/min. The rhamnose, arabinose, glucose, xylose, galactose and apparent viscosity (ηa,100) at 100 s−1 was calculated using the magni-
mannose were used as reference standards. tudes of K and n.
1176 S.Y. Seo et al. / International Journal of Biological Macromolecules 115 (2018) 1174–1182

2.8. Dynamic shear rheological properties Table 1


Chemical composition and monosaccharide composition of water soluble-
sage seed gum (WSG).
2.8.1. Frequency sweep
Frequency sweep measurements were performed to investigate Chemical composition WSG
the dynamic rheological properties [storage modulus (G′), loss modulus Moisture content (%) 8.73 ± 0.54
(G″), complex viscosity (η*) and tan δ (G″/G′)] of pectin standard and Ash content (%) 14.16 ± 0.27
WSG solutions (2, 4, 6, 8, 10, and 12%, w/w) on a strain controlled Protein content (%) 16.62 ± 0.19
Fat content (%) ND
MCR 102 rheometer (Anton Paar, Graz, Austria). Prior to the frequency
Total sugar (%) 62.79 ± 0.38
sweep tests, a strain sweep test at a constant frequency of 6.3 rad/s Monosaccharide relative (%)
was carried out to determine the linear viscoelastic region. The fre- Fucose 0.13 ± 0.00
quency sweep tests were performed at a strain value of 0.02 (2%) Rhamnose 0.80 ± 0.10
(within the linear viscoelastic region). Frequency sweep tests of WSG Arabinose 2.44 ± 0.04
Galactose 2.18 ± 0.23
were performed using a plate/plate geometry (50 mm diameter, Glucose 2.48 ± 0.41
0.5 mm gap) at 25 °C and frequencies (ω) from 0.63 to 63 rad/s. Xylose 22.89 ± 0.75
Total uronic acid (%) 31.75 ± 0.45
2.8.2. Time sweep All measurements are on a dry weight basis except for moisture.
For time sweep measurements in the aging process, pectin standard Data presented are means of three replicates ± the standard deviations.
and WSG solutions (8 and 10%, w/w) were loaded onto the rheometer ND: not detected.
plate at 4 °C. The exposed sample edge was covered with a thin layer
of light paraffin oil to prevent evaporation during measurements. The rhamnose, arabinose, galactose and glucose were also found in WSG.
G′ and G″ values were monitored during aging at 4 °C for 1 h at The presence of various neutral sugars in WSG can be attributed to a
6.28 rad/s and a strain value of 0.01 (1%) (within the linear viscoelastic more complex polysaccharide structure. This finding was contradictory
region). with the previous results from Razavi et al. [4], who reported that the
mannose and galactose (mostly galactomannans) were the main
2.8.3. Temperature sweep component of monosaccharides in the sage seed gum extracted from
The temperature sweep measurements were performed at strain Salvia macrosiphon. Therefore, it was found that different Salvia species
value of 0.01 (1%) (within the linear viscoelastic region), while the fre- had the different monosaccharide composition.
quency was fixed at 6.28 rad/s. To investigate the effect of temperature
on the rheological properties of pectin standard and WSG solutions (8 3.2. Fourier transform-infrared spectroscopy (FTIR)
and 10%, w/w), a program was set up. WSG was transferred to the rhe-
ometer plate at 4 °C. The exposed sample edge was covered with a thin The FTIR spectrum in Fig. 1(A) shows the characteristic absorption
layer of light paraffin oil to prevent evaporation during measurements. peaks of WSG. The peaks observed in the regions between 1000 and
Briefly, the programs included: (1) a heating step with a linear temper- 1150 cm−1 were corresponded to the skeletal C\\O and C\\C vibration
ature increased from 4 to 95 °C, using heating rate of 5 °C/min, (2) a bands of glycosidic bonds and pyranoid ring, respectively [24–26]. Wu
cooling step with a linear temperature decreased from 95 to 4 °C, et al. [27] reported that pectic saccharides extracted from boat-fruited
using cooling rate of 5 °C/min. sterculia seeds showed the bands at 1146, 1070, and 1035 cm−1,
which were characteristic peaks of pectic polysaccharides.
2.9. Emulsion capacity and stability The peaks in the regions between 1200 and 1800 cm−1 reflected the
stretching modes of carboxyl groups [24]. The peak at about 1740 cm−1
The emulsion capacity and stability were determined based on the was attributed to C_O stretching vibration of the methyl esterified car-
method described by Chang et al. [23]. boxylic acid, while the peak at about 1415 cm−1 were attributed to
C_O stretching vibration of nonesterified carboxyl groups [28,29].
2.10. Statistical analysis The strong and broad absorption peak at 3432.03 cm−1 was assigned
to the O\\H stretch vibration of hydroxyl groups [30] and the signal at
All statistical analyses were performed using SAS version 9.3 (SAS 2922.10 cm−1 was attributed to C\\H stretching (including CH, CH2,
Institute Inc., Cary, NC, USA). Analysis of variance (ANOVA) was per- and CH3) vibration of the methyl groups [24].
formed using the general linear models (GLM) procedure to determine
significant differences among the samples. Means were compared by 3.3. 1H and 13C NMR analysis
using Fisher's least significant difference (LSD) procedure. Significance
was defined at the 5% level. The signals of WSG in 1H NMR spectra (Fig. 1(B)) were assigned
based on the chemical shifts compared with the previous literatures.
3. Results and discussion The signal at 4.73 ppm can be assigned to the strong chemical shift of
solvent D2O. In the anomeric protons, the signals at 5.08 and 4.84 ppm
3.1. Chemical composition were assigned to H-1 of nonesterified and methyl esterified 1,4-linked
α-D-GalpA, respectively [31]. The signal at 3.76 ppm of WSG was
The extraction yield of WSG was 7.93% (data are not shown). The attributed to H-2 of methyl groups binding to carboxyl groups of 1,4-
chemical composition of WSG is shown in Table 1. WSG contained linked α-D-GalpA, corresponding to the previous results reported by
8.73% moisture, 14.16% crude ash, 16.62% crude protein and 62.79% Košťálová et al. [32]. The low-intensity signal at around 2.20 ppm was
total sugar (Table 1). WSG had 31.75% of total uronic acid contents, derived from O-acetyl groups of GalpA [33]. The methyl rhamnose sig-
reflecting a poly-electrolyte nature and the relative amount of acidic nal at 1.05 ppm of WSG indicated the O-2,4-linked α-L-Rhap [34]. The
polysaccharides in the gum. Razavi et al. [4] also reported that sage two distinguished peaks at 4.42 and 4.19 ppm of WSG apparently
seed gum extracted from Salvia macrosiphon contained 28.2–32.2% of indicated the unesterified and esterified β-D-Xylp, respectively [35].
uronic acid. The 13C NMR spectrum of WSG and its derivatives is presented
Neutral sugar analysis by high performance anion exchange chro- in Fig. 1(C). In intense anomeric regions, the signals at 100.2 and
matography (HPAEC) revealed the presence of xylose as the main neu- 99.0 ppm were attributed to C-1 of methyl esterified and non-
tral sugar component in WSG (Table 1). Minor amounts of fructose, esterified 1,4-linked α-D-GalpA, respectively [36]. The signals at 71.01,
S.Y. Seo et al. / International Journal of Biological Macromolecules 115 (2018) 1174–1182 1177

Fig. 1. FTIR and 1D NMR spectra of water soluble-sage seed gum (WSG); (A) FTIR, (B) 1H NMR, and (C) 13C NMR.

71.8, 73.17, and 76.95 ppm of WSG were attributed to the C-2, C-3, C-4, weight (Mw), z-average molecular weight (Mz), polydispersity
and C-5 of α-D-GalpA, respectively [29]. Signals at 176.2 and 174.7 ppm (Mw/Mn), and number average molecular weight (Mn) (Table 2).
were attributed to C-6 of the carboxyl groups of 1,4-linked α-D-GalpA The molecular weight (Mw) (3.214 × 10 11) of WSG was higher
and methyl estered carboxyl group of α-D-GalpA, respectively [37]. than that (1.268 × 1011 ) of pectin standard. The number-average
molecular weight (Mn) and z-average molecular weight (Mz) of
3.4. Molecular weight distribution WSG were also higher than those of pectin standard. Comparison re-
vealed that the polydispersity (Mw/Mn) of WSG was lower than that
The results on molecular weights of commercial pectin standard of pectin standard, indicating a broad range of molecular weight dis-
and WSG showed that the different molecular weights were ob- tribution [38]. Therefore, WSG showed a macromolecular compound
tained by various parameters, such as weight average molecular in comparison to pectin standard.
1178 S.Y. Seo et al. / International Journal of Biological Macromolecules 115 (2018) 1174–1182

Table 2
Molecular characterization, emulsion capacity and emulsion stability of pectin standard and water soluble-sage seed gum (WSG).

Sample Molecular characterization Emulsion capacity (%) Emulsion stability (%)

Mw × 1011 Mn × 1010 Mz × 1012 Mw/Mn × 10

Pectin standard 1.268 0.380 1.020 3.340 46.13 ± 0.09b 36.93 ± 1.06b
WSG 3.214 1.847 1.669 1.740 57.77 ± 1.61a 55.80 ± 0.63a

Mw, weight average molecular weight; Mn, number average molecular weight; Mz, z-average molecular weight; Mw/Mn, polydispersity.
Data presented are means of three replicates ± the standard deviations.
Means (n = 3) within the same column with different superscript letters differ significantly (p b 0.05).

3.5. Steady shear rheological analysis The ηa,100 and K values for pectin standard and WSG solutions were
significantly affected by their concentration (Table 3). For instance, ele-
The analysis of flow behavior of pectin standard and WSG solu- vating concentration of all sample solutions from 2 to 8% resulted in a
tions by power law model (Eq. (1)) is shown in Table 3. In general, substantial increase in the values of ηa,100 and K. The increase in the
power law model presented coefficients of determination (R2 ) N values of ηa,100 and K for all sample solutions at the higher concentration
0.97 for all samples. The flow behavior indexes (n), which indicate could be attributed to the greater number of junction zones which were
the extent of shear thinning behavior as it deviates away from 1, formed in polysaccharide solutions. These junction zones in polysaccha-
ranged from 0.44 to 0.85, reflecting pseudoplastic characteristics of ride solutions are known to cause flow interruption [43]. Accordingly, it
all samples. The observed shear thinning behavior of pectin standard is plausible in the present study that the molecules could be drawing
and WSG solutions can be explained by disruption of intermolecular closer to one another and junction zones could be more readily formed
entanglements between polymer chains during shearing, as de- with increasing the concentration, which could cause an increase in the
scribed by Morris [39]. In the same concentration range, the n values ηa,100 and K values of all sample solutions at higher concentration.
of WSG solutions were significantly lower than those of pectin
standard solutions, explaining a more pronounced shear thinning 3.6. Frequency sweep measurements
behavior of WSG solutions. Morris et al. [40] reported that polysac-
charides with high molecular weight presented a more pronounced Fig. 2 shows changes in storage modulus (G′), loss modulus (G″),
shear thinning behavior. Therefore, the result of the higher shear and complex viscosity (η*) as a function of frequency (ω) for pectin
thinning behavior of WSG solutions may be explained by the higher standard and WSG solutions at different concentration (2, 4, 6, 8, 10,
molecular weight of WSG compared to pectin standard, as indicated and 12%, w/w) at 25 °C. As seen in Fig. 2, pectin standard and WSG so-
in Table 2. lutions at the identical concentration showed the same trends, but the
A concentration (2–8%) had a distinct effect on the n values of pectin magnitudes of G′ and G″, and η* of WSG solutions were much higher
standard and WSG solutions (Table 3). For instance, an increase in the than those of pectin standard solutions in the frequency range from
concentration of pectin standard and WSG solutions from 2 to 8% caused 6.3 to 63 rad/s. The magnitudes of G′ and G″ increased with an increase
a significant reduction of n values from 0.85 to 0.60 and 0.67 to 0.44, re- in frequency, showing a frequency dependency for all concentration of
spectively, showing the higher shear thinning behavior of all sample so- sample solutions. Furthermore, for all concentration of sample solu-
lutions at the higher concentration. Anvari et al. [41] noted that the tions, the η* values decreased linearly with an increase in frequency, in-
higher shear thinning behavior at the higher concentration can be ex- dicating a strong shear thinning behavior of them. The increase in the
plained by the increased amounts of high molecular weight molecules magnitudes of G′, G″, and η* at the higher concentration of sample solu-
in the liquid phase. tions was also confirmed. The foregoing results could be explained by
In the same concentration range, the apparent viscosity (ηa,100) and the greater network development in higher concentration of sample so-
consistency index (K) values of WSG solutions were significantly higher lutions. The network development is known to be formed by the strong
than those of pectin standard solutions (Table 3). According to Wood interaction between closed-packed molecules in solution [44].
[42], an increase in the molecular weight of polysaccharides caused In the experimental range of frequency indicated in Fig. 2, the visco-
the increase in their viscosity. Therefore, it was speculated that higher elastic behavior of all sample solutions was obviously dependent on
ηa,100 and K values of WSG solutions might have resulted from higher sample concentration. At low sample concentration (b6%), G″ was supe-
molecular weight of WSG compared to that of pectin standard as rior to G′, showing that sample solutions tended to present mainly vis-
shown in Table 2. cous characteristics instead of a clear tendency to form a gel. Whereas,

Table 3
Apparent viscosity (ηa,100), consistency index (K), flow behavior index (n), and coefficients of determination (R2) of pectin standard and water soluble-sage seed gum (WSG) solution with
different concentrations at 25 °C.

Concentration (%) Apparent viscosity Consistency index Flow behavior index R2


n
ηa,100 (Pa·s) K (Pa·s ) n (–)

Pectin standard
2 0.08 ± 0.00d 0.16 ± 0.00d 0.85 ± 0.01a 1.00 ± 0.00
4 0.64 ± 0.00c 1.41 ± 0.01c 0.83 ± 0.00b 1.00 ± 0.00
6 2.30 ± 0.03b 7.97 ± 0.07b 0.73 ± 0.00c 0.99 ± 0.00
8 5.62 ± 0.07a 35.75 ± 0.37a 0.60 ± 0.00d 0.98 ± 0.00

WSG
2 0.22 ± 0.01d 1.00 ± 0.06d 0.67 ± 0.00a 1.00 ± 0.00
4 1.28 ± 0.03c 8.38 ± 0.20c 0.59 ± 0.00b 1.00 ± 0.00
6 4.32 ± 0.02b 40.33 ± 0.39b 0.51 ± 0.00c 0.99 ± 0.00
8 9.95 ± 0.35a 129.57 ± 4.84a 0.44 ± 0.01d 0.97 ± 0.00

Data presented are means of three replicates ± the standard deviations.


Means (n = 3) within the same column with different superscript letters differ significantly (p b 0.05).
S.Y. Seo et al. / International Journal of Biological Macromolecules 115 (2018) 1174–1182 1179

Fig. 2. Plots of log G′ (▲), G″ (□), and η* (+) versus log ω for pectin standard and water soluble-sage seed gum (WSG) at different concentration (2, 4, 6, 8, 10, and 12%) at 25 °C.

sample solutions with concentration of 8 and 10% initially exhibit vis- [22]. Finally, in sample solution with concentration of 12%, G′ exceeded
cous behavior and, above the cross-over point, G′ values dominated G″ over most of the frequency range investigated, suggesting a clear ten-
over G″ values, representing elastic behavior. These rheological patterns dency to form weak-gel network.
of sample solutions (8 and 10%) were typical of viscoelastic fluids which The G′, G″, and η* values of pectin standard and WSG solutions at
were consistent with those of other gums as describe in previous studies 6.3 rad/s are shown in Table 4. The G′ and G″ values of WSG solutions

Table 4
Storage modulus (G′), loss modulus (G″), complex viscosity (η*), and tan δ at 6.3 rad/s of pectin standard and water soluble-sage seed gum (WSG) at 25 °C.

Sample Concentration (%) G′ (Pa) G″ (Pa) η* (Pa s) tan δ

Pectin standard 2 0.20 ± 0.05e 0.65 ± 0.04e 0.11 ± 0.01e 3.43 ± 0.75a
4 2.43 ± 0.46e 6.93 ± 0.30e 1.17 ± 0.07e 2.92 ± 0.47a
6 14.32 ± 0.51d 31.83 ± 0.85d 5.54 ± 0.16d 2.22 ± 0.02b
8 60.53 ± 1.76c 97.75 ± 2.63c 18.25 ± 0.50c 1.62 ± 0.00bc
10 194.85 ± 15.24b 225.95 ± 13.21b 47.36 ± 3.16b 1.16 ± 0.02cd
12 412.38 ± 3.98a 389.09 ± 4.27a 89.99 ± 0.91a 0.94 ± 0.00d
WSG 2 3.55 ± 0.49e 5.81 ± 0.58e 1.08 ± 0.12d 1.64 ± 0.07a
4 17.01 ± 0.16e 27.52 ± 0.37e 5.14 ± 0.06d 1.62 ± 0.01a
6 90.18 ± 1.94d 120.06 ± 1.33d 23.83 ± 0.35d 1.33 ± 0.01b
8 337.53 ± 8.92c 361.24 ± 3.46c 78.48 ± 1.35c 1.07 ± 0.02c
10 952.93 ± 4.44b 854.78 ± 5.95b 203.20 ± 1.05b 0.90 ± 0.00d
12 2096.01 ± 84.56a 1621.66 ± 48.09a 408.03 ± 32.76a 0.78 ± 0.01e

Data presented are means of three replicates ± the standard deviations.


Means (n = 3) within the same column with different superscript letters differ significantly (p b 0.05).
1180 S.Y. Seo et al. / International Journal of Biological Macromolecules 115 (2018) 1174–1182

at all the concentration were significantly higher than those of pectin indicating that viscous properties predominated over the elastic compo-
standard solutions, indicating that WSG solutions exhibited the greater nent G′. Whereas, WSG solution with concentration of 10% had elastic
viscoelastic properties compared to pectin standard solutions. The η* properties rather than viscous properties because G′ was higher than
values for WSG solutions were significantly higher than those for pectin G″ during all the aging time, representing a progressive reinforcement
standard solutions at the same concentration. This finding was consis- of the polymer networks. Similar trends were also observed in pectin
tent with our previous results on apparent viscosity where WSG solu- standard solutions. However, G′ and G″ values of WSG solutions were
tions exhibited the significantly higher apparent viscosity than pectin largely higher than those of pectin standard solutions throughout the
standard solutions (Table 3). entire time window. These results were probably due to the formation
The tan δ values of WSG solutions with concentration of 2, 4, 6, and of a more rigid polymeric network in WSG solutions compared to pectin
8% were in the range of 1.07–1.64, which demonstrated that the sample standard solutions [45]. Moreover, under the experimented conditions,
solutions exhibited mainly viscous properties instead of elastic proper- the magnitudes of G′ and G″ of WSG solutions were increased with an
ties. On the other hand, the tan δ values of WSG solutions with concen- increase in sample concentration, suggesting enhanced formation of
tration of 10 and 12% were 0.90 and 0.78, respectively, indicating the macromolecular network at higher concentration.
sample solutions were more elastic than viscous (Table 4). For pectin
standard solutions, only sample solution with concentration of 12%
exhibited elastic properties instead of viscous properties at 6.3 rad/s 3.8. Temperature sweep measurements
(tan δ = 0.94). In view of foregoing results, superior viscoelastic proper-
ties were confirmed in WSG solutions compared to pectin standard Changes in magnitudes of G′ and G″ during heating from 4 to 95 °C
solutions, reflecting the greater network development in WSG for pectin standard and WSG solutions with different concentration (8
solutions. The differences in macromolecular conformation between and 10%) are shown in Fig. 4. During the initial heating, G′ and G″ values
pectin standard and WSG might be a reason for the foregoing results decreased quickly as temperature increased, reaching a minimum value
because the conformation of macromolecule could contribute to the at 95 °C. The decrease in G′ can be related to the increase in fluidity with
inter- and/or intra-molecular interactions essential for strengthening increasing temperature. Also, this decrease may be attributed to the en-
polymer network [43]. Moreover, the viscoelastic properties of pectin ergy dissipation movement of the molecules and a decrease in intermo-
standard and WSG solutions were strongly influenced by the sample lecular interactions, which in turn decrease the energy needed for the
concentration. flow, thus decreasing the interference of the hydrodynamic domains.
The reduction in G′ values with increasing temperature is well-known
3.7. Time sweep measurements and usually explained by an increase in thermal activity of molecules,
resulting in an increase in molecule free volume and a simultaneous
Changes in G′ and G″ as a function of aging time for 1 h at 4 °C were decrease in intermolecular and/or intramolecular hydrogen bonding
measured for pectin standard and WSG solutions at different concentra- interactions [46,47].
tion (8 and 10%), as illustrated in Fig. 3. For both pectin standard and The G′ and G″ values of all samples steadily increased with decreas-
WSG solutions, the G′ and G″ values were not remarkably changed dur- ing temperature from 95 to 4 °C (Fig. 4). In all samples, the G′ values dra-
ing aging for 1 h. This phenomenon can support that WSG solutions as matically increased when temperature was decreased up to a critical
well as pectin standard solutions could be very stable during storage value. Such increase in G′ values can be attributed to the formation of
at refrigeration temperature (4 °C). For WSG solution with concentra- a three-dimensional network structure due to the strong interactions
tions of 8%, G″ values were higher than G′ during all the aging time, in the samples. Hesarinejad et al. [14] reported that the increase in G′
values was associated with the structural changes, such as formation
of hydrogen bonds, occurred during cooling of sample solutions.
The gelation temperature for a biopolymer is defined as the temper-
ature where the G′ value equals the G″ value (the sol-gel transition)
[48]. For WSG solutions with concentration of 8 and 10%, during heating
and cooling, G′ values were higher than G″ values in the whole range of
investigated temperatures. There was no cross-over point for G′ and G″,
which means WSG tended to exist in aggregated helical structures to
form three-dimensional networks. In other words, WSG solutions
with concentration of 8 and 10% showed a typical gel-like behavior.
On the other hand, for pectin standard solution with concentration
of 8%, G″ values were higher than G′ values during heating and cooling.
It means that 8% pectin standard solution tended to exist in a random
coil state, subsequently presenting the predominance of the viscous
character. In the case of pectin standard solution with concentration of
10%, the sample solution showed gel-like behavior in the beginning of
the heating step (G′ N G″). The G′ and G″ values decreased with an in-
crease in temperature and G′ values decreased much faster than G″
values. Then, a cross-over point between G′ and G″ values were ob-
served at around 20 °C (melting point). After the melting point, the vis-
cous behavior became predominant until 95 °C (G″ N G′), which means
that the dissociation of the aggregated helices happened, consequently
resulting in helix to coil transition (gel-sol transition). During cooling
from 95 to 4 °C for 10% pectin standard, the sol-gel transition occurred
at around 20 °C (gelation point) and the elastic behavior became pre-
dominant after gelation point. It seems that pectin standard solution
Fig. 3. Time evolution of storage (G′; open symbols) and loss (G″; filled symbols) moduli
with concentration of 10% might undergo a structural transformation
for pectin standard (○) and water soluble-sage seed gum (WSG, □) solutions at different from random coil to helical structure. Moreover, at lower temperature
concentration (8 and 10%, w/w) at 4 °C for 1 h. than gelation point, the side by side association of the helical structure
S.Y. Seo et al. / International Journal of Biological Macromolecules 115 (2018) 1174–1182 1181

Fig. 4. Temperature dependence of storage (G′; △) and loss (G″; ○) moduli for pectin standard (A) and water soluble-sage seed gum (WSG; B) solutions at different concentration (8 and
10%, w/w) during initial heating from 4 to 95 °C and subsequent cooling from 95 to 4 °C at a rate of 5 °C/min.

might occur to form three-dimensional network, as seen in the higher 57.77 and 55.80%, respectively, and were significantly higher than
G′ values than G″ values. those of pectin standard. In general, it has been widely known that
A thermal hysteresis was found for all concentration of pectin stan- many commercial polysaccharides, such as gums and pectins, include
dard and WSG solutions during the testing temperature range. More- a small amount of protein, either as a contaminant or as an intrinsic
over, thermal hysteresis area decreased when the concentration of part of their structure. These proteins can provide good emulsifying
sample solutions increased from 8 to 10%. The decreasing hysteresis ability to the polysaccharides, since those are usually strongly hydro-
area might be due to a denser and greater network formation in higher phobic. That is, the hydrophobic proteins in polysaccharides can adsorb
concentration of sample solutions. at oil-water interfaces to form stabilized layers around oil droplets [21].
Finally, during heating and cooling procedures, the dynamic moduli Due to their ability to act as the strong anchor at an oil-water interfacial
(G′ and G″) of WSG solutions were remarkably higher than those of pec- film, gums have often been used as ingredients to stabilize oil-in-water
tin standard solutions. According to Yaich et al. [49], the variation in dy- food emulsions [50]. Therefore, when compared to pectin standard, the
namic moduli was associated with the molecular weight. In other higher emulsion capacity and stability of WSG can be due to the higher
words, when polysaccharides were degraded, a decrease in the dynamic protein content (16.62%, Table 1) of WSG than that of pectin standard
moduli could be observed due to the reduced interactions between (1.47%, data are not shown).
small ions and charged chains in solutions. Furthermore, Yaich et al.
[49] described that the higher dynamic moduli might be caused by the 4. Conclusion
greater network formation in sample solution due to the existence of
expanded and high molecules which can lead to polymer-polymer asso- The water soluble gum (WSG) from sage (Salvia splendens) seed was
ciations. Accordingly, in the present study, higher dynamic moduli of isolated and its structural, physicochemical, and rheological properties
WSG compared to those of pectin standard could be associated with were studied. Molecular weight distribution analysis indicated that
greater molecular weight of WSG (Table 2). Also, the dynamic moduli WSG had slightly higher molecular weight than pectin standard. FTIR,
1
of pectin standard and WSG solutions increased with an increase in H NMR, and 13C NMR spectra confirmed successful extraction of WSG
concentration of sample solutions. It can be explained that the sample from sage seed gum. According to the results of steady shear rheological
solution with higher concentration has a denser network with more analysis, compared to commercial pectin standard, WSG solutions
interconnected chains. The denser network can retain the greater showed higher shear thinning behaviors, apparent viscosity (ηa,100),
amount of aqueous phase, resulting in formation of stronger gel [49]. and consistency index (K). Furthermore, the dynamic shear rheological
analyses indicated that WSG tended to form the more viscoelastic solu-
3.9. Emulsion capacity and emulsion stability tions compared to commercial pectin standard. Besides, we confirmed
that WSG had better emulsifying capacities than pectin standard. There-
The emulsion capacity and stability of pectin standard and WSG are fore, it is suggested in the present study that WSG could be used in food
shown in Table 2. The emulsion capacity and stability of WSG were industry as a potential surfactant, thickening, stabilizing, and gelling
1182 S.Y. Seo et al. / International Journal of Biological Macromolecules 115 (2018) 1174–1182

agents. In the future, the detailed structural characterization of WSG [25] Y. Liu, M. Qiang, Z. Sun, Y. Du, Optimization of ultrasonic extraction of polysaccha-
rides from Hovenia dulcis peduncles and their antioxidant potential, Int. J. Biol.
should be carried out to elucidate the structure-function relationship Macromol. 80 (2015) 350–357.
of this gum. [26] J. Zhao, F. Zhang, X. Liu, K.S. Ange, A. Zhang, Q. Li, R.J. Linhardt, Isolation of a lectin
binding rhamnogalacturonan-I containing pectic polysaccharide from pumpkin,
Carbohydr. Polym. 163 (2017) 330–336.
Acknowledgements [27] Y. Wu, S.W. Cui, J. Tang, Q. Wang, X. Gu, Preparation, partial characterization and
bioactivity of water-soluble polysaccharides from boat-fruited sterculia seeds,
This research was supported by Basic Science Research Program Carbohydr. Polym. 70 (4) (2007) 437–443.
[28] F. Jafari, F. Khodaiyan, H. Kiani, S.S. Hosseini, Pectin from carrot pomace: optimiza-
through the National Research Foundation of Korea (NRF) funded by tion of extraction and physicochemical properties, Carbohydr. Polym. 157 (2017)
the Ministry of Education (NRF-2015R1D1A1A01059564). 1315–1322.
[29] Y. Wang, F. Mao, X. Wei, Characterization and antioxidant activities of polysaccha-
rides from leaves, flowers and seeds of green tea, Carbohydr. Polym. 88 (1)
References
(2012) 146–153.
[30] G.D. Manrique, F.M. Lajolo, FT-IR spectroscopy as a tool for measuring degree of
[1] D.L. Wu, S.W. Hou, P.P. Qian, L.D. Sun, Y.C. Zhang, W.J. Li, Flower color chimera and
methyl esterification in pectins isolated from ripening papaya fruit, Postharvest
abnormal leaf mutants induced by 12C6+ heavy ions in Salvia splendens Ker-Gawl,
Biol. Technol. 25 (1) (2002) 99–107.
Sci. Hortic. 121 (4) (2009) 462–467.
[31] Y. Wu, L. Ai, J. Wu, S.W. Cui, Structural analysis of a pectic polysaccharide from boat-
[2] M.H. Li, J.M. Chen, Y. Peng, P.G. Xiao, Distribution of phenolic acids in chinese salvia
fruited sterculia seeds, Int. J. Biol. Macromol. 56 (2013) 76–82.
plants, World Sci. Technol. 10 (2008) 46–52.
[32] Z. Košťálová, Z. Hromádková, A. Ebringerová, Structural diversity of pectins isolated
[3] J. Kang, L. Li, D. Wang, H. Wang, C. Liu, B. Li, Y. Yan, L. Fang, G. Du, R. Chen, Isolation
from the Styrian oil-pumpkin (Cucurbita pepo var. styriaca) fruit, Carbohydr. Polym.
and bioactivity of diterpenoids from the roots of Salvia grandifolia, Phytochemistry
93 (1) (2013) 163–171.
116 (2015) 337–348.
[33] Z. Zhao, J. Li, X. Wu, H. Dai, X. Gao, M. Liu, P. Tu, Structures and immunological activ-
[4] S.M.A. Razavi, S.W. Cui, Q. Guo, H. Ding, Some physicochemical properties of sage
ities of two pectic polysaccharides from the fruits of Ziziphus jujuba Mill. cv.
(Salvia macrosiphon) seed gum, Food Hydrocoll. 35 (2014) 453–462.
jinsixiaozao Hort, Food Res. Int. 39 (8) (2006) 917–923.
[5] S.M. Razavi, A. Bostan, R. Rahbari, Computer image analysis and physico-mechanical
[34] W. Wang, X. Ma, P. Jiang, L. Hu, Z. Zhi, J. Chen, T. Ding, X. Ye, D. Liu, Characterization
properties of wild sage seed (Salvia macrosiphon), Int. J. Food Prop. 13 (2) (2010)
of pectin from grapefruit peel: a comparison of ultrasound-assisted and conven-
308–316.
tional heating extractions, Food Hydrocoll. 61 (2016) 730–739.
[6] F. Salehi, M. Kashaninejad, Effect of drying methods on rheological and textural
[35] E.I. Nep, S.M. Carnachan, N.C. Ngwuluka, V. Kontogiorgos, G.A. Morris, I.M. Sims, A.M.
properties, and color changes of wild sage seed gum, J. Food Sci. Technol. 52 (11)
Smith, Structural characterisation and rheological properties of a polysaccharide
(2015) 7361–7368.
from sesame leaves (Sesamum radiatum Schumach. & Thonn.), Carbohydr. Polym.
[7] A. Alghooneh, S.M.A. Razavi, F. Behrouzian, Rheological characterization of hydrocol-
152 (2016) 541–547.
loids interaction: a case study on sage seed gum-xanthan blends, Food Hydrocoll. 66
[36] Y.F. Zou, Y.P. Fu, X.F. Chen, I. Austarheim, K.T. Inngjerdingen, C. Huang, L.D. Eticha, X.
(2017) 206–215.
Song, L. Li, B. Feng, C.L. He, Z.Q. Yin, B.S. Paulsen, Purification and partial structural
[8] B. Vardhanabhuti, S. Ikeda, Isolation and characterization of hydrocolloids from
characterization of a complement fixating polysaccharide from rhizomes of
monoi (Cissampelos pareira) leaves, Food Hydrocoll. 20 (6) (2006) 885–891.
Ligusticum chuanxiong, Molecules 22 (2) (2017) 287–298.
[9] B. Abbastabar, M.H. Azizi, A. Adnani, S. Abbasi, Determining and modeling rheolog-
[37] D. Liu, Q. Sun, J. Xu, N. Li, J. Lin, S. Chen, F. Li, Purification, characterization, and bio-
ical characteristics of quince seed gum, Food Hydrocoll. 43 (2015) 259–264.
activities of a polysaccharide from mycelial fermentation of Bjerkandera fumosa,
[10] A. Haddarah, A. Bassal, A. Ismail, C. Gaiani, I. Ioannou, C. Charbonnel, T. Hamieh, M.
Carbohydr. Polym. 167 (2017) 115–122.
Ghoul, The structural characteristics and rheological properties of Lebanese locust
[38] F. Ma, Y. Zhang, Y. Yao, Y. Wen, W. Hu, J. Zhang, X. Liu, A.E. Bell, C. Tikkanen-
bean gum, J. Food Eng. 120 (2014) 204–214.
Kaukanen, Chemical components and emulsification properties of mucilage from
[11] Q. Wu, H. Qu, J. Jia, C. Kuang, Y. Wen, H. Yan, Z. Gui, Characterization, antioxidant
Dioscorea opposita Thunb, Food Chem. 228 (2017) 315–322.
and antitumor activities of polysaccharides from purple sweet potato, Carbohydr.
[39] E.R. Morris, Polysaccharide solution properties: origin, rheological characterization
Polym. 132 (2015) 31–40.
and implications for food systems, Front. Carbohydr. Res. 1 (1989) 132–163.
[12] A.H. Clark, S.B. Ross-Murphy, Structural and mechanical properties of biopolymer
[40] E.R. Morris, A.N. Cutler, S.B. Ross-Murphy, D.A. Rees, J. Price, Concentration and
gels, Adv. Polym. Sci. 83 (1987) 57–192.
shear rate dependence of viscosity in random coil polysaccharide solutions,
[13] Rao, M.A. 1999. Flow and functional models for rheological properties of fluid foods.
Carbohydr. Polym. 1 (1) (1981) 5–21.
In: Rheology of Fluid and Semisolid Foods: Principles and Applications (edited by
[41] M. Anvari, M. Tabarsa, R. Cao, S. You, H.S. Joyner, S. Behnam, M. Rezaei, Composi-
M.A. Rao). Gaithersburg: Aspen Publishers. Raton, F.L., USA: CRC Press. pp. 25–58.
tional characterization and rheological properties of an anionic gum from Alyssum
[14] M.A. Hesarinejad, A. Koocheki, S.M.A. Razavi, Dynamic rheological properties of
homolocarpum seeds, Food Hydrocoll. 52 (2016) 766–773.
Lepidium perfoliatum seed gum: effect of concentration, temperature and heating/
[42] P.J. Wood, Cereal β-glucans in diet and health, J. Cereal Sci. 46 (3) (2007) 230–238.
cooling rate, Food Hydrocoll. 35 (2014) 583–589.
[43] F.F. Simas-Tosin, R.R. Barraza, C.L.O. Petkowicz, J.L.M. Silveira, G.L. Sassaki, E.M.R.
[15] X. Huang, Y. Kakuda, W. Cui, Hydrocolloids in emulsions: particle size distribution
Santos, P.A.J. Gorin, M. Iacomini, Rheological and structural characteristics of
and interfacial activity, Food Hydrocoll. 15 (4) (2001) 533–542.
peach tree gum exudate, Food Hydrocoll. 24 (5) (2010) 486–493.
[16] J.P. Osano, S.H. Hosseini-Parvar, L. Matia-Merino, M. Golding, Emulsifying properties
[44] Y. Brummer, W. Cui, Q. Wang, Extraction, purification and physicochemical charac-
of a novel polysaccharide extracted from basil seed (Ocimum bacilicum L.): effect of
terization of fenugreek gum, Food Hydrocoll. 17 (3) (2003) 229–236.
polysaccharide and protein content, Food Hydrocoll. 37 (2014) 40–48.
[45] A.M. Sousa, J. Borges, A.F. Silva, M.P. Gonçalves, Influence of the extraction process
[17] A.R. Yousefi, S.M. Razavi, S.K. Aghdam, Influence of temperature, mono-and divalent
on the rheological and structural properties of agars, Carbohydr. Polym. 96 (1)
cations on dilute solution properties of sage seed gum, Int. J. Biol. Macromol. 67
(2013) 163–171.
(2014) 246–253.
[46] Y.F. Tang, Y.M. Du, X.W. Hu, X.W. Shi, J.F. Kennedy, Rheological characterisation of a
[18] AOAC International, Official Method of Analysis of AOAC (Method 925.09B, 923.03,
novel thermosensitive chitosan/poly (vinyl alcohol) blend hydrogel, Carbohydr.
979.09, 920.39C), 18th ed. Association of Official Analytical Chemists, Gaithersburg,
Polym. 67 (4) (2007) 491–499.
MD, USA, 2005.
[47] K. Bekkour, D. Sun-Waterhouse, S.S. Wadhwa, Rheological properties and cloud
[19] M.K. Dubois, J.K. Gilles, J.K. Hamilton, P.A. Rebers, F. Smith, Pfizer flocon 4800
point of aqueous carboxymethyl cellulose dispersions as modified by high or low
procedure-phenol/sulfuric acid method, Anal. Chem. 28 (1956) 350–352.
methoxyl pectin, Food Res. Int. 66 (2014) 247–256.
[20] N. Blumenkrantz, G. Asboe-Hansen, New method for quantitative determination of
[48] E. Flores-Huicochea, A.I. Rodríguez-Hernández, T. Espinosa-Solares, A. Tecante,
uronic acids, Anal. Biochem. 54 (2) (1973) 484–489.
Sol-gel transition temperatures of high acyl gellan with monovalent and diva-
[21] S.W. Cui, Y.H. Chang, Emulsifying and structural properties of pectin enzymatically
lent cations from rheological measurements, Food Hydrocoll. 31 (2) (2013)
extracted from pumpkin, LWT- Food Sci. Technol. 58 (2) (2014) 396–403.
299–305.
[22] Y.H. Chang, S.W. Cui, Steady and dynamic shear rheological properties of extrusion
[49] H. Yaich, H. Garna, S. Besbes, J.P. Barthélemy, M. Paquot, C. Blecker, H. Attia, Impact
modified fenugreek gum solutions, Food Sci. Biotechnol. 20 (6) (2011) 1663–1668.
of extraction procedures on the chemical, rheological and textural properties of
[23] Y.H. Chang, S.W. Cui, K.T. Roberts, P.K.W. Ng, Q. Wang, Evaluation of extrusion-
ulvan from Ulva lactuca of Tunisia coast, Food Hydrocoll. 40 (2014) 53–63.
modified fenugreek gum, Food Hydrocoll. 25 (5) (2011) 1296–1301.
[50] S. Naji, S.M. Razavi, H. Karazhiyan, Effect of thermal treatments on functional prop-
[24] Y. Chen, H. Zhang, Y. Wang, S. Nie, C. Li, M. Xie, Sulfated modification of the polysac-
erties of cress seed (Lepidium sativum) and xanthan gums: a comparative study,
charides from Ganoderma atrum and their antioxidant and immunomodulating ac-
Food Hydrocoll. 28 (1) (2012) 75–81.
tivities, Food Chem. 186 (2015) 231–238.

You might also like