You are on page 1of 9

International Journal of Biological Macromolecules 128 (2019) 629–637

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules

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

Physicochemical and rheological characterization of pectin-rich fraction


from blueberry (Vaccinium ashei) wine pomace
Liyuan Feng a,d,1, Yun Zhou a,c,1, Tolulope Joshua Ashaolu a, Fayin Ye a, Guohua Zhao a,b,⁎
a
College of Food Science, Southwest University, Chongqing 400715, PR China
b
Chongqing Engineering Research Center of Regional Foods, Chongqing 400715, PR China
c
Guangxi Key Laboratory of Agricultural Resources Chemistry and Biotechnology, Yulin 537000, PR China
d
School of Resources and Environment Sciences, Baoshan University, Baoshan 678000, PR China

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

Article history: Blueberry wine pomace is rich in pectin, which has been extensively used as a functional dietary fiber and a rhe-
Received 8 October 2018 ological modifier in the food industry. This paper reports a method to extract the pectin-rich fraction (PF) with a
Received in revised form 22 January 2019 mediate degree of esterification of 51.66% from blueberry wine pomace and provides insight into its relationship
Accepted 28 January 2019
between the structure and rheological properties. The impacts of related extrinsic factors, such as sucrose, ions
Available online 29 January 2019
and pH, were also studied in view of food applications. The viscosity of PF aqueous dispersion gradually increased
Keywords:
with its concentration. The addition of sucrose, CaCl2 or NaCl to the solution resulted in increased viscosity. How-
Pectin ever, the elevations in temperature and pH led to decrease in solution viscosity. The viscoelastic property of PF
Blueberry dispersion displayed strong temperature dependence but weak frequency dependence. This was largely due to
Rehological properties PF concentration, sucrose, CaCl2 and solution pH. The present study revealed the unique characteristic of
medium-methoxylated pectin fraction and the obtained results are helpful in value-added utilization of blue-
berry wine pomace.
© 2019 Elsevier B.V. All rights reserved.

1. Introduction
Innovative ways are sought to utilize berry pomace in food process-
Blueberry (Vaccinium spp.), a very popular berry fruit, is widely con- ing to minimize the amount of wastes, thereby generating a new food
sumed as fresh fruit and in processed products such as fruity yogurt, jam source. Press residues of red/green currant and sea buckthorn berry
and beverages. The health significance of blueberry is attributed to the could be used as stabilizers and food supplements [4]. Swamp cranberry
presence of a diverse range of phytochemicals, such as anthocyanins, pomace extracts could assist in the antimicrobial protection of minced
flavonols and chlorogenic acids and pectin-rich dietary fibers. On the pork meat [5] and inhibit unfavorable storage changes of lipids and
average, the pectin content in fresh blueberry fruits is up to 4 g/kg [1]. muscle pigments as food additives for meat products [6]. Incorporation
Blueberry is very suitable for wine processing. During wine process- of dried black currant pomace in wheat dough is a possible way to in-
ing, the fruits are pressed to release the sweet juice used for fermenta- crease the water absorption and development time of dough [7]. A pre-
tion along with their skins and seeds. Similar to grapes, anthocyanins liminary study on the blueberry wine pomace showed that the
of blueberries are present mainly in the skins, while tannins and pheno- phenolics extracts obtained by supercritical CO2 fluid extraction exhib-
lic acids are abundantly predominant in the seeds [2]. As fermentation ited a higher antioxidant activity than that recorded by the conventional
proceeds, these compounds are transferred into wine contributing to methanol extraction [8]. A pressurized liquid extraction method was
the color, mouthfeel, and antioxidant properties of wine. The skin of used to recover phenolic compounds, antioxidants and anthocyanins
blueberries are abundant in dietary fibers and other bioactive com- from blueberry residues [9].
pounds [3]. However, the residues of wine processing, the blueberry The blueberry wine pomace is rich in dietary fibers, including
pomace, which accounts for approximately 60% of the total waste, is a 5.12 g/100 g soluble dietary fiber and 77.74 g/100 g insoluble dietary
low-value by-product and usually discarded. fiber (Supplementary material 1). The chemical composition of dietary
fiber was also investigated, indicating that there was 3.33 wt% pectin,
8.38 wt% cellulose and 20.91 wt% hemicellulose. The content of free
⁎ Corresponding author at: College of Food Science, Southwest University, 2 Tiansheng
Road, Chongqing 400715, PR China.
phenol and bound phenol of blueberry wine pomace was 150.83 mg
E-mail address: zhaogh@swu.edu.cn (G. Zhao). and 252.73 mg of garlic acid equivalents per 100 g extract, respectively,
1
These authors contributed equally to this work. on a dry-weight basis. It can be observed that blueberry wine pomace is

https://doi.org/10.1016/j.ijbiomac.2019.01.166
0141-8130/© 2019 Elsevier B.V. All rights reserved.
630 L. Feng et al. / International Journal of Biological Macromolecules 128 (2019) 629–637

rich in phenol compounds, especially the bound ones. Regenerating di- was colorimetrically determined as 65.88 (wt%) by using a meta-
etary fibers from blueberry wine pomace could improve the commercial phenyldiphenol assay.
values of wine wastes and the utilization rate of related functional com-
ponents of blueberry. 2.3. Physicochemical properties
Pectins, a heterogeneous polysaccharide found in the primary cell
wall and the middle lamella, has been extensively used as a rheology 2.3.1. Galacturonic acid (GalA) content
modifier in the food industry. Depending on the degree of esterification The content of Gal-A was determined by the colorimetrically meta-
(DE), pectin can be divided into high methoxyl pectin (HMP) with DE phenyldiphenol assay at 525 nm, using GalA as the standard and
N 50% (N7% methoxyl), and low methoxyl pectin (LMP) with DE b 50% NaOH solution as the blank group [15].
(3%–6% methoxyl) [10,11]. DE value of pectin could determine its rheo-
logical properties and gel characteristics of pectin-incorporated food 2.3.2. Monosaccharide composition
matrix, along with other intrinsic molecular parameters including mo- Monosaccharide composition analysis was performed as described
lecular weight, monosaccharide composition, and extrinsic conditions by Zhang et al. [16]. Briefly, 2 mg pectin was hydrolyzed with 2 mL of
like temperature and pH [12]. The concentration of pectin and co- 2 M TFA at 110 °C for 3 h. The monosaccharides were derivatized by
solutes could also influence the viscoelasticity of pectin solution and the 1-phenyl-3-methyl-5-pyrazolone (PMP) method. The derivatives
gels by influencing intermolecular interactions such as hydrogen bond- were analyzed by HPLC on a Thermo BDS-C18 column (250 × 4.6 mm
ing, electrostatic and hydrophobic interactions. i.d, 5 μm) with a UV detector (Shimadzu, Tokyo, Japan) at 30 °C. 20 μL
Commercial pectin are mostly obtained from citrus peel and apple of PMP derivative were injected, eluted with continuous gradient elu-
pomace. Pectins extracted from various plant species show different tion mobile phase (A phase: 15% (v/v) acetonitrile + 0.05 mol/L
physicochemical properties. Gels formed by pectins extracted from KH2PO4-NaOH buffer, pH 7.1; B phase: 40% (v/v) acetonitrile + 0.05
berry press residues by microwave-assisted extraction were weaker in mol/L KH2PO4-NaOH buffer, pH 7.1) at a flow rate of 0.7 mL/min and
gel strength than that of commercial citrus pectin, but stronger than monitored by UV absorbance at 250 nm. Elution time stage is 0 min–
that of pectin extracted from apple [13]. In this paper, pectin-rich con- 10 min-40 min-50 min-57 min, corresponding to the B concentration
centrate from blueberry wine pomace was obtained by acid-soluble al- gradient change being 0%—10%—30%—0%—0%. The correction factor is
cohol precipitation, and its physicochemical characteristics were defined by the equation of f(Lactose) = (Alac × Cmon)/(Amon × Clac),
analyzed. The influences of some extrinsic factors on the thickness and where f (Lactose) is the correction factor. Alac and Amon are the peak
gel-forming properties of pectin were explored in detail. areas for lactose and a standard monosaccharide, respectively. Clac and
Cmon are the values of concentration (mmol/L) of lactose and a standard
monosaccharide, respectively.
2. Materials and methods
The linear standard curve equations are expressed by the concentra-
tion (mmol/L) of standard monosaccharide as x, while the ratio value of
2.1. Materials
peak area for corresponding monosaccharide of tested samples and lac-
tose as y.
Fresh blueberry wine pomace (BWP) was provided by Tiantaishan
Blueberry Wine Co., Ltd. (Chongqing, China). The pomace was immedi-
2.3.3. Molecular weight
ately dried at 50 °C in a thin layer in the oven for 48 h. The moisture con-
The molecular weight was measured by high-performance size-
tent of the dried pomace was determined as 4.30 g/100 g (wet basis).
exclusion chromatography (HPSEC) on a TSK G4000 PWXL-SEC column
The contents of ash, crude protein, fat, total fiber, and pectin were
(7.8 × 300 mm i.d.) [17]. Pectin-rich fraction (0.5 mg/mL) was dissolved
0.47 g/100 g, 4.46 g/100 g, 5.54 g/100 g, 82.86 g/100 g and
in deionized water and filtered through a 0.45 μm cellulose acetate
3.33 g/100 g (dry basis), respectively. Analytical grade chemicals were
membrane. 20 μL of sample was injected, eluted with 0.02% NaN3 (w/
purchased from Chengdu Kelong Chemical Reagent Co., Ltd. (Chengdu,
v) at a flow rate of 0.6 mL/min and monitored using a refractive index
China). Galacturonic acid standard (97%) for galacturonic acid content
RID-10A detector (Shimadzu, Tokyo, Japan). Dextran of molecular
determination was purchased from Sigma-Aldrich Co., Ltd. (USA). The
weight of 5, 50, 150, 670, 110 kDa were used for building the standard
standard substances of molecular weight (5, 50, 150, 670 and
curve. The linear standard curve was expressed by the equation of y
110 kDa) and monosaccharides (GalA, Man, Ara, Glu, Rha, Fuc, GluA
= −0.4497x +10.11, R2 = 0.9986, where x represents retention time,
and Gal) were abtained from Sigma-Aldrich Co., Ltd. (USA) and
y as log Mw.
YIFANG S & T Co., Ltd. (Tianjin, China) for monosaccharide composition
and molecular weight characterization, respectively. All the HPLC-grade
2.4. Rheological properties
reagents were purchased from Tjshield Co., Ltd. (Hebei, China).
All rheological measurements were carried out on a DHR rheometer
2.2. Extraction of pectin-rich fraction (TA Instruments, USA). The parallel plate geometry (60 mm diameter)
was used for both steady shear and oscillatory tests. The prepared sam-
The pectin-rich fraction was extracted according to Fig. 1, which is a ples were equilibrated at 25 °C for 1 h before being loaded to the rheom-
modified protocol of Pasandide et al. [14]. Approximately 100 g blue- eter plate with a spoon. Prior to each measurement, sample was left for
berry wine pomace was dried at 5 °C for 24 h until the attainment of a 10 min to recover the structural breakdown upon loading. The exposed
constant weight. The dried blueberry wine pomace was grounded and sample edge was covered with a thin layer of silicon oil in order to pre-
milled through a 0.15 mm sieve. Hydrochloric acid solution was added vent moisture evaporation.
to the blueberry wine pomace powder (pH = 1, 1:15 g/mL) and stirred
for 2 h at 85 °C. The resulting slurries were centrifuged at 4000 rpm for 2.4.1. Flow behavior
20 min to remove sediments. The supernatants were filtered and pre- In steady shear measurements, pectin dispersions with varying con-
cipitated with absolute ethanol (1:1, v/v) for 12 h at ambient tempera- centrations were prepared by suspending PF in distilled water with con-
ture. The ethanol extraction was repeated twice. The pectin tinuous magnetic stirring for 2 h at room temperature. The effects of PF
concentrates were filtered and washed successively with 80%, 95% and concentration (0–2.5%, w/w), CaCl2 (0–81.0 mM), NaCl (0–0.7%, w/w),
100% ethanol (1:2, v/v). The pectin-rich fraction was dried by vacuum sucrose (5–20%, w/w), temperature (25–100 °C) and pH (2.9–10.0)
lyophilization for further use. The extraction yield was estimated to be were investigated. The steady shear flow analyses were performed at
1.62 wt% on a dry basis and the pectin concentration in the fraction 25 °C via a ramp of shear rate from 18 s−1 to 1000 s−1in 10 min. The
L. Feng et al. / International Journal of Biological Macromolecules 128 (2019) 629–637 631

Fig. 1. Schematic flow chart of the extraction of pectin-rich fraction (PF) from blueberry wine pomace (BWP).

results were fitted with Power law models: τ = K × γ_ n, where τ is the investigated. In preparing samples for oscillatory tests, PF and sucrose
shear stress (Pa), γ_ is shear rate (s−1), K is the consistency index were first dissolved in 0.1 M citric acid sodium buffer with a specific
(Pa·sn) and n is the dimensionless flow behavior index. pH. Then the dispersions were heated until boiling and thoroughly
mixed with a portion of CaCl2 solution to result in a pre-settled concen-
tration. The resultant dispersion was then cooled to room temperature
2.4.2. Viscoelastic behavior and rested for 24 h at 4 °C prior to measurements. Temperature
In visoelastic properties of PF gels were investigated by temperature sweep was performed from temperature decreasing from 85 °C to 25
sweep and frequency tests. The effects of PF concentrations (0–2.5%, w/ °C at the rate of 5 °C/min at 1 Hz and 0.5% strain. For frequency sweep,
w), CaCl2 (0–81.0 mM), sucrose (10–50%, w/w) and pH (3.0–7.0) were it was operated from 0.1 to 20 Hz at 5 °C. All oscillatory measurements
632 L. Feng et al. / International Journal of Biological Macromolecules 128 (2019) 629–637

were carried out at a strain % of 0.5% within the linear viscosity region. methoxyl pectin, since its DE was very close to 50%. It is traditionally
The traces of G′ and G″ as a function of temperature or frequency were recognized that soluble solid can facilitate the gelation of high methoxyl
obtained. pectin by reducing water activity and thus increasing the size and num-
ber of cross-linking zones. For a specific pectin solution, a decrease in pH
2.5. Statistical analysis not only enhances the hydrogen bonding between water molecules and
the carboxyl groups of pectin molecules but also weakens the electro-
The level of statistically significant difference was set at p b 0.05 by static repulsion between pectin molecules. Regarding the gelation of
Duncan's procedure using the SPSS statistics 16. All experiments were low methoxyl pectin, it usually requires the presence of metal cations,
done in triplicate and results were expressed as mean ± standard especially divalent calcium ions. In this case, one calcium ion can elec-
deviation. trostatically interact with two negatively charged carboxyl groups of
pectin molecules and this ultimately triggers the formation of a three-
3. Results and discussion dimensional mesh structure called “egg box” [23]. Therefore, the factors
influencing both high methoxyl pectin and low methoxyl pectin should
3.1. Characterization of pectin-rich fraction (PF) be taken into account in evaluating the functionality of PF.
The monosaccharide composition of PF was analyzed by a PMP-pre-
The chemical characteristics of PF are summarized in Table 1. In column derivatization HPLC method. The results revealed that PF con-
order to evaluate the purity of PF, hydroxybiphenyl method was used tains seven monosaccharides, namely GalA, Man, Ara, Glu, Rha, Fuc
to determine its galacturonic acid (GalA) content. The results showed and GluA in a decreasing molar ratio order (Supplementary material
that GalA content (65.88 wt%) in PF extracted from blueberry wine 3). GalA was the most abundant among these monosaccharides, ac-
pomace was slightly higher than the minimum value (≥65 wt%) re- counting for 44.14 mol% while Man is the richest neutral monosaccha-
quired by Food Chemicals Codex (FCC) for commercial pectin products. ride, although it is relatively low in gold kiwifruit pectins [22]. It is
However, the present GalA content was higher than that of the pectin worth noting that galactose (Gal) could not be found in the present PF
from whole blueberry power (51.6 wt%) [18]. samples, although it was present in other pectin-rich pomaces of vari-
Generally, high-performance size-exclusion chromatography was ous fruits sources [22]. This shows that the monosacchride profile of a
used to characterize the molecular weight distribution of various poly- pectin matter certainly depends on its plant origin [24].
saccharides. For the molecular weight distribution pattern of the pres- Rha/GalA value was usually applied to identify the structure of pec-
ent PF, a dominant big peak as well as several small peaks were tin polymers. The present PF was concluded with a Rha/GalA value of
observed in the resultant eluting curve (Supplementary material 2). 0.03 (b0.05), indicating it belongs to HG and RG-II groups [25]. On the
The weight-average molecular weigh (Mw), number-average molecular other hand, the value of (Gal + Ara)/Rha always reflect the extension
weigh (Mn) and polydispersity index (PDI, Mw/Mn) of PF were 641 kDa, of the neutral side chains in a pectin matter [26]. The present PF showed
460 kDa and 1.39, respectively (Table 1). These data were comparable to a relatively low (Gal + Ara)/Rha value (10.45) suggesting the shortness
those reported for the pectin extracted from whole blueberry powder of its side chains of RG-I.
[18]. It was also observed that Mw of PF was higher than that of commer-
cial pectins such as citrus pectin (100–200 kDa) and sugar beet pectin
(400 kDa) [19–21], but lower than that of the pectin from gold kiwifruit 3.2. Flow behaviors of PF solutions
pomace (670 kDa) [22].
As determined by FCC titration method, the degree of esterification Almost all polymer solutions present shear thinning behavior, i.e.
(DE) of PF was 51.66%, which was much higher than that (26%–36%) apparent viscosity decreases as shear rate increases. This behavior is
of the pectins extracted from blueberry by different methods [18]. This usually interpreted as a consequence of insufficient polymer chain en-
indicated the pectin from blueberry wine pomace could be classified tanglements. However, in most cases of pectin-containing samples,
as high methoxyl pectin. However, it could be expected that the present co-solutes and environmental factors such as sugar concentration and
pectin fraction might show some specific properties or functions of low temperature delivered substantial impacts on the rheological properties
of their solutions. As shown in Fig. 2, PF solutions under different condi-
Table 1
tions clearly exhibited a shear thinning behavior, signifying that the
Chemical characterization of pectin-rich fraction from blueberry wine samples were pseudoplastic. However, the situations were more or
pomace. less different among the cases with different co-solutes and environ-
mental parameters. For the cases on PF concentration, sucrose, NaCl,
Character Value
temperature and pH, the shear thinning behavior was mainly observed
Galacturonic acid (wt%)a 65.88 ± 0.06
in low shear rate scope (b30 s−1). However, continuous and significant
Mw (×105 Da) 6.41 ± 0.18
Mn (×105 Da) 4.60 ± 0.21 shear thinning behaviors were observed for CaCl2 cases across the
Mw/Mn (PDI) 1.39 ± 0.11 whole shear rate scope.
DE (%)b 51.66 ± 0.05 The steady rheological characteristics of PF solutions were described
Monosaccharides (mol %) by fitting the shear rate and shear stress to power law equation and the
Galacturonic acid (GalA) 44.14 ± 2.63 results are shown in Table 2. When PF concentration is under consider-
Mannose (Man) 33.59 ± 3.72 ation, the results clearly revealed that the higher PF concentration was
Arabinose (Ara) 16.11 ± 1.73 inevitably accompanied with a higher value in consistency coefficient,
Glucose (Glu) 3.22 ± 0.26
K. The K value increased from 0.872 to 4.777 as PF concentration varied
Rhamnose (Rha) 1.54 ± 0.11
Fructose (Fuc) 0.84 ± 1.12 from 0.1 wt% to 2.5 wt% (Fig. 2a). At the same time, the n value de-
Glucuronic acid (GluA) 0.58 ± 0.12 creases from 1.008 to 0.967. This attested that PF solutions become
Rha/GalA 0.03 ± 0.00 more non-Newtonian as pF concentration increased and the present
(Gal + Ara)/Rha 10.45 ± 0.37
PF system shows the typical behavior of polymer-rich systems. With
HGc 42.60 ± 2.52
RG-Id 19.18 ± 1.95 the aid of statistical analysis, it was found that the K value of the tested
a
systems is linearly correlated to PF concentration (R = 0.9863). This fact
Based on dry matter.
b
DE refers to the degree of esterification.
suggests that the present PF solutions behave as diluted solutions. In
c
HG = GalA-Rha. other words, even at a concentration of 2.5 wt%, the present PF could
d
In this case, RG-I = 2Rha + Gal + Ara. not form a highly viscous liquid via molecular entanglements.
L. Feng et al. / International Journal of Biological Macromolecules 128 (2019) 629–637 633

Fig. 2. The steady shear flow curves of the dispersion of pectin-rich fraction (PF) from blueberry wine pomace as affected by substrate concentration (a, natural pH), sucrose (b, PF 1.5%,
pH 2.85), CaCl2 (c, PF 1.5%, pH 2.85), NaCl (d, PF 1.5%, pH 2.85), temperature (e, PF 1.5%, pH 2.85) and pH (f, PF 1.5%, pH 2.85).

Regarding the impacts of sucrose, the experiments were carried out (b30 s−1) or high (N30 s−1) shear rate scope, the resultant amplitudes
at a PF concentration of 1.5 wt%. The obtained data showed that a signif- of shear thinning gradually weakened as the sucrose concentration in-
icant increasing tendency in K value (2.85 → 3.79) but a weak increasing creased. At the end of shear, increase in the sucrose concentration of
tendency in n value (0.976 → 0.999) were concluded as the sucrose in the solution led to increased final viscosity. A previous study revealed
the solutions increased from 0 wt% to 20 wt%. In details, either in low that the presence of sucrose could accelerate preferential hydration

Table 2
Rheological parameters of solutions of PF based on Power Law equation.

Systema τ = K × γ_ n

K (mPa·sn) n R2

PF 0.1% 0.87 ± 0.12 1.008 ± 0.027 0.9986


PF 0.5% 1.23 ± 0.17 0.994 ± 0.013 0.9995
PF 1.0% 1.80 ± 0.06 0.979 ± 0.024 0.9976
PF 1.5% 2.85 ± 0.11 0.976 ± 0.005 0.9997
PF 2.0% 3.41 ± 0.40 0.978 ± 0.008 0.9999
PF 2.5% 4.78 ± 0.28 0.967 ± 0.008 0.9999
PF 1.5%-Sucrose 5% 2.92 ± 0.07 0.983 ± 0.005 0.9998
PF 1.5%-Sucrose 10% 3.18 ± 0.08 0.988 ± 0.005 0.9999
PF 1.5%-Sucrose 15% 3.52 ± 0.02 0.990 ± 0.001 0.9999
PF 1.5%-Sucrose 20% 3.79 ± 0.06 0.999 ± 0.006 0.9999
PF 1.5%-CaCl2 13.5 mM 141.6 ± 10.5 0.524 ± 0.011 0.9964
PF 1.5%- CaCl2 27.0 mM 86.3 ± 2.40 0.598 ± 0.004 0.9981
PF 1.5%- CaCl2 40.5 mM 61.2 ± 5.03 0.641 ± 0.010 0.9990
PF 1.5%- CaCl2 54.0 mM 45.9 ± 4.71 0.673 ± 0.013 0.9984
PF 1.5%- CaCl2 67.5 mM 41.6 ± 4.23 0.680 ± 0.012 0.9987
PF 1.5%- CaCl2 81.0 mM 46.4 ± 7.85 0.671 ± 0.023 0.9991
PF 1.5%-NaCl 0.1% 2.79 ± 0.06 0.982 ± 0.005 0.9997
PF 1.5%-NaCl 0.3% 3.58 ± 0.11 0.960 ± 0.005 0.9999
PF 1.5%-NaCl 0.5% 4.69 ± 0.20 0.928 ± 0.011 0.9999
PF 1.5%-NaCl 0.7% 5.89 ± 0.11 0.900 ± 0.002 0.9999
PF 1.5%-40 °C 3.03 ± 0.25 0.971 ± 0.008 0.9997
PF 1.5%-60 °C 2.60 ± 0.06 0.983 ± 0.006 0.9999
PF 1.5%-80 °C 2.59 ± 0.09 0.989 ± 0.000 0.9998
PF 1.5%-100 °C 2.42 ± 0.08 0.994 ± 0.004 0.9997
PF 1.5%-pH 2.85 2.85 ± 0.11 0.976 ± 0.005 0.9997
PF 1.5%-pH 4.07 2.88 ± 0.06 0.975 ± 0.005 0.9997
PF 1.5%-pH 6.01 2.58 ± 0.10 0.968 ± 0.011 0.9996
PF 1.5%-pH 8.10 2.30 ± 0.02 0.980 ± 0.001 0.9997
PF 1.5%-pH 10.04 2.40 ± 0.00 0.983 ± 0.000 0.9998
a
The concentrations of PF and sucrose were in wt%.
634 L. Feng et al. / International Journal of Biological Macromolecules 128 (2019) 629–637

[27], by which more water molecules bind to sucrose via hydrogen pH 3.0 [35]. The effects of media pH on the viscosity of PF solution were
bonds [28]. This promoted the crosslinking formation among polymer demonstrated in Fig. 2 f. Obviously, there were no significant difference
chains and gave rise to increased viscosity of high methoxyl pectin solu- between pH 2.85 and pH 4.07 across the whole shear rate range. How-
tions [29]. Simultaneously, the hydrophobic interactions among pectin ever, lower viscosities were observed for the pectin solutions at pH 6, 8
chains were promoted with sucrose incorporation and thus again pro- and 10. The present result is in agreement with the results reported for
mote crosslinking within polymer chains [30]. sunflower head pectin, which presented the highest viscosity at pH 3
Calcium ion (Ca2+) can generate a bridge between the two carboxyl [36,37]. At pH higher than 4.5, the galacturonic acid molecules, forming
groups of pectin molecules and thus trigger the formation of a gel net- the pectin chains, were fully charged leading to a high electrostatic repul-
work structure with a semi-stable configuration between pectin chains sion and consequent viscosity reduction [36,38,39]. On the contrary, at pH
and Ca2+, known as “egg box” model [31]. As illustrated in Fig. 2c, the lower than 3.5, the\\COOH groups are available for hydrogen bonding
addition of CaCl2 at any concentration (13.5–81.0 mM) brought about thus increasing the viscosity value [40].
a substantial increase in the viscosity of pectin solution. However, the
resultant amplitude in viscosity does not linearly depend on the concen-
tration of CaCl2 applied. Specifically, the CaCl2 addition at 13.5 mM pro- 3.3. Viscoelastic behavior of PF solutions
duced the largest amplitude in viscosity changes. With the further
increase in CaCl2 concentration from 13.5 mM to 54.0 mM, the resultant 3.3.1. Temperature sweep
viscosity changes decreased gradually. No significant changes in viscos- The evolution of viscoelastic moduli against temperature was ob-
ity were observed upon further increase of CaCl2 concentration from tained through a temperature sweep test from 85 °C to 25 °C, at 5 °C/
54.0 mM to 81.0 mM. The significant viscosity increase at 13.5 mM min and a constant frequency of 0.5 Hz. This was done to investigate
should be ascribed to the crosslinking of pectin molecules via calcium the phase transition occurring in the systems. The temperature-
bridge. However, as previously pointed out that excessive Ca2+ will induced phase transition was ultimately a consequence of the effects
bring about the agglomeration of pectin molecules, thus weakening of temperature on the intermolecular interactions involved in pectin
the viscosity [11,32]. These results convey that the present PF indeed molecules. With temperature increase, the intermolecular hydrophobic
showed an ability to chelate divalent cation. This was assumed as a con- interactions occurring among the methoxyl groups of pectin molecules
sequence of a certain number of consecutive free carboxyl groups, as were enhanced, but the hydrogen bonding among the hydrophilic car-
available frequently in low methoxyl pectin. For high methoxyl pectin, boxyl and hydroxyl groups of pectin molecules grew weak. Upon the
it also can form calcium bridges upon the demethoxylation achieved cooling, the hydrophobic interactions were loosened and even cleaved
by using pectin methyl esterases [33]. but the hydrogen bonding was strengthened. The results showed that
It has been proven that pectin, an anionic biopolymer, tends to form both of the storage modulus (G′) and loss modulus (G″) increased as
aggregates upon salts addition, being responsible for the increase in vis- the temperature decreased (Fig. 3). The crossover point of two moduli
cosity of the pectin involved system [34]. The natural pH of 1.5 wt% PF (G′ and G″) indicated the occurrence of phase transition in the system,
solution was determined as 2.85 in the present study. Under these cir- namely from liquid-like (high temperature) to solid-like (low tempera-
cumstances, the chains of PF polymers would be negatively charged at ture) or vice versus.
a low density [35]. Upon the addition of NaCl over 0.3 wt%, the resultant Regarding the effects of PF concentration on its viscoelastic behavior,
increase in ion strength of the system significantly screened the electro- the results in Fig. 3a showed that, regardless of the PF concentration,
static repulsion among pectin molecules. This brought the adjacent both moduli gradually decreased upon the cooling from 85 °C to 25 °C.
polymer chains closer to each other and, consequently, strengthened In contrast to 0.5 wt% PF solution, the bigger changes in both moduli
the polymer-based three-dimensional network within the system. were concluded for 1.5 wt% and 2.5 wt% PF solutions. For a specific mod-
This was ultimately confirmed by a significantly elevated viscosity of ulus either G′ or G″, its amplitude increased as PF concentration in-
the system at the end of shear (Fig. 2d). Therefore, the addition of creased. This is consistent with the previous observation in flow
salts like NaCl to a pectin solution manifested an obvious viscosity en- behavior measurements that a higher apparent viscosity could be ex-
hancement via the partial shielding of electrostatic repulsive forces as pected by increasing the concentration of PF. At high polymer concen-
well as the significant enhancement of hydrogen bonding and trations, the side chains of pectin molecules play a key role in further
rearranging of polymer chains [10]. tightening the network through entanglements and topological con-
To investigate the effects of heating treatment on the rheological straints as those explained in “reptation model” [41]. Moreover, the
properties of PF solution, an 1.5 wt% PF solution was preprepared and temperature corresponding to the cross of the traces of G′ and G″ was
incubated at different temperatures of 25 °C, 40 °C, 60 °C, 80 °C and highly depended on PF concentration. Approximately, the cross of the
100 °C, respectively, for 20 min before viscosity measurement. The re- traces of two moduli occurred at 35 °C, 60 °C and 85 °C for the PF solu-
sults were presented in Fig. 2e. Normally, the heating treatment cer- tions at 0.5 wt%, 1.5 wt% and 2.5 wt%, respectively. This suggested that a
tainly affected the polymer conformation and the intermolecular more concentrated PF solution could form a gel at a higher temperature.
interactions, such as hydrogen bonds, hydrophobic association, electro- The effects of sucrose concentration on the viscoelastic behavior of
static repulsion and Van der Waals forces [36]. This finally resulted in PF solution were illustrated in Fig. 3b. The crossover point of two moduli
more or less changes in the viscosity of pectin solution. Clearly, the for PF solutions containing 10%, 30% and 50% sucrose occurred at 42.6 °C,
heat treatment at 40 °C caused a slight increase in viscosity, but the 72.1 °C, and 79.4 °C, respectively. Obviously, the lower values in both
treatments at 60 °C, 80 °C and 100 °C brought minor declines in viscos- moduli of the system containing 10% sucrose indicated its very soft tex-
ity. For the viscosity increase induced by the heat treatment at 40 °C, it ture nature. The critical concentration of sucrose for the cold-set gela-
was assumed as a consequence of the enhanced pectin solubilization. In tion of pectin is about 50–55 wt% [29,42], allowing the necessary
other words, PF did not dissolve completely without heat treatment and contact of the pectin chains for junction zones formed via hydrophobic
the heat treatment increased the actual pectin concentration in the interactions and hydrogen bonds. In the early stage of cooling, both G′
tested system [37]. However, for the viscosity decrease induced by the and G″ of 50% sucrose system were sharply raised to 7.12 Pa (72.95
heat treatments over 60 °C, they could be ascribed to the heat-induced °C) and 2.83 Pa (76.72 °C). In the subsequent cooling, they increased
damages on the intermolecular interactions of pectin molecules, thus in less degrees. Generally, the cold set gels from high methoxyl pectin
weakening the three-dimensional network within the pectin solution. are thermo-irreversible, which attested that the gel structure would
Generally, pH is crucial to the charge distribution of the chains of not be re-built during annealing. As observed for konjac glucomannan,
pectin molecules. The pKa value of the carboxyl groups of pectin mole- the conformational transformation and intermolecular aggregations
cules is about 3.5 and approximately 50% of them were protonated at are responsible for its thermo-irreversible gelation [43].
L. Feng et al. / International Journal of Biological Macromolecules 128 (2019) 629–637 635

Fig. 3. Evolution of storage modulus G′ (filled symbols) and loss modulus G″ (open symbols) of the dispersion of pectin-rich fraction (PF) from blueberry wine pomace with temperature
(from 85 °C to 25 °C at a rate of 10 °C/min) as affected by PF concentration (a, 20 mM CaCl2, 30% sucrose, pH 4.0), sucrose (b, 20 mM CaCl2, 1.5% PF, pH 4.0), pH (c, 1.5% PF, 20 mM CaCl2, 30%
sucrose) and CaCl2 (d, 1.5% PF, 30% sucrose, pH 4.0). The determinations were performed at 1 Hz and 0.5% strain.

Generally, pH exerts a great impact on the gel formation of pectin many small voids but less water. Eventually, the gelling process is re-
materials. It was previously reported that the gelation pH for high tarded and the mechanical properties of the resultant product might
methoxyl pectin is lower than 3.5, while that for low methoxyl pectin be undermined.
is in a wider range of pH 2.6–7 with or without sugar [29,42]. In our
work, the evolution of G′ and G″ during cooling from 85 °C to 25 °C 3.3.2. Frequency sweep
was studied for solutions with pH of 3, 4, 5 and 7. Clearly, among The rheological properties play an essential role in evaluating the ef-
these systems, the one at pH 4 presented the highest gel strength after ficacy of pectin in modifying food matrix. Typical mechanical spectra for
being cooled to room temperature (Fig. 3c). The gel strength at pH 5 PF solutions present a solid-like nature, that is, the elasticity reflected by
was less than that at pH 3.0. even at pH 7, the PF solution cannot form G′ dominates over the viscocity reflected by G″. The evolution of storage
a gel at any temperature with its G″ values consistently higher than modulus G′ and loss modulus G″ against frequency (Hz) of different PF
the corresponding G′ values. The effects of pH should be related to its in- gels are shown in Fig. 4.
fluence on the dissociation degree of the carboxyl groups in pectin mol- Over the tested frequency range, all PF systems at concentrations of
ecules. At a lower pH such as 3.0, the carboxyl groups in pectin 0.5 wt%, 1.5 wt% and 2.5 wt% demonstrated a predominance of elastic-
molecules are highly protonized (\\COOH) and there are not enough ity. This applauded the gelling capacity of the present PF. Moreover,
dissociated carboxyl groups (\\COO−) to interact with calcium ions to the PF concentration had a significant effect on its viscoelastic moduli
induce the gel formation. On the other hand, as the pH increases to 5 (Fig. 4a). The higher the PF concentration, the stronger the gel nature
or 7, the dissociation degree of the carboxyl groups in pectin molecules of the system was. The increase in the concentration of pectin certainly
will be substantially improved, which highly increased the electrostatic resulted in a higher amount of crosslinks among polymers, which are
repulsion among the negatively charged pectin molecules and thus im- crucial to build up the gel network of certain extent of stiffness. These
paired the gel network structure or even disabled the formation of poly- results indicated that more stable and stronger gels can be obtained at
mer based three-dimension structure [31]. higher PF concentrations, which was in line with the results obtained
The presence of calcium ions or other divalent metal ions could form in temperature sweep test.
intermolecular ionic “egg-box” junction zones, which is responsible for The impacts of sucrose on both moduli seemed to be much more
the gel formation of pectin solution [31]. Besides the electrostatic inter- complicated than that of PF concentration (Fig. 4b). As for sucrose con-
actions, van der Waals interactions and hydrogen bonds could stabilize centration, the system with 30 wt% sucrose presented the highest value
the chains when egg-boxes were formed between neighboring chains. in G′ while the system with of 10 wt% sucrose presented the lowest
Fig. 3d showed that the optimal CaCl2 concentration for the gelation of value in G″. For all systems, the G″ value increased with the frequency
the present PF is around 10 mM. The gel strength with 30 mM CaCl2 and the similar trend was observed for G′ of the system with 30 wt% su-
was lower than that with 10 mM. The previous study declared that crose. However, sharp drops in G′ were observed at approximately
the application of excessive CaCl2 in pectin system might cause sheet- 7–8 Hz for the systems with 10 wt% and 50 wt% sucrose. This indicated
like aggregation and precipitation instead of gel formation [44]. In the fragility of network structure formed in these systems. An appropri-
these aggregates or precipitates, a large number of junction zones ate level of sucrose could enhance the hydrophobic interactions among
with short distances formed between pectin molecules. They contained pectin molecules. However, the excessive sucrose could damage the
636 L. Feng et al. / International Journal of Biological Macromolecules 128 (2019) 629–637

Fig. 4. The frequency dependences of G′ (filled symbols) and loss modulus G″ (open symbols) of the dispersion of pectin-rich fraction (PF) from blueberry wine pomace as affected by PF
concentration (a, 20 mM CaCl2, 30% sucrose, pH 4.0), sucrose (b, 20 mM CaCl2, 1.5% PF, pH 4.0), pH (c, 1.5% PF, 20 mM CaCl2, 30% sucrose) and CaCl2 (d, 1.5% PF, 30% sucrose, pH 4.0). The
determinations were performed at 0.5% strain.

already formed gel structure by extracting water from it and made it absolute values were arranged in a CaCl2 order of 10 mM N 30 mM
more fragile. It seemed to be inconsistent with the previous discussion N 5 mM. Notably, the gelation of high methoxyl pectin could be acceler-
of viscoelastic moduli on cooling process regarding the impact of su- ated by CaCl2, which strengthened the molecular interactions of pectin
crose on gel-formation. On one hand, the increase in sucrose can pro- molecules and favored the approach of pectin molecules [46]. An opti-
mote the hydrophobic interactions favoring the approach of polymers. mum distance between neighboring junction zones is important for
On the other hand, the balance between hydrogen bonding and hydro- both the formation and mechanical stability of a pectin gel. If the dis-
phobic interaction is more crucial to a gel, which finally determine the tance between the junction zones is too long, as in the case of 5 mM
restraints of a considerable number of water molecules within the del- CaCl2 gels, a very fragile soft gel with low mechanical stability will be
icate network. Once the hydrophobic interaction is overwhelming, in- formed. Gels with lots of rod-like junction zones most likely formed in
tensive oscillation might lead to the fracture of architecture 20 mM and 40 mM CaCl2 but failed to hold water during the redistribu-
maintained by hydrogen bonds, which in turn accelerates the collapse tion of water taking place in gels aging. They are soft and vulnerable to
of spongy microstructure. frequency sweep.
The effects of matrix pH on the stability of pectin gels were illus-
trated in Fig. 4c. For the system at pH 7, the instability of the gels was ev- 4. Conclusion
ident even at low frequency and the corresponding crossover of G′ and
G″ took place at 0.141 Hz. A sharply drop of G′ was observed for this sys- The present work attempts to clarify the possibility of applying the
tem at approximately 2.82 Hz. The systems at pH 4 and pH 5 eventually pectin-rich fraction (PF) extracted from blueberry wine by-product as
displayed a solid-state nature and no crossover of two moduli was ob- a gelling agent in the food industry. Chemical characterization suggests
served within the tested frequency range. However, for the system at it is medium-methoxylated (DE 51.66%), mainly composed of HG and
pH 3, a transition from gel to sol was observed at 4.7 Hz. This is possibly RG-II with short side-chain of RG-I. In general, dispersions prepared at
due to the fact that the pectin chains were much less charded at pH 3, higher concentrations of PF showed higher shear viscosity, but failed
thus weakening electrostatic repulsion among pectin molecules and im- to form very viscous liquid without divalent cations such as CaCl2 and
proving the sensitivity of pectin to Ca2+. While at pH N 4.5, the electro- certain amount of soluble solids (often N50% sucrose). Both cold set ge-
static attractions might be intensified due to the fully charged pectin lation and ionotropic gelation could induce gelling of PF. Pectin gels of
chains, and as a result, galacturonic residues turned to be insensitive stable viscoelastic gel-like behaviors as a function of frequency could
to Ca2+ [24,38,39]. The small-angle x-ray scattering analysis suggested be formed with 50% sucrose at 10 mM CaCl2 and pH 4. These results
that both rod-like junction zones and point-like crosslinks between were attributed to the unique characteristics of PF, which is bound in
neighboring chains were formed in calcium-induced low methoxyl pec- medium-methoxylated pectin.
tin gels at pH 6. However, more rod-like junction zones were replaced
by the point-like crosslinks at higher CaCl2 concentration [45]. Acknowledgements
It is interesting to note that the 10 mM CaCl2 gels experienced a gel
to sol transition at 11.2 Hz (Fig. 4d), even though 10 mM CaCl2 was op- This work was supported by National Key Research and Develop-
timal for gel formation and gel strength enhancement. Besides, both G′ ment Plan (2016YFD0400204-2), Fundamental Research Funds for the
and G″ moduli increased with the increase in shear frequency, and their Central Universities (SWU117046), Open Funds provided by Guangxi
L. Feng et al. / International Journal of Biological Macromolecules 128 (2019) 629–637 637

Key Laboratory of Agricultural Resources Chemistry and Biotechnology [22] O. Yuliarti, K.K. Goh, L. Matia-Merino, J. Mawson, C. Brennan, Extraction and charac-
terisation of pomace pectin from gold kiwifruit (Actinidia chinensis), Food Chem. 187
(KF06) and Development and Research Center of Sichuan Cuision (2015) 290–296.
(CC18Z13). [23] E.R. Morris, D.A. Powell, M.J. Gidley, D.A. Rees, Conformations and interactions of
pectins: i. polymorphism between gel and solid states of calcium polygalacturonate,
J. Mol. Biol. 155 (4) (1982) 507–516.
Appendix A. Supplementary data [24] J. Singthong, S. Ningsanond, S.W. Cui, H.D. Goffc, Extraction and physicochemical
characterization of Krueo Ma Noy pectin, Food Hydrocoll. 19 (5) (2005) 793–801.
Supplementary data to this article can be found online at https://doi. [25] H.A. Schols, A.G.J. Voragen, Complex pectins: structure elucidation using enzymes,
Prog. Biotechnol. 14 (1995) (1996) 3–19.
org/10.1016/j.ijbiomac.2019.01.166. [26] J.A. Ferreira, I. Mafra, M.R. Soares, D.V. Evtuguin, M.A. Coimbra, Dimeric calcium
complexes of arabinan-rich pectic polysaccharides from Olea europaea L. cell
References walls, Carbohydr. Polym. 65 (4) (2006) 535–543.
[27] J.C. Lee, S.N. Timasheff, The stabilization of proteins by sucrose, J. Biol. Chem. 256
[1] V.P. Gouw, J. Jung, Y. Zhao, Functional properties, bioactive compounds, and invitro, (14) (1981) 7193.
gastrointestinal digestion study of dried fruit pomace powders as functional food in- [28] Y.I. Matveev, V.Y. Grinberg, V.B. Tolstoguzov, The plasticizing effect of water on pro-
gredients, LWT Food Sci. Technol. 80 (2017) 136–144. teins, polysaccharides and their mixtures. Glassy state of bio- polymers, food and
[2] K. Riihinen, L. Jaakola, S. Kärenlampi, A. Hohtola, Organ-specific distribution of phe- seeds, Food Hydrocoll. 14 (2000) 425–437.
nolic compounds in bilberry (Vaccinium myrtillus) and “northblue” blueberry [29] D. Giacomazza, D. Bulone, P.L. San Biagio, R. Lapasin, The complex mechanism of HM
(Vaccinium corymbosum × V. angustifolium), Food Chem. 110 (1) (2008) 156–160. pectin self-assembly: a rheological investigation, Carbohydr. Polym. 146 (2016)
[3] H. Kim, G.E. Bartley, A.M. Rimando, W. Yokoyama, Hepatic gene expression related 181–186.
to lower plasma cholesterol in hamsters fed high-fat diets supplemented with blue- [30] I. Braccini, M.A. Rodríguez-Carvajal, S. Pérez, Chain-chain interactions for methyl
berry peels and peel extract, J. Agric. Food Chem. 58 (7) (2010) 3984–3991. polygalacturonate: Models for high methyl-esterified pectin junction zones,
[4] A. Puganen, H. Kallio, K.M. Schaich, J.P. Suomela, B. Yang, Red/green currant and sea Biomacromolecules 6 (3) (2008) 1322–1328.
buckthorn berry press residues as potential sources of antioxidants for food use, J. [31] G.T. Grant, E.R. Morris, D.A. Rees, P.J.C. Smith, D. Thom, Biological interactions be-
Agric. Food Chem. 66 (13) (2018). tween polysaccharides and divalent cations: the egg-box model, FEBS Lett. 32 (1)
[5] A. Stobnicka, M. Gniewosz, Antimicrobial protection of minced pork meat with the (1973) 195–198.
use of Swamp Cranberry (Vaccinium oxycoccos, L.) fruit and pomace extracts, J. [32] D.A. Rees, Polysaccharide conformation in solutions and gels-recent results on pec-
Food Sci. Technol. 55 (1) (2018) 62–71. tin, Carbohydr. Polym. 2 (4) (1982) 254–263.
[6] P. Kathirvel, Y. Gong, M.P. Richards, Identification of the compound in a potent cran- [33] A.B. O'Brien, K. Philp, E.R. Morris, Gelation of high-methoxy pectin by enzymic
berry juice extract that inhibits lipid oxidation in comminuted muscle, Food Chem. deesterification in the presence of calcium ions: a preliminary evaluation,
115 (3) (2009) 924–932. Carbohydr. Res. 344 (14) (2009) 1818–1823.
[7] S. Struck, D. Straube, S. Zahn, H. Rohm, Interaction of wheat macromolecules and [34] S.H. Yoo, M.L. Fishman, H. Atjr, H.G. Lee, Viscometric behavior of high-methoxy and
berry pomace in model dough: rheology and microstructure, J. Food Eng. 223 low-methoxy pectin solutions, Food Hydrocoll. 20 (1) (2006) 62–67.
(2018) 109–115. [35] P.M. Gilsenan, R.K. Richardson, E.R. Morris, Associative and segregative interactions
[8] L.E. Laroze, B. Díazreinoso, A. Moure, M.E. Zúñiga, H. Domínguez, Extraction of anti- between gelatin and low-methoxy pectin: part 2—co-gelation in the presence of ca,
oxidants from several berries pressing wastes using conventional and supercritical Food Hydrocoll. 17 (6) (2003) 739–749.
solvents, Eur. Food Res. Technol. 231 (5) (2010) 669–677. [36] X. Hua, K. Wang, R. Yang, J. Kang, J. Zhang, Rheological properties of natural low-
[9] J. Paesa, R. Dotta, G.F. Barberob, J. Martíneza, Extraction of phenolic compounds and methoxyl pectin extracted from sunflower head, Food Hydrocoll. 44 (44) (2015)
anthocyanins from blueberry (Vaccinium myrtillus L.) residues using supercritical 122–128.
CO2 and pressurized liquids, J. Supercrit. Fluids 95 (2014) 8–16. [37] G.J. Li, K.C. Chang, Viscosity and gelling characteristics of sunflower pectin as af-
[10] A.L. Kjøniksen, M. Hiorth, N. Bo, Association under shear flow in aqueous solutions fected by chemical and physical factors, J. Agric. Food Chem. 45 (12) (1997)
of pectin, Eur. Polym. J. 41 (4) (2005) 761–770. 4785–4789.
[11] S.A. El-Nawawi, Y.A. Heikal, Factors affecting the production of low-ester pectin gels, [38] F. Capel, T. Nicolai, D. Durand, P. Boulenguer, V. Langendorff, Influence of chain
Carbohydr. Polym. 26 (3) (1995) 189–193. length and polymer concentration on the gelation of (amidated) low-methoxyl pec-
[12] D. Lootens, F. Capel, D. Durand, T. Nicolai, P. Boulenguer, V. Langendorff, Influence of tin induced by calcium, Biomacromolecules 6 (6) (2005) 2954–2960.
pH, Ca concentration, temperature and amidation on the gelation of low methoxyl [39] X. Li, S. Al-Assaf, Y. Fang, G.O. Phillips, Characterisation of commercial lm-pectin in
pectin, Food Hydrocoll. 17 (3) (2003) 237–244. aqueous solution, Carbohydr. Polym. 92 (2) (2013) 1133–1142.
[13] K. Bélafi-Bakó, P. Cserjési, S. Beszédes, Z. Csanádi, C. Hodúr, Berry pectins: [40] H. Kastner, K. Kern, R. Wilde, A. Berthold, U. Einhorn-Stoll, S. Drusch, Structure for-
microwave-assisted extraction and rheological properties, Food Bioprocess Technol. mation in sugar containing pectin gels - influence of tartaric acid content (pH) and
5 (3) (2012) 1100–1105. cooling rate on the gelation of high-methoxylated pectin, Food Chem. 144 (2014)
[14] B. Pasandide, F. Khodaiyan, Z.E. Mousavi, S.S. Hosseini, Optimization of aqueous pec- 44–49.
tin extraction from Citrus medica peel, Carbohydr. Polym. 178 (2017) 27–33. [41] P.G.D. Gennes, Reptation of a polymer chain in the presence of fixed obstacles, J.
[15] N. Blumenkrantz, G. Asboe-Hansen, New method for quantitative determination of Chem. Phys. 55 (2) (1971) 572–579.
uronic acids, Anal. Biochem. 54 (2) (1973) 484–489. [42] C. Löfgren, S. Guillotin, H. Evenbratt, H. Schols, A.M. Hermansson, Effects of calcium,
[16] L. Zhang, X. Ye, T. Ding, X. Sun, Y. Xu, D. Liu, Ultrasound effects on the degradation pH, and blockiness on kinetic rheological behavior and microstructure of HM pectin
kinetics, structure and rheological properties of apple pectin, Ultrason. Sonochem. gels, Biomacromolecules 6 (2) (2005) 646–652.
20 (1) (2013) 222–231. [43] Y. Zhou, R. Jiang, W.S. Perkins, Y. Cheng, Morphology evolution and gelation mech-
[17] M.S. Shin, S. Lee, K.Y. Lee, H.G. Lee, Structural and biological characterization of anism of alkali induced konjac glucomannan hydrogel, Food Chem. 269 (2018)
aminated-derivatized oat beta-glucan, J. Agric. Food Chem. 53 (14) (2005) 5554. 80–88.
[18] Z. Lin, J. Fischer, L. Wicker, Intermolecular binding of blueberry pectin-rich fractions [44] B.R. Thakur, R.K. Singh, A.K. Handa, Chemistry and uses of pectin—a review, Crit. Rev.
and anthocyanin, Food Chem. 194 (2016) 986–993. Food Sci. Nutr. 37 (1) (1997) 47–73.
[19] M. Correding, W. Kerr, L. Wicker, Molecular characterization of commercial pectins [45] I. Ventura, J. Jammal, H. Bianco-Peled, Insights into the nanostructure of low-
by separation with linear mix gel permeation columns inline with multi-angle light methoxyl pectin–calcium gels, Carbohydr. Polym. 97 (2) (2013) 650–658.
scattering detection, Food Hydrocoll. 14 (1) (2000) 41–47. [46] Y. Yang, G. Zhang, Y. Hong, Z. Gu, F. Fang, Calcium cation triggers and accelerates the
[20] J. Jung, L. Wicker, Laccase mediated conjugation of sugar beet pectin and the effect gelation of high methoxy pectin, Food Hydrocoll. 32 (2) (2013) 228–234.
on emulsion stability, Food Hydrocoll. 28 (1) (2012) 168–173.
[21] Y. Kim, Q. Teng, L. Wicker, Action pattern of Valencia orange PME deesterification of
high methoxyl pectin and characterization of modified pectins, Carbohydr. Res. 340
(17) (2005) 2620–2629.

You might also like