Sie sind auf Seite 1von 9

International Journal of Biological Macromolecules 79 (2015) 894–902

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules


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

Characterization of fish gelatin–gum arabic complex coacervates


as influenced by phase separation temperature
Mohammad Anvari a,b , Cheol-Ho Pan c,d , Won-Byong Yoon e , Donghwa Chung f,g,∗
a
Department of Marine Food Science and Technology, Gangneung-Wonju National University, Gangneung 210-702, Republic of Korea
b
School of Food Science, University of Idaho, Moscow, ID 83843, USA
c
Laboratory of Biomodulation, KIST Gangneung Institute of Natural Products, Gangneung 210-340, Republic of Korea
d
Department of Biological Chemistry, Korea University of Science and Technology (UST), Daejeon 305-350, Republic of Korea
e
Department of Food Science and Biotechnology, Kangwon National University, Chuncheon 200-701, Republic of Korea
f
Graduate School of International Agricultural Technology, Seoul National University, Pyeongchang 232-916, Republic of Korea
g
Institute of Food Industrialization, Institutes of Green Bio Science and Technology, Seoul National University, Pyeongchang 232-916, 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 rheological and structural characteristics of fish gelatin (FG)–gum arabic (GA) complex coacervate
Received 29 March 2015 phase, separated from an aqueous mixture of 1% FG and 1% GA at pH 3.5, were investigated as influenced
Received in revised form 30 May 2015 by phase separation temperature. Decreasing the phase separation temperature from 40 to 10 ◦ C lead to:
Accepted 2 June 2015
(1) the formation of a coacervate phase with a larger volume fraction and higher biopolymer concentra-
Available online 6 June 2015
tions, which is more viscous, more structural resistant at low shear rates, more shear-thinning at high
shear rates, and more condensed in microstructure, (2) a solid-like elastic behavior of the phase separated
Keywords:
at 10 ◦ C at a high oscillatory frequency, (3) the increase in gelling and melting temperatures of the coacer-
Fish gelatin–gum arabic complex
coacervation
vate phase (3.7−3.9 ◦ C and 6.2−6.9 ◦ C, respectively), (4) the formation of a more rigid and thermo-stable
Phase separation temperature coacervate gel. The coacervate phase is regarded as a homogeneously networked biopolymer matrix dis-
Oscillatory shear persed with water vacuoles and its gel as a weak physical gel reinforced by FG–GA attractive electrostatic
interactions.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction two-phase system where one phase is enriched in insoluble


complexes and the other depleted [3,6]. The sequential process,
Protein–polysaccharide interactions in an aqueous environment composed of (1) the formation of charge neutralized insoluble com-
have been extensively studied in recent years because of their plexes and (2) the subsequent spontaneous macroscopic phase
high potential applications in the development of bioactive deliv- separation to complex-rich and solvent-rich phases, is known as
ery devices, fat replacers, meat analogs, gels, emulsions, edible associative separation, which can be also called complex coacer-
films, and coatings in food and pharmaceutical industries [1–3]. vation or precipitation for the separation of liquid-state insoluble
The interactions are either attractive or repulsive, depending on complexes (complex coacervates) or solid-state insoluble com-
biopolymer characteristics (e.g., molecular weight, charge den- plexes (precipitates), respectively [3,6].
sity, conformation, and flexibility), solvent properties (e.g., pH and The formation of insoluble protein–polysaccharide complexes
ionic strength), and mixing conditions (e.g., protein to polysaccha- can be significantly influenced by temperature through the
ride ratio, total biopolymer concentration, temperature, stirring, temperature-dependent alterations not only in the conformation
and pressure) [4,5]. Attractive interactions, primarily attributed of biopolymers but also in the noncoulombic interactions between
to electrostatic attractions between oppositely charged biopoly- the biopolymers, such as hydrogen bonding and hydrophobic inter-
mers, induce the formation of biopolymer complexes, which actions. Harding et al. [7] demonstrated that the conformational
could be either soluble in a single phase or insoluble to build a change of bovine serum albumin by heat-induced denaturation
enhanced the formation of insoluble complexes with alginate. Fang
et al. [8] showed that the conformational change of pigskin gelatin
∗ Corresponding author at: Graduate School of International Agricultural Tech-
by heat-induced transition from helix to random coil could also
nology, Seoul National University, Pyeongchang 232-916, Republic of Korea.
enhance the complexation with ␬-carrageenan. Hydrogen bonding,
Tel.: +82 33 339 5793; fax: +82 33 339 5716. in principle, is favored at low temperatures, whereas hydropho-
E-mail address: dchung@snu.ac.kr (D. Chung). bic interactions become stronger with increasing temperature,

http://dx.doi.org/10.1016/j.ijbiomac.2015.06.004
0141-8130/© 2015 Elsevier B.V. All rights reserved.
M. Anvari et al. / International Journal of Biological Macromolecules 79 (2015) 894–902 895

because of the exposure of more interaction sites by heat-induced respectively [3]. Sodium azide (NaN3 ) was obtained from Daehung
conformational changes of the biopolymers [1,9,10]. Liu et al. [11] Chemicals & Metals (Siheung, Korea). Fluorescein isothiocyanate
also reported that a decrease of temperature from 23 to 6 ◦ C sig- (FITC), rhodamine B isothiocyanate (RITC), and dimethyl sulfoxide
nificantly improved the complexation between pea protein isolate (DMSO) were purchased from Sigma Chemical Co. (St. Louis, MO,
and gum arabic by strengthening hydrogen bonding as a sec- USA). All other reagents were of analytical grade purity.
ondary force, whereas hydrophobic interactions appeared to be
more prevalent in a higher temperature range of 23−60 ◦ C and 2.2. Preparation of FG–GA coacervate phase
to be associated with the stability, rather than the formation of
complexes. Accurately weighed amounts of FG and GA were separately dis-
The spontaneous macroscopic phase separation of insoluble solved in 100 mL of distilled water at 40 ◦ C and mixed together at a
complexes may be also strongly dependent on temperature. The weight ratio of FG to GA (FG:GA) of 1:1 and a total biopolymer con-
phase separation phenomenon, mainly driven by entropy, takes centration of 2% (w/v), followed by adjusting the pH to 3.5. Sodium
place by a nucleation-and-growth-like mechanism, as a result of the azide was added as a preservative at a concentration of 0.02% (w/v).
rearrangement, coalescence, Ostwald ripening, or aggregation of The FG–GA mixture was incubated at 40 ◦ C in a shaking water bath
insoluble complexes, mostly via the reformation of intermolecular at 100 rpm for 24 h for sufficient electrostatic interactions between
interactions and the expulsion of water molecules [1,12,13]. It was FG and GA molecules, and then placed statically at 10, 30, or 40 ◦ C
suggested by Nigen et al. [14] that the initial complexation between for another 24 h for the macroscopic phase separation of FG–GA
two oppositely charged globular proteins, ␣-lactalbumin and insoluble complexes to reach equilibrium state.
lysozyme, might be caused by electrostatic attractions, however,
the growth of complexes could be largely driven by hydrogen bond- 2.3. Measurements of phase volume and composition
ing at low temperatures, resulting in the formation of reversible
aggregates, but by hydrophobic interactions at higher temperatures The volume fraction of coacervate phase was determined as
to form coacervate-like structures. Kayitmazer et al. [15] showed follows:
that the shear-dependent viscosity and dynamic moduli of bovine
Vc
serum albumin-chitosan coacervate phase was significantly depen- Volume fraction (%) = × 100 (1)
Vt
dant on temperature, although the temperature was not the phase
separation temperature but the temperature of measurements. It is where Vc is the volume of coacervate phase and Vt is the total vol-
therefore expected that the rheology and structure of complex-rich ume of FG–GA mixture. The amounts of total biopolymers and FG
phase obtained by protein–polysaccharide complex coacervation in solvent-rich phase were measured by oven drying at 105 ◦ C for
and the sol–gel transformation of such coacervate phase, which are 24 h and by the Lowry method, respectively [3,16]. The values mea-
directly related to the success of some valuable applications such as sured above were used to calculate the amount and concentration
microencapsulation and emulsion stabilization, could be consider- of biopolymers in the coacervate phase and the FG:GA value of
ably influenced by the temperature of phase separation. However, coacervate phase. All measurements were performed at least in
such aspects of complex coacervation have not been systematically triplicate.
studied.
Fish gelatin (FG) is regarded as a promising alternative to mam- 2.4. Fourier transformed infrared (FTIR) spectroscopy
malian gelatins, because it can be produced from fish skin and
bones, comprising nearly 30% of fish processing byproducts, and FTIR spectroscopy was used to examine the changes in the
does not have safety- or religion-related consumer concerns [3]. secondary structure of FG and molecular interactions between
It was found by our group that the FG from cold water fish skin biopolymers, such as electrostatic interactions, hydrogen bonding,
(58 kDa) underwent complex coacervation with gum arabic (GA, or hydrophobic interactions, in the coacervate phase. The coacer-
382 kDa), one of the most widely used anionic polysaccharides, vate phases obtained were dried in a vacuum oven at their phase
in aqueous solutions at 40 ◦ C [3]. The formation of insoluble or separation temperatures. Aqueous solutions of FG and GA, sepa-
soluble FG–GA complexes was strongly influenced by pH and FG rately prepared at 1% (w/v) and pH 3.5, were also dried as controls.
to GA weight ratio, as often reported for other protein–anionic The dried samples of 2 mg in approximately 100 mg potassium
polysaccharide pairs, and the total concentration and composition bromide were placed in discs, and FTIR spectra were obtained in
of biopolymers in the coacervate phase was dependent on the total the wavenumber range of 4000–500 cm−1 using an infrared spec-
concentration of biopolymers in the initial mixture. In the present trophotometer (Tensor 27, Bruker Instruments, Billerica, MA, USA)
study, we aimed to investigate the rheological and structural at a scanning rate of 10 kHz with 64 scans and a resolution of 4 cm−1 .
characteristics of FG–GA complex coacervate phase as influenced The background was obtained against pure potassium bromide pel-
by phase separation temperature, using dynamic oscillatory and let. All data treatments were carried out using OPUS 7.0 software
steady shear rheological measurements, Fourier transform infrared (Bruker Instruments, Billerica, MA, USA).
spectroscopy (FTIR), confocal scanning laser microscopy (CSLM),
and scanning electron microscopy (SEM). 2.5. Confocal scanning laser microscopy (CSLM)

CSLM was used to examine the microstructure of FG–GA coac-


2. Materials and methods ervate phase. FG and GA were covalently stained with fluorescent
markers, FITC and RITC, respectively. The excitation/emission
2.1. Materials wavelengths of FITC and RITC are 485/530 nm and 530/590 nm,
respectively. The staining was performed as described in Schmitt
Fish gelatin (FG, from cold water fish skin) and gum arabic (GA, et al. [9]. Briefly, aqueous solutions of FG and GA were separately
from Acacia tree) were purchased from Sigma Chemical Co. (St. prepared at 2% (w/v), and their pH values were adjusted to 8.5
Louis, MO, USA) and Carl Roth GmbH (Karlsruhe, Germany), respec- and 10.5, respectively, to facilitate the staining reaction. There-
tively. The average molecular weights of FG and GA, determined after, 25 ␮L of 2% (w/w) FITC or RITC solution in DMSO was added
by viscometry using an Ostwald viscometer and Mark–Houwink to 100 mL of the FG or GA solution, respectively, and the staining
equation in our preliminary experiments, were 58 and 382 kDa, reaction was allowed to proceed for 1.5 h with gentle stirring at
896 M. Anvari et al. / International Journal of Biological Macromolecules 79 (2015) 894–902

Table 1 Secondly, the dynamic oscillatory behavior of the three FG–GA


Volume fraction and composition of FG–GA coacervate phasesa separated at three
coacervate phases was examined by frequency sweep tests, where
different temperatures.
the elastic or storage modulus (G ) and viscous or loss modulus (G )
Parameter Phase separation temperature of each phase were determined as a function of angular frequency
40 ◦ C 30 ◦ C 10 ◦ C (ω) in the range of 0.1–100 rad/s. The linear viscoelastic region,
where G and G are independent of strain (), was determined by
Volume fraction (%, v/v) 7.0 ± 0.1b
7.1 ± 0.3b
7.8 ± 0.1a
FG content (mg) 233.3 ± 1.8b 235.2 ± 1.3b 239.0 ± 0.8a strain sweep tests in the  range of 0.01–100% and at ω = 2 rad/s.
GA content (mg) 170.5 ± 1.3c 174.9 ± 0.8b 231.0 ± 2.0a For the coacervate phases separated at 10, 30, and 40 ◦ C, the lin-
Total biopolymer content (mg) 403.8 ± 1.9c 410.0 ± 2.5b 470.0 ± 1.7a ear viscoelastic region was found to be up to 2.0%, 1.3%, and 1.2%,
FG concentration (%, w/v) 8.3 ± 0.0a 8.3 ± 0.0a 7.7 ± 0.0b respectively (data not shown), and therefore, a  value of 1% was
GA concentration (%, w/v) 6.1 ± 0.0c 6.2 ± 0.0b 7.4 ± 0.0a
Total biopolymer 14.4 ± 0.1b 14.4 ± 0.1b 15.1 ± 0.1a
used for the frequency sweep and temperature sweep tests.
concentration (%, w/v) Thirdly, the sol–gel transformation of the three FG–GA coacer-
FG to GA weight ratio (FG:GA) 1.4 ± 0.0a 1.3 ± 0.0a 1.0 ± 0.1b vate phases was investigated by temperature sweep tests, where G
a
All the phases were separated from the FG–GA mixture prepared at a total and G were monitored at ω = 2 rad/s and  = 1% with decreasing
biopolymer concentration of 2% (w/w), FG:GA = 1:1, pH 3.5, and 40 ◦ C. Values are temperature from each phase separation temperature to 3 ◦ C and
expressed as mean ± standard deviation of at least three different samples. Val- then back to the phase separation temperature at a cooling or heat-
ues with different superscript letters in the same row are significantly different at ing rate of 0.016 ◦ C/min, and the gelling and melting points (Tg and
p < 0.05.
Tm , respectively) of each coacervate phase were determined. Com-
plex modulus (G*) was obtained at 3 ◦ C and used to compare the
overall rigidity of the gels transformed from the coacervate phases.

40 ◦ C. The two solutions were then mixed together as described 


in Section 2.2 to obtain FG–GA coacervate phases separated at G∗ = (G )2 + (G )2 (2)
three different temperatures (10, 30, and 40 ◦ C). The microstruc-
ture of stained FG–GA coacervate phases were imaged using FV-300 Finally, the dynamic oscillatory behavior of the gels obtained by
confocal scanning unit (Olympus, Tokyo, Japan) equipped with an cooling the coacervate phases to 3 ◦ C was also examined by fre-
Olympus IX71 inverted microscope and an argon ion laser. The quency sweep tests at  = 1% in the range of 0.1–100 rad/s. All the
laser was adjusted in the green/red fluorescence mode, yielding rheological measurements were performed at least in triplicate.
two excitation wavelengths of 485 and 530 nm, to obtain green and
red fluorescence images from two separate channels. All images
were taken at a magnification of 40× (oil immersion, numeric aper- 3. Results and discussion
ture 1.30) and processed using Fluoview Software (Olympus, Tokyo,
Japan). 3.1. Volume fraction and composition of FG–GA coacervate phase

The volume fraction of FG–GA coacervate phase was found to


2.6. Scanning electron microscopy (SEM)
increase from 7.0% to 7.8% with decreasing phase separation tem-
perature from 40 to 10 ◦ C (Table 1). The values were close to the
SEM was also used to examine the microstructure of the FG–GA
value (about 8.0%) reported for the coacervate phase separated
coacervate phases separated at three different temperatures (10,
from the mixture of gelatin (unknown source) and GA, which was
30, and 40 ◦ C). Each sample was fixed in 2% (w/w) glutaraldehyde
prepared under similar conditions as in the present study: a weight
solution, dehydrated using different aqueous ethanol solutions, and
ratio of gelatin to GA of 1:1, a total biopolymer concentration of 2%,
placed in acetone as described in Espinosa-Andrews et al. [17]. Each
pH about 3.5, and a mixing and phase separation temperature of
sample was dried in a vacuum oven for 5 min, fragmented, and
40 ◦ C [18]. The composition analysis shows that the amount and
mounted on stubs with the fractured face upwards, and gold-coated
concentration of total biopolymer and GA in the coacervate phase,
with a fine coat ion sputter (JFC 1100, JEOL Ltd., Akishima, Japan).
as well as the amount of FG in the phase significantly increased
A low vacuum scanning electron microscope (Inspect F50, FEI Co.,
with decreasing phase separation temperature, while the concen-
Hillsboro, OR, USA), was used at 15 kV to obtain the SEM images
tration of FG was reduced from 8.3% to 7.7% (w/v), resulting in
of the cross-section of each sample at magnifications of 100× and
the considerable decrease in the weight ratio of FG to GA (FG:GA)
1000×, respectively.
from 1.4 to 1.0 (Table 1). The results indicate that decreasing phase
separation temperature from 40 to 10 ◦ C not only increase the vol-
2.7. Rheological analysis ume fraction of coacervate phase, but also enhance the partition
of the two biopolymers, especially GA molecules, into the coac-
All rheological measurements were carried out using a TA- ervate phase. The volume increase of coacervate phase may be
AR 2000 rheometer (TA Instruments Inc., New Castle, DE, USA), partly attributed to the partition of more GA molecules (382 kDa),
equipped with a cone–plate geometry (diameter 40 mm, angle 2◦ , which are much larger than FG molecules (58 kDa), into the phase at
and gap 53 ␮m) and a moisture trap. Firstly, the steady shear behav- lower phase separation temperature [13]. The phase separation of
ior of the FG–GA coacervate phases separated at three different protein–polysaccharide mixture is known to be an entropy-driven
temperatures (10, 30, and 40 ◦ C) was examined by measuring the process occurring through a nucleation-and-growth like mecha-
apparent viscosity (a ) of each phase as a function of shear rate nism, where the biopolymer complexes formed during the mixing
().
˙ Each sample was sheared from 0.01 to 100 1/s at its phase sep- undergo rearrangement, coalescence, Ostwald ripening, or aggre-
aration temperature, during which 61 data points were recorded gation, mostly via the reformation of intermolecular interactions
at 6 s intervals. As controls, three aqueous mixtures of FG and and the expulsion of water molecules [1,12,13]. Different phase
GA were prepared with the same biopolymer compositions as the volumes and compositions of FG–GA coacervate phases separated
coacervate phases, determined in Section 2.3 (Table 1), at pH 8.0, at different temperatures suggest that temperature could be an
where there are no significant attractive interactions between FG important factor influencing the phase separation mechanism of
and GA. FG–GA complexes.
M. Anvari et al. / International Journal of Biological Macromolecules 79 (2015) 894–902 897

3.2. Fourier transformed infrared (FTIR) spectroscopy 10


o
FG-GA at 40 C
o
Fig. 1 shows the FT-IR spectra of FG, GA, and the FG–GA FG-GA at 30 C
o
coacervate phases separated at three different temperatures. The FG-GA at 10 C
changes in the secondary structure of FG and the molecular interac- Control at 40 oC
1 Control at 30 oC
tions between biopolymers, such as electrostatic interactions and

ηa (Pa.s)
Control at 10 oC
hydrogen bonding, in the coacervate phases were investigated by Carreau fitting curve
examining three band regions: amide A (3600–2300 cm–1 ) for N H
stretching, amide I (1600–1700 cm–1 ) for primarily C O stretching,
and amide II (1335–1560 cm–1 ) for N H bending and C N stretch- 0.1
ing [18–21]. At 40 ◦ C, FG molecules showed a typical amide I peak
at 1635 cm−1 , which is regarded as a characteristic for the random
coil structure of gelatin [19,22,23], and GA molecules showed two
peaks at 1612 and 1423 cm−1 due to the asymmetric and symmet- 0.01
ric stretching of carboxyl acid salt COO− , respectively (Fig. 1) [17]. 0.01 0.1 1 10 100
The FG–GA coacervate phase obtained at 40 ◦ C did not show the two .
GA peaks, implying the involvement of COO− of GA molecules in γ (1/s)
electrostatic interactions, and its amide I peak at 1632 cm−1 indi- Fig. 2. Apparent viscosity (a ) versus shear rate () ˙ for FG–GA coacervate phases
cates that the FG molecules complexed with GA molecules also had separated at 10, 30, or 40 ◦ C and their corresponding controls. The coacervate phases
a random coil structure at 40 ◦ C. were separated from the FG–GA mixture prepared at a total biopolymer concentra-
At 30 ◦ C, FG showed a similar spectrum to that at 40 ◦ C, while tion of 2% (w/w), FG:GA = 1:1, pH 3.5, and 40 ◦ C. The controls are the FG–GA mixtures
prepared with the same biopolymer compositions as the coacervate phases, but at
the FG–GA coacervate phase exhibited an amide A peak at a
pH 8.0.
lower wavenumber (3420 cm−1 ) and an amide II peak at a higher
wavenumber (1539 cm−1 ) (Fig. 1). Amide A and amide II bands are
attributed to the stretching vibration of N H group bonded to O H downshift of amide A peak and the upshift of amide II peak suggest
group and the C N stretching combined with N H deformation the increase of hydrogen bonding in the coacervate phase [20,24].
of peptide group, respective, and therefore, both the wavenumber The FG molecules in the coacervate phase may also have a random
coil structure at 30 ◦ C, considering the amide I peak at 1639 cm−1 .
When the temperature was further reduced to 10 ◦ C, FG showed
amide I and amide II peaks at higher wavenumbers (1656 and
amide A

amide II

1541 cm−1 , respectively) (Fig. 1). The amide I peak at 1656 cm−1
amide I

is regarded as a characteristic for the helical structure of collagen


[23], and it is known that cold water fish gelatin undergoes a struc-
tural transition from a random coil to a helical conformation at
FG 40 oC
around 15 ◦ C [25]. The spectral changes, therefore, may indicate
3429 1635 1521 that FG molecules had an ordered and entangled helical secondary
structure at 10 ◦ C. The FG–GA coacervate phase at 10 ◦ C exhibited
FG 30 oC an amide A peak at a lower wavenumber (3410 cm−1 ) compared
to those prepared at higher temperatures (Fig. 1), suggesting that
1521
3427 1633 FG–GA coacervate phase had increased hydrogen bonding [24]
o
FG 10 C at a lower phase separation temperature. The stronger hydrogen
bonding in FG–GA coacervate phase at a lower phase separation
temperature, especially at 10 ◦ C, could be a reason for the enhanced
Transmittance (%)

partition of the biopolymers, especially GA molecules, into the


1541
1656 coacervate phase with a lower FG:GA value at a lower tempera-
o
3423 GA 40 C ture (Table 1). The FG–GA coacervate phase at 10 ◦ C also showed an
amide I peak at 1654 cm−1 (Fig. 1), indicating that the FG molecules
1612 1423 in the coacervate phase at 10 ◦ C also had a helical secondary struc-
3440
ture.
FG-GA 40 oC
3.3. Steady shear behavior of FG–GA coacervate phase
1523
1632
3427
Fig. 2 shows the apparent viscosity (a ) of the FG–GA coacer-
FG-GA 30 oC
1539 vate phases separated at three different temperatures as a function
1639
3420 of shear rate ().
˙ Three corresponding control samples, which are
the FG–GA mixtures prepared with the same biopolymer compo-
FG-GA 10 oC sitions as the coacervate phases, but at pH 8.0, where there is no
1541 significant attractive electrostatic interactions between FG and GA
1654
3410 molecules, were also investigated. At all the tested temperatures,
the coacervate phases showed much higher values of a than the
4000 3500 3000 2500 2000 1500 1000 500 controls, although the biopolymer composition was the same. This
may be because the resistance to flow was elevated due to the
Wavenumber (cm-1) FG–GA electrostatic interactions in the coacervate phases. Wein-
breck et al. [26] also reported that the coacervate phase of whey
Fig. 1. FTIR spectra of FG, GA, and FG–GA coacervate phases separated at 10, 30, or
40 ◦ C. The coacervate phases were separated from the FG–GA mixture prepared at protein and gum arabic at pH 4.0 had much higher viscosity than
a total biopolymer concentration of 2% (w/w), FG:GA = 1:1, pH 3.5, and 40 ◦ C. the control prepared at pH 7.0 with similar composition.
898 M. Anvari et al. / International Journal of Biological Macromolecules 79 (2015) 894–902

Table 2 100
Carreau model parameters estimated for FG–GA coacervate phasesa separated at G' (40 oC)
three different temperatures and their controls. G" (40 oC)
o
Sample 0 (Pa s) ∞ (Pa s) N −1
˙ c (s ) R2 10 G' (30 C)
o
G" (30 C)
FG–GA at 40 ◦ C 0.250 0.046 0.36 0.0110 0.9973 G' (10 oC)

G' or G'' (Pa)


FG–GA at 30 ◦ C 0.327 0.061 0.38 0.0380 0.9989
1 G" (10 oC)
FG–GA at 10 ◦ C 1.246 0.150 0.43 0.0827 0.9999
Control at 40 ◦ C 0.042 0.016 0.22 0.0025 0.9992
Control at 30 ◦ C 0.079 0.017 0.28 0.0028 0.9997
Control at 10 ◦ C 0.198 0.019 0.30 0.0031 0.9998 0.1

a
All the phases were separated from the FG–GA mixture prepared at a total
biopolymer concentration of 2% (w/w), FG:GA = 1:1, pH 3.5, and 40 ◦ C. The controls 0.01
are the FG–GA mixtures prepared at pH 8.0 with the same biopolymer compositions
as the coacervate phases, determined in Table 1.

0.001
All the samples exhibited a shear-thinning flow behavior, 0.1 1 10 100
and therefore, the well-known Carreau model was employed to ω (rad/s)
describe the viscosity profiles in Fig. 2 [27].
Fig. 3. Storage (G ) and loss (G ) moduli versus angular frequency (ω) for FG–GA
0 − ∞
coacervate phases separated at 10, 30, or 40 ◦ C. The coacervate phases were sepa-
a = ∞ +   2 N (3)
rated from the FG–GA mixture prepared at a total biopolymer concentration of 2%

1+ ˙ c
(w/w), FG:GA = 1:1, pH 3.5, and 40 ◦ C.

where 0 is the zero-shear viscosity (Pa s), i.e., the limit viscosity at separation temperature (Table 2), implying the increase of New-
low shear rates, ∞ is the infinite-shear viscosity (Pa s), i.e., the limit tonian plateau region at low shear rates. This is quite interesting
viscosity at high shear rates, ˙ c is the critical shear rate (s−1 ) mark- because the result indicates that at a lower temperature, the coac-
ing the onset of shear-thinning region, and N is a dimensionless ervate phase was more resistant to the structural deformation at
exponent related to the slope of shear-thinning region. The Car- low shear rates, although it showed stronger shear-thinning behav-
reau model showed a good fitting to the experimental data with ior once the structural deformation and rearrangements occurred
high R2 values (≥0.9973), and the model parameters estimated are above ˙ c . This is probably ascribed to the stronger hydrogen bond-
listed in Table 2. ing and the reduced molecular mobility at a lower temperature. The
The values of 0 and ∞ for FG–GA coacervate phase controls also exhibited an increase in ˙ c from 0.0025 to 0.0031 s−1
increased by about 5-fold (0.250–1.246 Pa s) and by about 3-fold with the temperature decrease, but not as dramatic as the coacer-
(0.046–0.150 Pa s), respectively, with decreasing phase separa- vate phase. At any tested temperature, much higher ˙ c value, i.e.,
tion temperature from 40 to 10 ◦ C (Table 2). This is probably wider low-shear Newtonian plateau, was observed for the coac-
because decreasing the temperature not only reduced the mobil- ervate phase than for the control. This is probably because at low
ity of the biopolymers but also increased the hydrogen bonding shear rates, the FG–GA attractive electrostatic interactions might
between the biopolymers, as observed in the FT-IR analysis. The disturb the deformation of GA molecules or other molecular struc-
higher mass fraction of GA, having about 6 times larger molecular ture, as suggested by Weinbreck et al. [26].
weight (382 kDa) compared to FG (58 kDa), at a lower temperature
(Table 1) may be also attributed to the increase of 0 and ∞ with 3.4. Oscillatory shear behavior of FG–GA coacervate phase
the temperature decrease. The controls also showed about 5-fold
increase in 0 (0.042–0.198 Pa s) but only a slight increase in ∞ Fig. 3 shows the frequency sweep of the FG–GA coacervate
(0.016–0.019 Pa s) with decreasing the temperature. phases separated at three different temperatures. For the phases
The value of N for FG–GA coacervate phase increased from 0.36 separated at 30 and 40 ◦ C, the storage modulus (G ) was found to
to 0.43 with decreasing phase separation temperature (Table 2), increase with the angular frequency (ω), but to decrease when
indicating that the coacervate phase was not only more viscous but the frequency exceeded 50 and 25 rad/s, respectively, implying
also more shear-thinning at a lower temperature. The coacervate the occurrence of structural breakdown in the phases due to the
phase had a higher GA fraction at a lower temperature (Table 1), reduced intermolecular connectivity at high frequency. However,
and it is known that polysaccharide relaxation is mainly responsible the G of the coacervate phase at 10 ◦ C continuously increased
for the shear thinning behavior of most concentrated biopolymer with the frequency without decreasing, and the coacervate phase
mixtures [26]. In addition, the hydrogen bonding became stronger obtained at a lower temperature showed a higher G at any fre-
with the temperature decrease, even leading to the formation of quency. The results confirm the increase of structuration in the
helical structure in FG at 10 ◦ C, as observed in the FT-IR analysis. coacervate phase with decreasing the temperature, probably due
Therefore, the biopolymers in the coacervate phase might undergo to the reduced molecular mobility, the stronger hydrogen bonding,
more conformational changes and rearrangements at a lower tem- and the higher GA fraction at a lower temperature, as previously
perature. The controls also showed an increase in N value from 0.22 mentioned.
to 0.30 with the temperature decrease. At any tested temperature, The loss modulus (G ) continuously increased with the fre-
the coacervate phase was found to have a higher N value, indicat- quency at all the tested temperatures, and showed larger values
ing greater shear-thinning behavior than the control, although the than the G values, indicating the liquid-like viscous behavior of
biopolymer composition was the same. This is probably because the the coacervate phases (Fig. 3). Decreasing the temperature also
complexes formed by attractive electrostatic interactions between raised the G value, but the increase of G was not as significant
FG and GA molecules, i.e., FG–GA complex coacervates, were sub- as that of G due to the structuration, resulting in the approach of
jected to more structural rearrangements than the biopolymers G to G with decreasing the temperature. The coacervate phase
mixed without significant attractive interactions. at 10 ◦ C even showed a crossover of G and G at a frequency of
The value of ˙ c for FG–GA coacervate phase significantly 73 rad/s, that is, exhibited liquid-like viscous behavior below the
increased by 7.5 times (0.0110–0.0827 s−1 ) with decreasing phase crossover frequency and solid-like elastic behavior above it. The
M. Anvari et al. / International Journal of Biological Macromolecules 79 (2015) 894–902 899

Fig. 4. CSLM micrographs of FG–GA coacervate phases separated at 40 ◦ C (A), 30 ◦ C (B), and 10 ◦ C (C). The images were obtained from FITC-stained FG molecules. The
coacervate phases were separated from the FG–GA mixture prepared at a total biopolymer concentration of 2% (w/w), FG:GA = 1:1, pH 3.5, and 40 ◦ C. Bar represents 50 ␮m.
(For interpretation of the references to color in text near the reference citation, the reader is referred to the web version of this article.)

FG–GA complexes in the phase at 10 ◦ C might have not enough which were found to be smallest in the coacervate phase at 10 ◦ C,
time to relax to a more favorable structure and dissipate energy indicating that the coacervate phase at 10 ◦ C had more condensed
above the crossover frequency, and consequently, act more like a microstructure than the phase at 30 or 40 ◦ C. A similar CSLM
rigid gel network, in which the energy imposed by the oscillatory microstructure was also reported by Schmitt et al. [29] for the
strain could be temporarily stored, leading to the solid-like elastic coacervate phase of ␤-lactoglobulin and gum arabic, which became
behavior of the coacervate phase [28]. denser with the loss of water vacuoles upon aging. Fig. 5 shows the
The coacervate phase of bovine serum albumin and chitosan also SEM micrographs of the three FG–GA coacervate phases. All the
showed a crossover frequency, below which G < G (viscous) but coacervate phases displayed a sponge-like porous microstructure
above which G > G (elastic), and both dynamic moduli increased due to the entrapment of water molecules within the coacervate
with the decrease of temperature, although the temperature was structure. The water vacuoles in the coacervate microstructure, i.e.,
not the phase separation temperature but the temperature of the pores in the micrographs, were observed to become smaller
measurements [14]. The coacervate phase of ␤-lactoglobulin and and more homogeneous in size with decreasing the temperature,
gum arabic also showed a transition from liquid-like to gel-like indicating more compact coacervate microstructure at a lower tem-
behavior with a crossover of G and G after an aging treatment perature.
for 48 h [29]. The coacervate phase of whey protein and gum According to our observation, therefore, the FG–GA coacervate
arabic also exhibited viscous behavior (G < G ) during the fre- phase is regarded as a dense and coalesced liquid phase, where FG
quency sweep, however, no crossover frequency was reported and and GA molecules are homogeneously complexed and networked
the effect of temperature was not investigated [26]. Neverthe- by electrostatic attractions and noncoulombic interactions, such as
less, the coacervate phases of bovine serum albumin-pectin and hydrogen bonding, to form a continuous matrix in which water
␤-lactoglobulin-pectin showed gel-like elastic behavior (G > G ) vacuoles are dispersed. In the present study, a close relationship
during the frequency sweep [30,31]. The different oscillatory was found between the microstructure and the rheological prop-
shear behaviors of the different coacervate phases reflect charac- erties of the coacervate phase. The denser the microstructure is,
teristically distinct structures of protein–polysaccharide complex the more viscous, the more shear-thinning, and the more elastic
coacervates. the coacervate phase is.

3.5. Microstructure of FG–GA coacervate phase


3.6. Sol–gel transformation of FG–GA coacervate phase
Fig. 4 shows the CSLM micrographs of the FG–GA coacervate
phases separated at three different temperatures, where the FG Fig. 6 shows the temperature dependence of G and G of the
molecules stained with FITC were represented in green. For all the FG–GA coacervate phases separated at three different tempera-
coacervate phases, the green-colored images were similar to the tures. The temperature was lowered from each phase separation
red-colored images from RITC-stained GA molecules (not shown), temperature to 3 ◦ C and then raised back to the phase separa-
demonstrating nearly homogeneous distribution of FG and GA tion temperature at a rate of 0.016 ◦ C/min. During the cooling, all
molecules in the coacervate phases. The black regions in the images the coacervate phases showed the following notable sequential
represent water vacuoles inside the coacervate microstructure, changes in the moduli; (1) a sudden increase in G immediately

Fig. 5. SEM micrographs of FG–GA coacervate phases separated at 40 ◦ C (A), 30 ◦ C (B), and 10 ◦ C (C). The coacervate phases were separated from the FG–GA mixture prepared
at a total biopolymer concentration of 2% (w/w), FG:GA = 1:1, pH 3.5, and 40 ◦ C. Bar represents 100 ␮m.
900 M. Anvari et al. / International Journal of Biological Macromolecules 79 (2015) 894–902

100 of G is probably due to the adjustment of the coacervate system


G* = 75.2 ± 1.8 Pa (A)
to the cooling conditions. As the cooling proceeds, the mobility of
Tg = 3.7 ± 0.1 oC biopolymers is reduced, and the hydrogen bonding between the
Tm = 6.2 ± 0.1 oC biopolymers becomes stronger. As discussed in previous sections,
10
these could induce the conformational transition of FG molecules
from random coli to helical structure, which was reported to occur
G' or G" (Pa)

Heating
at around 15 ◦ C for cold water fish gelatin [25]. The resulting FG
1 helices may participate in the formation of junction zones for ther-
moreversible gel networks, as known for mammalian gelatins [33].
Consequently, more and stronger inter- and intra-macromolecular
connectivities might be formed with the cooling, leading to the
0.1
G' (Cooling) Cooling
remarkable increase, first in viscous fluidity (G ), and then in elastic
G" (Cooling) solidity (G ), followed by the crossover of G and G , i.e., gelation of
G' (Heating)
the coacervate phase.
G" (Heating)
0.01 Decreasing the phase separation temperature from 40 to 10 ◦ C
0 10 20 30 40 was found to slightly increase the gelling point (Tg ) of the coacer-
Temperature ( C) o vate phase from 3.7 to 3.9 ◦ C and also increase the complex modulus
(G*) obtained at 3 ◦ C from 75.2 to 85.7 Pa (Fig. 6). This indicates
100 that the coacervate phase obtained at a lower temperature had a
G* = 79.9 ± 1.8 Pa (B) higher tendency for gelation and formed a more rigid gel, which
Tg = 3.8 ± 0.0 oC
may be also attributed to the reduced molecular mobility and the
Tm = 6.6 ± 0.0 oC stronger hydrogen bonding in the coacervate phase. In particular,
10
the coacervate phase separated at 10 ◦ C, which is below the random
coil-to-helix transition temperature of FG molecules, was found to
G' or G" (Pa)

Heating contain helical FG molecules in our FT-IR analysis, and therefore, the
1 cooling could easily associate the helices for the formation of junc-
Cooling tion zones, leading to the stronger and faster gelation, compared to
the coacervate phases separated at higher temperatures.
During the heating of the coacervate phases, the following
0.1
G' (Cooling) Heating sequential changes were observed in the moduli; (1) a crossover
G" (Cooling) of G and G between 6.2 and 6.9 ◦ C with sharp decreases in both G
G' (Heating) and G , (2) a sudden change of G from a steep to a gradual decrease
0.01
G" (Heating)
between 9.0 and 10.3 ◦ C, (3) a sudden change of G from a steep to
0 5 10 15 20 25 30 a gradual decrease between 9.8 and 10.6 ◦ C (not observed for the
coacervate phase separated at 10 ◦ C). The crossover temperature,
Temperature (oC)
at which G went below G , was defined as the melting point (Tm )
100 of the coacervate phase [32]. On the contrary to the case of cool-
G* = 85.7 ± 2.4 Pa (C) ing, the heating increases the molecular mobility and weakens the
Heating
hydrogen bonding between the biopolymers. These could loosen
Tm = 6.9 ± 0.1 oC
the junction zones of gel networks, stabilized by hydrogen bonding,
10
and shift the conformation of FG molecules from an ordered helical
to a disordered random coil state. Consequently, the heating might
G' or G" (Pa)

Tg = 3.9 ± 0.0 oC
reduce and weaken the inter- and intra-macromolecular connec-
1 Cooling
tivities, leading to the melting of the coacervate gel networks and
the dramatic losses in both elastic solidity (G ) and viscous fluid-
ity (G ). Decreasing the phase separation temperature from 40 to
10 ◦ C was found to increase the melting point (Tm ) from 6.2 to 6.9 ◦ C
0.1
G' (Cooling) (Fig. 6), also indicating that the coacervate phase separated at a
G" (Cooling) lower temperature formed a more rigid and thermo-stable gel.
G' (Heating) The Tm values were higher than the Tg values for all the three
G" (Heating)
0.01 coacervate phases due to the thermal hysteresis, which is a com-
2 4 6 8 10 monly observed indication of reluctance to the thermoreversible
o sol–gel transformation of a polymeric system. Decreasing the phase
Temperature ( C)
separation temperature from 40 to 10 ◦ C increased the difference
Fig. 6. Changes in storage (G ) and loss (G ) moduli during cooling and heating of between gelling and melting points (Tm –Tg ) from 2.5 to 3.0 ◦ C, also
FG–GA coacervate phases separated at 40 ◦ C (A), 30 ◦ C (B), and 10 ◦ C (C) at a rate indicating the formation of a more ordered and thermo-stable gel
of 0.016 ◦ C/min. The coacervate phases were separated from the FG–GA mixture by the coacervate phase separated at a lower temperature.
prepared at a total biopolymer concentration of 2% (w/w), FG:GA = 1:1, pH 3.5, and
It would be also worthwhile to examine the effect of GA, a
40 ◦ C. The moduli were monitored at ω = 2 rad/s and  = 1%, and complex modulus
(G*) was obtained at 3 ◦ C. nongelling biopolymer, on the sol–gel transformation. The temper-
ature dependence of G and G was also examined for the following
two controls: (1) an aqueous FG–GA mixture prepared at pH 8.0,
after the cooling started, (2) a rapid increase in G between 6.5 and where there is no significant FG–GA attractive interactions, with
7.7 ◦ C, (3) a second sharp increase in G between 5.2 and 5.9 ◦ C, and the same biopolymer compositions as the coacervate phase sepa-
(4) a crossover of G and G between 3.7 and 3.9 ◦ C. The crossover rated at 10 ◦ C (7.7% FG and 7.4% GA, Table 1), (2) an aqueous FG
temperature, at which G exceeded G , was defined as the gelling solution prepared at the same concentration as that in the coacer-
point (Tg ) of the coacervate phase [32]. The initial sharp increase vate phase separated at 10 ◦ C (7.7%). The FG–GA mixture prepared
M. Anvari et al. / International Journal of Biological Macromolecules 79 (2015) 894–902 901

at 10 ◦ C, having a FG:GA value of 1.0 (Table 1), at which the FG–GA


electrostatic interactions at pH 3.55 was reported to be the most
intense [3].

3.7. Viscoelastic behavior of FG–GA coacervate gel

Fig. 7 shows the frequency sweep (at  = 1%) of the three


FG–GA coacervate gels obtained by cooling the coacervate phases
separated at three different temperatures to 3 ◦ C. All the three coac-
ervate gels exhibited a slight increase in storage modulus (G ) and
a sharp increase in loss modulus (G ) with increasing angular fre-
quency (ω) (Fig. 7A). The values of G were higher than those of G
at any frequency, and the crossover of G and G was not observed.
Accordingly, the loss tangent (tan ı = G /G ), an indicative of liquid-
like behavior, showed a steep increase with frequency but did not
exceed unity (Fig. 7B). The results imply that the three FG–GA
coacervate gels can be classified as a weak physical gel formed by
noncovalent interactions, such as hydrogen bonding, hydrophobic
interactions, and electrostatic interactions [35], having a tendency
toward more liquid-like behavior at a higher frequency due to the
breakdown of network structure. It was also found that the gel
obtained from the coacervate phase separated at a lower temper-
ature had a smaller loss tangent at any frequency (Fig. 7B), also
demonstrating that reducing the phase separation temperature can
form a more rigid FG–GA coacervate gel network, as previously
discussed in the section of sol–gel transformation.

4. Conclusions

The present study demonstrated that decreasing the phase sep-


aration temperature from 40 to 10 ◦ C significantly influenced the
phase volume, composition, steady and oscillatory shear behaviors,
microstructure, and sol–gel transformation of FG–GA coacervate
Fig. 7. Storage (G ) and loss (G ) moduli (A) and tan ı (B) versus angular frequency phase, as well as the viscoelastic behavior of FG–GA coacervate
(ω) for the gels obtained by cooling the coacervate phases separated at 10, 30, or 40 ◦ C gel. The decrease in the temperature increased the volume frac-
to 3 ◦ C. The coacervate phases were separated from the FG–GA mixture prepared at tion of coacervate phase, and enhanced the partition of the two
a total biopolymer concentration of 2% (w/w), FG:GA = 1:1, pH 3.5, and 40 ◦ C. biopolymers, especially GA molecules, into the coacervate phase.
This is probably because the hydrogen bonding, which is known to
at pH 8.0 showed a higher gelling point (3.7 ◦ C) than the FG solution be majorly responsible for the reversible aggregation of biopolymer
(3.1 ◦ C), and also higher G values, indicating that the presence of complexes during the nucleation-and-growth like process of phase
GA molecules facilitated the gelation of FG molecules. This could separation, became stronger with decreasing the phase separa-
be explained that the volume exclusion effect between the two tion temperature, as shown by FT-IR spectroscopy. The coacervate
biopolymers, where the free volume of FG molecules is reduced phase became more viscous and more resistant to the structural
due to their thermodynamic incompatibility with GA molecules, deformation at low shear rates with decreasing the phase sepa-
increased the probability of association of FG molecules, and the ration temperature, but showed stronger shear-thinning behavior
resulting filler effect of the dispersed GA-rich phase reinforced the once the structural deformation and rearrangements occurred
FG network and increased the modulus [34]. above ˙ c . The decrease in phase separation temperature signif-
The FG–GA mixture prepared at pH 8.0 had a gelling point close icantly increased the G of coacervate phase in the frequency
to that (3.9 ◦ C) of the coacervate phase separated at 10 ◦ C, but much sweep test, and the phase separated at 10 ◦ C even showed a
lower G values. Its complex modulus (G*) at 3 ◦ C was measured to solid-like elastic behavior at a frequency of above 73 rad/s. In addi-
be 36.7 Pa, which was only 43% of that (85.7 Pa) of the coacervate tion, the coacervate phase was found to have more condensed
phase at 10 ◦ C. The results clearly demonstrated that the presence of microstructure at a lower temperature. The decrease in phase sepa-
GA molecules could significantly reinforce the gel network by their ration temperature increased the values of Tg , Tm , and (Tm –Tg ) but
attractive electrostatic interactions with FG molecules, much more decreased the tan ı of the FG–GA coacervate gels formed at 3 ◦ C,
than by their filler effect resulting from the thermodynamic incom- indicating that the coacervate phase separated at a lower temper-
patibility between the two biopolymers. In the coacervate phase, ature had a higher tendency for gelation and formed a more rigid
the GA molecules electrostatically interacted with FG molecules and thermo-stable gel. The major reasons for all the above results
seem to participate in the formation of gel network together with may be that decreasing the phase separation temperature reduced
FG molecules to form more compact network structure. the mobility of the biopolymers, increased the mass fraction of
The coacervate phase separated at 10 ◦ C showed a higher com- GA molecules, strengthened the hydrogen bonding between the
plex modulus (G*) at 3 ◦ C than the coacervate phases separated at biopolymers, and formed FG helical structure at 10 ◦ C, as shown
higher temperatures (Fig. 6), although it had a lower FG concentra- by FT-IR spectroscopy. The FG–GA coacervate gels were classi-
tion (Table 1). This may be attributed not only to the existence of FG fied as a weak physical gel formed by noncovalent interactions,
helices at 10 ◦ C, as previously mentioned, but also to the stronger and the presence of GA molecules was found to significantly rein-
FG–GA electrostatic interactions in the coacervate phase separated force the gel network probably due to their attractive electrostatic
902 M. Anvari et al. / International Journal of Biological Macromolecules 79 (2015) 894–902

interactions with FG molecules. The results obtained in the cur- [13] F. Weinbreck, R.H. Tromp, C.G. de Kruif, Biomacromolecules 5 (2004)
rent study provide basic knowledge necessary for the use of 1437–1445.
[14] M. Nigen, T. Croguennec, D. Renard, S. Bouhallab, Biochemistry 46 (2007)
FG–GA complex coacervation in many useful applications, such as 1248–1255.
microencapsulation, textural modification, emulsion stabilization, [15] A.B. Kayitmazer, S.P. Strand, C. Tribet, W. Jaeger, P.L. Dubin,
hydrogel formation, and meat analog development. Biomacromolecules 8 (2007) 3568–3577.
[16] O.H. Lowry, N.J. Rosebrough, A.L. Farr, R.J. Randall, J. Biol. Chem. 193 (1951)
265–275.
Acknowledgement [17] H. Espinosa-Andrews, O. Sandoval-Castilla, H. Vázquez-Torres, E.J.
Vernon-Carter, C. Lobato-Calleros, Carbohydr. Polym. 79 (2010) 541–546.
[18] D.J. Burgess, J.E. Carless, J. Colloid Interface Sci. 98 (1984) 1–8.
This research was supported by the Korea Sea Grant Program [19] M. Ahmad, S. Benjakul, Food Hydrocolloids 25 (2011) 381–388.
(Gangwon Sea Grant) funded by the Ministry of Oceans and Fish- [20] S. Benjakul, K. Oungbho, W. Visessanguan, Y. Thiansilakul, S. Roytrakul, Food
eries in Korea. Chem. 116 (2009) 445–451.
[21] J.H. Muyonga, C.G.B. Cole, K.G. Duodu, Food Chem. 86 (2004) 325–332.
[22] M.H. Uriarte-Montoya, H. Santacruz-Ortega, F.J. Cinco-Moroyoqui, O.
References Rouzaud-Sández, M. Plascencia-Jatomea, J.M. Ezquerra-Brauer, Food Res. Int.
44 (2011) 3243–3249.
[1] S.L. Turgeon, C. Schmitt, C. Sanchez, Curr. Opin. Colloid Interface Sci. 12 (2007) [23] K.J. Payne, A. Veis, Bipolymers 27 (1988) 1749–1760.
166–178. [24] Z. Dong, Q. Wang, Y. Du, J. Membr. Sci. 280 (2006) 37–44.
[2] F. Weinbreck, R. de Vries, P. Schrooyen, C.G. de Kruif, Biomacromolecules 4 [25] I.J. Haug, K.I. Draget, in: G.O. Philips, P.A. Williams (Eds.), Handbook of
(2003) 293–303. Hydrocolloids, 2nd ed., CRC Press, Boca Raton, USA, 2009, pp. 142–163.
[3] Y. Yang, M. Anvari, C.H. Pan, D. Chung, Food Chem. 135 (2012) 555–561. [26] F. Weinbreck, R.H.W. Wientjes, H. Nieuwenhuijse, G.W. Robijn, C.G. de Kruif, J.
[4] D. McClements, J. Biotechnol. Adv. 24 (2006) 621–625. Rheol. 48 (2004) 1215–1228.
[5] C. Schmitt, S.L. Turgeon, Adv. Colloid Interface Sci. 167 (2011) 63–70. [27] M.A. Rao, in: M.A. Rao (Ed.), Rheology of Fluid and Semisolid Foods, 2nd ed.,
[6] H.G. Bungenberg de Jong, in: H.R. Kruyt (Ed.), Colloid Science, vol. II, Elsevier, Springer, New York, USA, 2007, pp. 27–58.
Amsterdam, The Netherlands, 1949, pp. 232–258. [28] J. Ma, Y. Lin, X. Chen, B. Zhao, J. Zhang, Food Hydrocolloids 38 (2014)
[7] S. Harding, K. Jumel, R. Kelly, E. Gudo, J.C. Horton, J.R. Mitchell, in: K.D. 119–128.
Schwenke, R. Mothes (Eds.), Food Protein, Structure, and Functionality, VCH, [29] C. Schmitt, E. Kolodziejczyk, M.E. Leser, in: E. Dickinson (Ed.), Food Colloids:
Weinheim, Germany, 1993, pp. 216–226. Interactions, Microstructure and Processing, The Royal Society of Chemistry,
[8] Y. Fang, L. Li, C. Inoue, L. Lundin, I. Appelqvist, Langmuir 22 (2006) 9532–9537. Cambridge, UK, 2005, pp. 284–300.
[9] C. Schmitt, C. Sanchez, S. Desobry-Banon, J. Hardy, Crit. Rev. Food Sci. Nutr. 38 [30] X. Wang, J. Lee, Y.-W. Wang, Q. Huang, Biomacromolecules 8 (2007) 992–997.
(1998) 689–753. [31] Q. Ru, Y. Wang, J. Lee, Y. Ding, Q. Huang, Carbohydr. Polym. 88 (2012)
[10] C. Schmitt, C. Sanchez, A. Lamprecht, D. Renard, C.-M. Lehr, C.G. de Kruif, J. 838–846.
Hardy, Colloids Surf. B 20 (2001) 267–280. [32] M. Gudmundsson, J. Food Sci. 67 (2002) 2172–2176.
[11] S. Liu, Y.-L. Cao, S. Ghosh, D. Rousseau, N.H. Low, M.T. Nickerson, J. Agric. Food [33] C. Michon, G. Cuvelier, P. Relkin, B. Launay, Int. J. Biol. Macromol. 20 (1997)
Chem. 58 (2010) 552–556. 259–264.
[12] C. Schmitt, L. Aberkane, C. Sanchez, in: G.O. Philips, P.A. Williams (Eds.), [34] S.R. Monteiro, C. Tavares, D.V. Evtuguin, N. Moreno, J.A.L. da Silva,
Handbook of Hydrocolloids, 2nd ed., CRC Press, Boca Raton, USA, 2009, pp. Biomacromolecules 6 (2005) 3291–3299.
421–476. [35] M.H. Tunick, J. Agric. Food Chem. 59 (2011) 1481–1486.

Das könnte Ihnen auch gefallen