Anda di halaman 1dari 8

Food Hydrocolloids 39 (2014) 272e279

Contents lists available at ScienceDirect

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

Rheology and structure of mixed kappa-carrageenan/iota-carrageenan


gels
Tom Brenner a, b, *, Rando Tuvikene c, d, Alan Parker e, Shingo Matsukawa b,
Katsuyoshi Nishinari a, f
a
Graduate School of Human Life Science, Osaka City University, 3-3-138 Sugimoto, Sumiyoshi, Osaka 558-8585, Japan
b
Department of Food Science and Technology, Tokyo University of Marine Science and Technology, 4-5-7 Konan, Minato-ku, Tokyo 108-8477, Japan
c
Institute of Mathematics and Natural Sciences, Tallinn University, Narva mnt 29, 10120 Tallinn, Estonia
d
National Institute of Chemical Physics and Biophysics, Akadeemia tee 23, 12618 Tallinn, Estonia
e
Firmenich SA, Materials Science Department, Corporate Research Division, Meyrin 2, Geneva, Switzerland
f
Glyn O. Phillips Hydrocolloids Research Centre, School of Food and Pharmaceutical Engineering, Faculty of Light Industry, Hubei University of Technology,
Wuchang, Wuhan 430068, PR China

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

Article history: The rheology of mixed k-carrageenan/i-carrageenan gels was investigated in the presence of 75 or
Received 17 November 2013 150 mM Kþ. The addition of i-carrageenan to k-carrageenan led to an initial weak increase of the Young’s
Accepted 14 January 2014 modulus. A maximum occurred at concentrations ratios close to 1:1, followed by a weak decrease at
higher i-carrageenan concentrations. When k- and i-carrageenan were mixed at a constant total poly-
Keywords: saccharide concentration, a plateau of the Young’s modulus and a clear local minimum in the fracture
Carrageenan mixtures
strain in both compression and extension were identified at ratios between 2:1 and 4:3 (i:k). At all other
Mixed gels
ratios, the expected strong increase of the Young’s modulus and decrease of the fracture strain were
Rheological properties
Phase separation
observed with an increasing k-carrageenan fraction. A weak maximum of the fracture stress in
Bicontinuous structure compression and extension was observed at ratios between 1:1 and 1:2 (i:k). The structural and rheo-
logical findings support segregative phase-separation of mixed carrageenan gels in the presence of Kþ.
Phase separation could be promoted through curing of the gel network of i-carrageenan at temperatures
above the coil-to-helix transition of k-carrageenan.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction Nilsson, & Muhrbeck, 1992). For IC, on the other hand, only weak
dependence of the coil-helix transition temperature on the identity
Red sea weed galactans, i.e., carrageenans and agars, form the of the alkaline cation is observed, and no thermal hysteresis is
second largest group (weight-wise) of polysaccharides in the food present between the DSC endothermic and exothermic peaks,
industry (Piculell, 2006). The main carrageenans, k and i, hence- suggesting that the helices are not (or only very weakly) stabilized
forth designated KC and IC, respectively, are linear polyelectrolytes by aggregation (Piculell, 2006; Piculell et al., 1992; Ridout, Garza,
of different charge densities. Ideally, carrageenans have a disac- Brownsey, & Morris, 1996).
charide repeating unit, comprising a 1,3 linked b-D-galactopyr- In many applications, a mixture of carrageenans is present, and
anose (G) and a 1,4 linked 3,6-anhydro-a-D-galactopyranose (AG). in general, KC and IC powders (especially food-grade) can contain
KC is sulphated at O-4 of the G residue, and IC carries one additional non-negligible contaminations of the other (Parker, Brigand,
sulphate group on O-2 of the AG residue. The different structures of Miniou, Trespoey, & Vallee, 1993; Piculell et al., 1992; Rochas,
KC and IC lead to very different gelling behaviours. The helix for- Domard, & Rinaudo, 1980; Rochas, Rinaudo, & Landry, 1989; van
mation and subsequent helix aggregation of KC is strongly pro- de Velde, Peppelman, Rollema, & Tromp, 2001). These impurities
moted by the larger alkaline cations Kþ, Rbþ and Csþ, and especially are present as entire chains of the contaminating carrageenan
at low ionic strength, by divalent cations (Piculell, 2006; Piculell, (Rochas et al., 1989) or as blocks of the contaminating carrageenan
within a chain of the predominant carrageenan (van de Velde et al.,
2001), forming a KC/IC hybrid.
* Corresponding author. Department of Food Science and Technology, Tokyo
University of Marine Science and Technology, 4-5-7 Konan, Minato-ku, Tokyo 108-
The more extensive aggregation of the helices of KC, compared
8477, Japan. Tel.: þ81 3 5463 0581. to that of IC helices, causes significant differences in the properties
E-mail address: physicalchemistrytom@gmail.com (T. Brenner). of their gels. In general, KC gels are hard and brittle, whereas IC gels

0268-005X/$ e see front matter Ó 2014 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.foodhyd.2014.01.024
T. Brenner et al. / Food Hydrocolloids 39 (2014) 272e279 273

are much weaker and fracture at larger deformations. Because of AG content was determined colorimetrically using a resorcinol-
the huge difference in elastic properties of KC and IC in the presence acetal method (Yaphe & Arsenault, 1965) with fructose as a stan-
of Kþ, rheological measurements have been suggested as a cheap dard sugar. Galactose content was obtained by reductive hydrolysis
alternative to more expensive 13C NMR determination of the of the polysaccharides followed by the acetylation process (Jol,
amount of KC contamination in IC samples (Parker, 1995). In mix- Neiss, Penninkhof, Rudolph, & De Ruiter, 1999); the obtained aldi-
tures containing Kþ, KC and IC show independent DSC endotherms tol acetate derivatives were analysed by gas chromatography
and exotherms (Ridout et al., 1996), and these correspond to in- coupled with flame ionization detection (GC-FID) using inositol as
dependent gel networks as observed rheologically (Parker et al., an internal standard (Stevenson & Furneaux, 1991). The ratio of G to
1993). Both these results provide evidence that the two gel net- AG units was 1.03 and 1.15 in the KC and IC samples. The lower AG
works are uncoupled. However, probably owing to the huge dif- content in the IC powder is a general feature of IC preparations, as
ference in elastic moduli between pure KC and IC gels in the they are known to contain non-trivial amounts of substituted gal-
presence of Kþ, determining whether the KC and IC elastic net- actopyranosyl units, carrying mainly methyl and pyruvate groups
works in mixed gels are phase-separated or interpenetrated has (see, e.g., Tuvikene et al., 2009; Usov, 2011), which we did not
proven rather difficult, with differing views expressed based on not quantify. We note that the presence of only trace amounts of G-
incompatible sets of data (Parker et al., 1993; Piculell et al., 1992; 2S,6S, i.e., the n-form carrageenan, could be inferred from the 1H
Ridout et al., 1996). NMR spectra. The total polysaccharide content (without counter-
Given the importance of KC/IC mixtures in applications, we ions), recalculated assuming idealized structure with the ratio of i
extensively map the small- and large-deformation rheological and k diads estimated from NMR, was 64.8% and 71.9% in the KC and
properties of their mixed gels, using commercial food-grade sam- IC samples, respectively. The moisture content of these powders
ples. Our results are compared to the literature, where different was respectively 8.7% and 10.3%.
investigators have mooted either a phase-separation of the two Inductively coupled plasma atomic absorption elemental anal-
polymers (Parker et al., 1993; Piculell et al., 1992) or their inter- ysis was carried out on an SPS7800 Plasma Spectrometer (SII
penetration (Ridout et al., 1996) as structural possibilities. The Group). About 0.1 g of powder was microwave-digested in 5 mL
structure of the mixed gels affects not only their rheology, but also nitric acid and 5 mL hydrogen peroxide, then diluted x10 with
the release of water or flavour compounds, and textural properties water and x5 with a mixture of 1/9 nitric acid/water. Contents of
that are difficult or impossible to relate to rheology (van den Berg, Naþ, Kþ, Ca2þ, Mg2þ and S are given in Table 1.
van Vliet, van der Linden, van Boekel, & van de Velde, 2007a; Chen, Size exclusion chromatography (SEC) of polysaccharides was
2009; Foegeding et al., 2011), like creaminess and spreadability. The performed on a chromatograph equipped with a PerkinElmer Se-
high permeability of bicontinuous structures improves flavour- ries 200 pump, a Knauer Smartline 2300 refractive index detector, a
release that is mediated through the release of solvent (van den Knauer Smartline column thermostat, and two Shodex OHpak SB-
Berg et al., 2007a; van den Berg, van Vliet, van der Linden, van 806MHQ columns in series. Elution was carried out using a 0.1 M
Boekel, & van de Velde, 2007b), and the emulsion-like (“water-in- NaNO3 solution as the mobile phase at a flow rate of 0.8 mL/min.
water emulsion”) structure of filled phase-separated gels contrib- The temperature of the columns was maintained at 60.0  C. A
utes to their creaminess (e.g., Le Révérend, Norton, Cox, & calibration curve was constructed using 12 pullulan standards
Spyropoulos, 2010). It is therefore clear that further evidence (peak molecular weight Mp values 2400, 1560, 710, 380, 200, 106,
regarding the structural transition in mixed KC/IC gels and the 45.9, 22.0, 11.2, 5.6, 1.08 and 0.342 kDa). The elution volume was
respective structures at high and low i:k ratios holds potential in- corrected to the internal marker of ethylene glycol (0.01% in sam-
terest for the food industry. ple) at 23.04 mL. The polymer concentration used was 0.07%, and
the sampling volume 100 mL. Powders were dissolved in the SEC
2. Materials and methods eluent solvent by vigorously stirring for 10 min at the boiling point.
The hot (60  C) sol was filtered through a 0.45 mm RC membrane,
2.1. Ingredients allowed to cool down, and then injected into the HPLC system.
Fig. 1 shows the SEC elution profiles of the samples. The peak
Food-grade powders of k-carrageenan (SAN SUPPORT Ò G-16) molecular weight values (Mp), relative to those of the pullulan
and i-carrageenan (SAN SUPPORT Ò G-17) were provided by San Ei standards, were 1860 and 1435 kDa for KC and IC, respectively.
Gen F.F.I. (Osaka, Japan) and used without further purification. KCl Another finding of note regarding the KC sample is the large (15%)
was purchased from SigmaeAldrich. peak for Cl (at 21.37 mL), which, based on the data of Table 1, is in
excess of the theoretical limit even if no other anions are present.
2.2. Chemical analysis This peak, however, decreases as a function of keeping time at high
temperature, indicating that some heat sensitive oligosaccharides
1
H NMR spectroscopic analyses were performed on a Bruker are eluted at the same retention volume as Cl. Our mono-
AVANCE III spectrometer operating at 800 MHz. The 1H NMR saccharide analysis also detected trace amounts of glucose in the KC
spectra from 0.5% polysaccharide solutions in D2O (w/w) were powder, supporting the conclusion that some oligosaccharide has
obtained at 65  C and 64 transients were collected with an inter- been added to the sample. An important finding of the IC powder is
pulse delay of 5 s. Sodium 4,4-dimethyl-4-silapentane-1-sulfonate the presence of excess SO2 4 (2.2%; retention volume 20.33 mL), so
served as an internal reference (d ¼ 0 ppm). The molecular weight
of the polysaccharide samples was reduced prior to the NMR
measurement by sonicating in a water-bath at 25  C for 100 min Table 1
Elemental analysis results of KC and IC powders.
with an ultrasound homogenizer (Bandelin Sonopuls UW2070,
70 W, operating at 80%), followed by centrifugation and freeze- Element IC (%) KC (%)
drying. The molar ratio of KC and IC was estimated by integrating Na 0.84 0.32
the anomeric signals corresponding to AG residues from the K 3.3 9.1
respective diads. Our NMR results indicated that, in molar terms, Ca 3.3 0.88
the KC sample contains 80.5% KC and 19.5% IC, whereas the sample Mg 0.21 0.18
S 13.1 7.7
of IC contains only 1.7% KC.
274 T. Brenner et al. / Food Hydrocolloids 39 (2014) 272e279

different parameters were chosen for technical reasons, and cy-


21.37 lindrical gel samples (height ¼ 9.0 mm, diameter ¼ 11 mm) were
23.04 compressed (10 mm/s, strain rate 1.1 s1) to fracture using a square
PMMA plunger. Ring extension was as described elsewhere
(Brenner, Wang, Achayuthakan, Nakajima, & Nishinari, 2013). En-
gineering fracture strain and stress were calculated from the
compression test (εfc and sfc), while for the ring extension test the
engineering fracture strain and stress (εfe and sfe) were approxi-
A
mated from the elongation of the inner side of ring as described
elsewhere (Brenner et al., 2013).
22.33

20.33 2.6. Rotational rheometry


B
The shear storage (G0 ) and loss (G00 ) moduli were measured on a
10 15 20 25 Haake Mars II rheometer (Thermo Scientific) using a chemically
roughened parallel plate geometry (diameter 35 mm). Contact with
Retention volume, mL
the sample was ensured through use of the Autotension option.
Fig. 1. Size exclusion chromatograms of KC (A) and IC (B) samples obtained at 60  C.
2.7. Turbidity measurements
we can conclude the presence of 13.1%e0.7% ¼ 12.4% S in the IC
polysaccharide. The turbidity of gels (300e1000 nm) was determined using a V-
630 Biospectrophotomer (Jasco, Tokyo, Japan) using 10 mm quartz
2.3. Preparation of the gels cells. Aqueous 0.15 M KCl solution was used as a blank.

Potassium chloride and carrageenan powders were added in 2.8. Differential scanning calorimetry
this order under stirring at room temperature. The mixtures were
heated with stirring for 120 min at 40  C and 30 min at 90  C. The Differential scanning calorimetry (DSC) was performed on a
stirring was stopped for the last 15 min of the heating procedure to Micro-DSC VII (Setaram, Caluire, France). Samples (0.8 mL) were
allow any trapped air to escape. Following heating, the hot solu- loaded into Hastelloy cells, heated to 85  C to erase the thermal
tions were poured into moulds of different dimensions and then history, and then cooled to 5  C at 0.5  C/min to identify the sol-to-
kept in a refrigerator (5  C) for 16e20 h, except for experiments on gel transition temperatures of the two polysaccharides.
the effect of the curing method. In these experiments, the mould
with hot solution was immediately placed in a water bath kept at 3. Results and discussion
52  C for a fixed time, and then immersed in ice water for 60 min
before being placed in a refrigerator (5  C) for 16e20 h. Gels were 3.1. Small and large deformation rheology
equilibrated to room temperature (20  2  C) (about 60e120 min)
before measurement. MQ-grade water was used, and the total Fig. 2 shows the Young’s modulus, E, for KC:IC mixtures as a
concentration of Kþ was fixed to 75 or 150 mM. function of the IC weight fraction, cIC ¼ CIC/(CIC þ CKC), in the
presence of 150 mM Kþ and at a fixed total polysaccharide content
2.4. Linear extensional rheology e Young’s modulus CIC þ CKC ¼ 1.6 wt%. A near-plateau of E is observed at IC:KC ratios of
2:1 to 4:3, i.e., cIC z 0.55e0.65. Fig. 3A and B shows the fracture
The complex Young’s modulus of the gels was measured as strain and stress of these mixed gels in compression and extension,
described by Nishinari et al. (1980). Cylindrical gels
(height ¼ 31 mm, diameter ¼ 20 mm) were glued to the upper and
bottom faces of a Rheolograph Gel (Toyo Seiki Seisakusho, Tokyo,
Japan) with alpha cyano acrylate glue (Aron Alpha; Toa Gosei Ltd.,
Japan). The Rheolograph measures the storage (E0 ) and loss (E00 )
Young’s moduli at frequency of 3 Hz and an amplitude of 0.1 mm,
which for these gel specimens corresponds to a strain of 0.32%. For
all gels it was found that E0 z E, while E00 was not reliably deter-
mined for most gels, i.e., it was lower than or comparable to the
lower limit of the Rheolograph, 0.1 kPa. Therefore, we will only
discuss E.

2.5. Mechanical tests

Compression and extension tests were carried out on an XT.T2


Texture Analyser (Stable Micro Systems, Surrey, UK), or on a RCT-
2002D-D Fudoh Rheometer (Rheotech, Tokyo, Japan). Lubrication
in compression was not necessary because the gels exude water. For
compression tests using the Texture Analyser, cylindrical gel sam- Fig. 2. Young’s modulus E (circles) as a function of cIC ¼ CIC/(CIC þ CKC) for a fixed total
carrageenan content of 1.6% and [Kþ] ¼ 150 mM. Dotted: upper (isostrain, x in Eq.
ples (height ¼ 15 mm, diameter ¼ 25 mm) were compressed (1) ¼ 1) bound, solid: lower (isostress, x ¼ 1) bound, dot-dashed: additivity, short
(10 mm/s, strain rate 0.67 s1) to fracture using a square aluminium dashed: bicontinuous (x ¼ 0.2), long dashed: fit to Eq. (1) with x ¼ 0.46. Error-bars
plunger. For compression tests using the Fudoh Rheometer, (from two preparations) are too small to be seen.
T. Brenner et al. / Food Hydrocolloids 39 (2014) 272e279 275

Fig. 3. Engineering fracture strain (open symbols) and stress (filled symbols). A, compression, B, extension. Error-bars are calculated from at least 4 samples from each of 2
preparations.

respectively, and demonstrate that at the same ratio range, a clear replaced 25%, 50% and 75% of the IC content with KC, in the pres-
local minimum of the fracture strain is observed, and as a result, ence of 0.08 M KCl. As in the other two studies, the data of Ridout
this is where the strongest decrease of the fracture stress with et al. (1996) lay between the upper and lower bounds predicted by
increasing IC content is observed. When a composite gel contains Eq. (1). However, because no obvious phase-inversion was indi-
two gelling components, the structure formed may comprise a cated by their data, Ridout et al. (1996) concluded that an inter-
single coupled network, two interpenetrating networks or two penetrating KC/IC network was formed. At this point we should
phase-separated networks (e.g., Morris, 1995). This simplified tax- clarify the distinction between the bicontinuous and inter-
onomy does not take into account the distribution of each penetrating cases for the networks of mixed carrageenans. Obvi-
component in different phases, phase continuity/connectivity, ously, each of the elastic networks is formed by phase-separating
different interactions (other than direct binding) or swelling/ from both the solvent and the other polysaccharide, as double-
deswelling. There are different models for estimating the elastic helices are formed strictly from two chains of the same poly-
modulus of the composite gel G0 from the individual moduli Gi and saccharide, and junction-zones of double-helices include only one
volume fractions Øi. For phase-separated networks, the elastic type of polysaccharide. The question to be answered is what the
contributions of each phase may be added arithmetically (isostrain) scale of phase-separation, or inhomogeneity, between the two
or harmonically (isostress) to give theoretical upper and lower polysaccharide networks is. In the limit of interpenetrating net-
bound estimates (e.g., Ross-Murphy, 1995): works, it is assumed that the density of each network is the same as
that in absence of the other network, i.e., no densification takes
1=x place, and a junction-zone of one network will be on average at the
G0 ¼ B1 Gx1 þ B2 Gx2 (1)
same distance from a junction-zone of the other polysaccharide
The exponent x determines the behaviour of the composite: the network as from another junction-zone of its own network when
values x ¼ 1 and x ¼ 1 correspond to the upper (isostrain) and the total density of junction-zones of both networks is equal. When
lower (isostress) bounds, respectively, which were suggested for one network causes densification of the other but both networks
filled gels where only one phase is continuous (the other one being keep their continuity, we can speak of phase-separation. In this
a filler). Isostrain behaviour is expected when the continuous phase case, at equal total junction-zone density of both networks, a
has the higher elastic modulus, as the deformation will be deter- junction-zone of one network will be on average closer to a
mined by the network. On the other hand, when the filler is more junction-zone of its own network than to a junction-zone of the
rigid than the network, its deformation is lower than that of the other networks, thus leading to well-defined areas (phases) where
network, and the limit of equal stress in both phases is more one network is concentrated and the other is depleted, on a length
appropriate (Ross-Murphy, 1995). When both networks are scale larger than that of inhomogeneities of either network. Of
continuous, i.e., for the bicontinuous case, Davies (1971) suggested course, the two cases of interpenetrating and phase-separated
the exponent x ¼ 1/5. networks are approximate limiting cases, while the reality of dis-
When two independent networks span the entire volume and tribution of the two networks can be much more complicated (see,
do not phase-separate on a length scale larger than the scale of e.g., discussion in Ross-Murphy, 1995).
heterogeneities in each network (the interpenetrating case; see Our data do not contradict the results reported by Parker et al.
discussion below), the starting point for calculating G0 is assuming (1993), Piculell et al. (1992) and Ridout et al. (1996), and
additivity, i.e.: resemble rheological data by Rochas et al. (1989), who, like us,
prepared mixed gels using a nearly pure i-form and a substantially
G0 ðC1 þ C2 Þ ¼ G1 ðC1 Þ þ G2 ðC2 Þ (2) impure (11% i-impurity) k-form. As in our Fig. 2, they reported a
monotonous decrease of the elastic modulus with increasing i-
where C1 and C2 are concentrations of the individual gelling fraction, however, did not identify a near-plateau as in our case.
(polymer) components (Zhang & Rochas, 1990). Furthermore, Rochas et al. (1989) also found, as in our Fig. 3A, a
Piculell et al. (1992) and Parker et al. (1993) substituted KC for maximum of the fracture stress in compression at about equal
up to 10% of the IC content in the presence of 0.1 M KCl and 0.2 M mixing ratios of the two polysaccharides.
NaCl, respectively, and concluded that the gels were phase- Our results indicate that the elastic modulus is indeed close to
separated, because the small KC fraction affected the rheology that predicted from simple additivity at low i:k ratios, but the effect
much more than expected if it were not excluded from the large of KC is far greater than predicted by additivity at low kappa con-
volume occupied by the dominant iota fraction. Ridout et al. (1996) tents. Without over-emphasizing the theoretical values for the
276 T. Brenner et al. / Food Hydrocolloids 39 (2014) 272e279

interpenetrating networks and bicontinuous cases as presented in attribute this maximum to both the opposing trends of the elastic
Fig. 2, we can say that both models predict the data better than modulus and the fracture strain and to the slightly stronger strain
either the lower or upper bounds. However, while the inter- hardening. We do not consider the data of Fig. 4 to contradict phase
penetrating networks and bicontinuous cases are both close to the separation of the two polysaccharides in the gel phase, as to our
data at low i:k ratios, only the bicontinuous model seems plausible knowledge, no universally accepted model has been put forth that
at high IC content. Indeed, the interpenetrating limit leads to a explains the large deformation rheology of mixtures.
minimum of the elastic modulus at high IC content, but we, like The Young’s modulus E of mixed gels as a function of KC and IC
Rochas et al. (1989), found that the elastic modulus monotonously concentrations is plotted in Fig. 5A. This figure clearly demonstrates
decreased with increasing IC content, with a weak plateau at the higher moduli of KC gels in comparison with IC gels, but also
around ratios of 2:1 to 4:3 (i:k). As shown in Fig. 2, the monotonous shows that when IC is added to KC at a constant KC concentration,
form Eq. (1) provides a satisfactory fit to the data, except for around an initial increase is observed, with a subsequent peak and an
the plateau, where the fit overestimates the data. The monotonous eventual weak decrease. This leads to a certain plateau in the 3D
change in the elastic modulus points to the bicontinuous phase, plot when following a straight line connecting points of the same
even if the exponent of the fit in Eq. (1) is different to that predicted value on the (positive) CKC and CIC axes, i.e., straight lines for con-
by (Davies, 1971). Furthermore, the transition seen in our data at stant total polysaccharide concentrations and different IC:KC ratios.
ratios of 2:1 to 4:3 (i:k) is indicated not only from measurements of This plateau appears at the same ratio range seen in Fig. 2, namely,
the elastic modulus (Fig. 2), but also from fracture measurements in about 4:3 to 2:1 (i:k). The findings of Fig. 5 are also indicative of a
compression and extension (Fig. 3). While KC forms relatively certain structural transition between low and high i:k ratios. The
brittle and strong gels in the presence of KCl, IC forms weaker gels fracture parameters from both compression and extension also
with higher fracture strains. The increase in E and decrease in support this conclusion, but because their three-dimensional rep-
fracture strain with a decreasing IC:KC ratio are therefore expected. resentation tends to be undescriptive, we chose to show only the
However, it is very difficult to explain the local minimum in the fracture stress in compression, see Fig. 5B. At constant KC concen-
fracture strain in compression and the near-plateaus in E and trations, the fracture stress goes through a maximum at around
fracture strain in extension around ratios of 2:1 to 4:3 (i:k) without equal additions of IC, emphasizing the general nature of this
invoking some structural transition. The presence of a transition finding, which was first noted from Fig. 3 at a fixed total carra-
could be an indicator of phase-separation. The degree of phase- geenan content.
separation is such that it affects the rheological properties in a Eq. (1) can be applied to binary mixtures of arbitrary concen-
manner that negates the assumption that neither network is trations of the two components, but this requires knowledge of the
affected by presence of the other, as in the interpenetrating model. elastic modulus of the pure components over the entire range of
We are therefore led to conclude the presence of sufficient densi- concentrations. In this study, we found the elastic modulus of KC to
fication of either polysaccharide network by the other, so that we scale as a power law of the KC concentration:
speak of “phase separation”.
2:55
Regarding fracture of the mixed gels in compression, Rochas EKC fCKC (3)
et al. (1989) claimed (without showing the data) that the fracture
strain was identical for all mixing ratios, which is in clear Because for all relevant concentrations, the elastic modulus of
disagreement with our data (Fig. 3A). In our case, the maximum of KC is much higher than that of IC, the elastic modulus will be
the fracture stress can be understood to arise from the two dependent only on the volume fraction and elastic modulus of the
opposing trends of increasing fracture strain and decreasing elastic KC phase, especially if Eq. (1) is valid with a sufficiently high value
modulus with increasing cIC. From the report by Rochas et al. of the exponent x. As we saw from the data obtained at a constant
(1989), however, it must have been concluded that the strain total polysaccharide content (Fig. 2), the best fit to Eq. (1) is ob-
hardening of the gels in compression increases with increasing tained with x ¼ 0.46. If the elastic modulus is determined by the
content of i-carrageenan, because otherwise the fracture stress volume fraction and elastic modulus of KC, then it is clear that x
could not increase with increasing i-carrageenan content at a should be larger than 1/2.55 z 0.4 to allow for an increase in the
constant fracture strain with decreasing elastic modulus. All our elastic modulus with increasing IC concentration at a fixed KC
gels showed similar strain hardening, although at cIC ¼ 0.5, strain concentration:
hardening was indeed slightly stronger, see Fig. 4. While Rochas   x 1=x
et al. (1989) concluded from the maximum of the fracture stress E ¼ ð1BKC ÞA ð1BKC Þ2:55 þBIC EIC
x

that interaction must be present between k- and i-chains, we   x 1=x


z ð1BKC ÞA ð1BKC Þ2:55

¼ A1=x ð1BKC Þð12:55xÞ=x (4)

where A is a constant. Thus, Eq. (1) with x ¼ 0.46, which is slightly


higher than 1/2.55, can explain the slight increase in E with
increasing CIC at a constant CKC, although it cannot explain the
maximum at around equal concentrations of KC and IC or the
subsequent decrease in E with increasing CIC.
For mixtures of agarose and KC, Zhang and Rochas (1990)
concluded the presence of an interpenetrating network, but their
investigation was mainly structural rather than rheological. As
expressed by Parker et al. (1993), it is interesting if the KC network
is indeed phase separated from that of its closest relative, IC, but is
Fig. 4. Engineering stress s, normalized with the Young’s modulus E and the engi-
miscible with the network of its more distant relative agarose. On
neering strain ε, as a function of the strain in compression. cIC values are indicated in the other hand, Piculell et al. (1992) noted that the separation of the
the figure. KC and IC networks is reasonable as they are both anionic
T. Brenner et al. / Food Hydrocolloids 39 (2014) 272e279 277

Fig. 5. Young’s modulus E (A) and fracture strain in compression sfc (B) as a function of KC and IC concentrations in the presence of 0.15 M Kþ.

polyelectrolytes, and that KC should indeed be closer to complete 19.5 mol% IC, the transition suggested from Figs. 2e4 around
miscibility with the neutral agarose. We also note that mainly cIC z 0.6 suggests a molar ratio close to 2:1 i:k.
based on structural investigations, KC was also concluded to form a
phase separated gel with gellan, another anionic polysaccharide 3.2. Structural investigations
(Nishinari, Watase, Rinaudo, & Milas, 1996), and parenthetically
note that the rheological additivity was not indicated correctly in To substantiate the presence of a structural transition, we
Fig. 6 of that publication. Our results indicate that a phase- measured the turbidity of gels in the UV-VIS range. While the
separated bicontinuous structure is formed in the KC/IC mixtures turbidity is a measure of structural homogeneity, it is not often
at high i:k ratios. While strictly, the equation suggested by Davies appreciated that many parameters determine the turbidity, among
(1971) for the elastic modulus of the bicontinuous structure un- which are the concentration of scattering units, the contrast be-
derestimates the values we find experimentally at low i:k ratios, we tween the scattering units and the medium (in this case, the sol-
recognize that conceptually, as an intermediate case between the vent), the length scale of concentration fluctuations and the
upper and lower bounds of the phase separated system, it explains compressibility of the system (see, e.g., De Kruif, 1993; Pouzot,
the values of the elastic modulus at all ratios of the two carra- Durand, & Nicolai, 2004), and in our case, also direct absorption
geenans. The slight underestimate of Davies’s (1971) model can be of light, and all of these factors have different wavelength de-
eliminated by using a higher exponent x in Eq. (1), as was shown in pendencies. For this reason, instead of looking at the turbidity at a
Fig. 2. As explained above, this higher exponent (x ¼ 0.46) can also single wavelength, we decided to search for clues of structural
explain the slight increase in the elastic modulus with increasing IC changes by calculating the ratio of the turbidity at two different
content at fixed CKC, but it cannot explain the peak and eventual wavelengths. Because we are not using any models to quantify the
decrease in E with further increase of CIC. turbidity, using such ratios of the turbidity offers no advantage over
The presence of a bicontinuous structure at all i:k ratios is the more convenient fractional ratio R, where we fix the ratio for
consistent with the phase-separation best indicated by the strong one pure component to zero and for the other pure component to
effect of low concentrations of KC on the elastic modulus of IC gels, unity:
and is also consistent with a structural transition between high and
low ratios i:k, reflected in both small and large deformation tests. si s
i ðICÞ
sj  s ðICÞ
We note that as our sample of KC contains 80.5 mol% KC and Ri=j ¼ s ðKCÞ j s ðICÞ (5)
sj ðKCÞ  sj ðICÞ
i i

where i and j refer to two wavelengths, si is the turbidity at


wavelength i, and s(KC) and s(IC) are the turbidities of the pure KC
and IC samples, respectively. The advantage of this fractional ratio
over the turbidity ratios themselves is that curves for different
combinations of wavelengths can be compared directly.
Fig. 6 shows the fractional ratios of the turbidity for
400&800 nm (open symbols) and 300&900 nm (closed symbols).
Our turbidity measurements at these two different combinations of
wavelengths show no evidence for the structural transition close to
cIC z 0.6 implied by our rheological measurements. We hope that
more detailed spectroscopic measurements we are making, on
which we will report separately, will resolve this enigma.

3.3. Controlling the structure through the curing method

As should be clear from the previous sections, our suggestion of


a bicontinuous structure of k- and i-carrageenan networks is based
Fig. 6. Fractional ratios of turbidity R; open symbols, R400/800, closed symbols, R300/ on rheological findings, as we have failed to produce convincing
900. Error-bars indicate one standard deviation. structural results to support this conclusion. However, as noted by a
278 T. Brenner et al. / Food Hydrocolloids 39 (2014) 272e279

k-carrageenan network. In theory, promoting the network forma-


tion of i-carrageenan before the k-carrageenan is allowed to gel
should lead to phase-separation on a larger length scale. In practice,
testing this idea becomes more feasible because the curing of IC is
very slow, as noted by Nono, Durand, and Nicolai (2012). We per-
formed DSC and rotational rheometry tests on a mixture
(CKC ¼ 0.4%, CIC ¼ 1.2%) containing 75 mM Kþ. The reason for the
lower potassium concentration chosen was to allow a larger sepa-
ration between the gelation temperatures of the i and k forms, as
the gelation temperature of the latter strongly increases with
increasing [Kþ] while the gelation temperature of the former is
hardly affected (see, e.g., Ridout et al., 1996). Our data clearly show
the distinct gelations of the two polysaccharides, with two separate
exotherms and two increases of G0 . The data are of course not novel,
as the distinct gelation of the i and k forms was demonstrated by
Fig. 7. Storage shear modulus G0 (circles, g ¼ 0.02, u ¼ 1 rad s1) and DSC heat flow other researchers using rotational rheometry (Parker et al., 1993)
(solid line) during cooling. The arrow indicates exothermic heat flow direction. and DSC (Ridout et al., 1996). The data enabled us, however, to
CKC ¼ 0.4%, CIC ¼ 1.2%, [Kþ] ¼ 75 mM. The cooling rate was 0.5  C/min in both
choose an optimal temperature for the curing experiments to
experiments.
promote IC gelation while KC chains are still in the coil conforma-
tion. Based on the data shown in Fig. 7, we chose to cure samples at
52  C for different periods of time, and, to maximize the effect of
the curing, we then quenched the samples in ice-water at 0  C, to
allow as fast a curing as possible of the k-carrageenan. After 60 min
at 0  C, gels were transferred to a refrigerator (5  C) for curing over
16e20 h. We found, however, that to obtain reliable stress-
relaxation data with the equipment available to us we needed to
prepare gels at higher KC concentrations than that used for
obtaining the data of Fig. 7. We therefore opted for a concentration
of 1.2% KC with 0.8% IC while keeping [Kþ] ¼ 75 mM. The stress
relaxation curves of this gel are shown in Fig. 8. We found that to
obtain clean signals a compression strain of 15% was needed. Ac-
cording to our preliminary tests, this strain is outside the linear
regime of the gels investigated. The same results could be obtained
using lower strains, but the signals tended to be noisier. From Fig. 8,
Fig. 8. Stress relaxation curves of a mixed gel (CKC ¼ 1.2%, CIC ¼ 0.8%, [Kþ] ¼ 75 mM) at we clearly see that increasing the curing time at 52  C had a weak
15% compression. Samples were cured at 52  C (curing time in min indicated in figure)
but consistent effect to increase the effective instantaneous
before being quenched in ice water; C indicates the control sample cured at 5  C.
modulus (i.e., the stress/strain at t ¼ 0) of the mixed gel. This
finding serves, therefore, as another evidence of phase separation;
reviewer of this paper, as i-carrageenan gels at a higher tempera- enhancing the phase separation by allowing selective curing of one
ture than k-carrageenan (unless the concentration of cations that network before the other affects the rheology. This effect is rather
promote k-carrageenan aggregation is very high), it should be weak, as could be expected from the data of Fig. 5A and the prior
possible to allow different degrees of formation of the i-carra- discussion of the Eq. (1) exponent x ¼ 0.46 (see Fig. 2), which was
geenan elastic network by keeping the mixtures at a temperature noted to be only slightly higher than 1/2.55 z 0.4. We will
between the gelation temperatures of the two polysaccharides comment parenthetically on the effect of the cooling rate on the
for varying lengths of time before inducing gelation of the rheology of the gel. Logically, the lateral aggregation of KC is

Fig. 9. Schematic representations of the bicontinuous structure suggested for KC/IC mixed gels. Left panel, structure obtained following fast curing; right panel, enhanced phase
separation following curing of IC at temperatures above the coil-to-helix transition temperature of KC.
T. Brenner et al. / Food Hydrocolloids 39 (2014) 272e279 279

arrested by gelation. Therefore, slower curing of the gel should Jol, C. N., Neiss, T. G., Penninkhof, B., Rudolph, B., & De Ruiter, G. A. (1999). A novel
high-performance anion-exchange chromatographic method for the analysis of
allow for more extensive KC aggregation that in turn will lead to a
carrageenans and agars containing 3,6-anhydrogalactose. Analytical Biochem-
higher elastic modulus. As seen from Fig. 8, the control sample that istry, 268(2), 213e222.
was cured in a refrigerator at 5  C, showed a higher effective Le Révérend, B. J. D., Norton, I. T., Cox, P. W., & Spyropoulos, F. (2010). Colloidal
modulus at t ¼ 0 than all the gels that were quenched in ice water, aspects of eating. Current Opinion in Colloid & Interface Science, 15(1e2), 84e89.
Morris, V. J. (1995). Synergistic interactions with galactomannans and gluco-
which is what we expect for a gel cured using a lower cooling rate. mannans. In S. E. Harding, S. E. Hill, & J. R. Mitchell (Eds.), Biopolymer mixtures
Our last conclusion, inferred from the findings of Fig. 8, can be (pp. 289e314). Nottingham: Nottingham University Press.
summarized schematically as in the drawings of Fig. 9. In the left Nishinari, K., Horiuchi, H., Ishida, K., Ikeda, K., Date, M., & Fukada, E. (1980). A new
apparatus for rapid and easy measurement of dynamic viscoelasticity for gel-
panel, we drew the bicontinuous structure we consider to be like foods. Journal of the Japanese Society for Food Science and Technology-
formed in the KC/IC mixed gels. In the right panel, the extent of the Nippon Shokuhin Kagaku Kogaku Kaishi, 27(5), 227e233.
phase separation is enhanced by an initial curing of IC at temper- Nishinari, K., Watase, M., Rinaudo, M., & Milas, M. (1996). Characterization and
properties of gellan-kappa-carrageenan mixed gels. Food Hydrocolloids, 10(3),
atures above the coil-to-helix transition temperature of KC, and this 277e283.
was represented by drawing each network on a larger scale. Nono, M., Durand, D., & Nicolai, T. (2012). Rheology and structure of mixtures of
iota-carrageenan and sodium caseinate. Food Hydrocolloids, 27(1), 235e241.
Parker, A. (1995). Rheology of mixed carrageenan gels: opposing effects of potas-
4. Summary sium and iodide ions. In Food macromolecules and colloids (pp. 495e498). The
Royal Society of Chemistry.
IC addition to KC led to an initial increase in the elastic modulus Parker, A., Brigand, G., Miniou, C., Trespoey, A., & Vallee, P. (1993). Rheology and
fracture of mixed iota-carrageenan and kappa-carrageenan gels e 2-step
that passed through a peak, followed by a decrease to values close
gelation. Carbohydrate Polymers, 20(4), 253e262.
to those of the pure KC gel. At small KC additions to IC, addition of Piculell, L. (2006). Gelling carrageenans. In A. M. Stephen, G. O. Phillips, &
the elastic moduli strongly underestimated the elastic modulus. P. A. Williams (Eds.), Food polysaccharides and their applications (pp. 239e288).
These results support the presence of a bicontinuous structure at Boca Raton: CRC Press.
Piculell, L., Nilsson, S., & Muhrbeck, P. (1992). Effects of small amounts of kappa-
high i:k ratios. A clear transition was seen at i:k ratios of 2:1 to 4:3 carrageenan on the rheology of aqueous iota-carrageenan. Carbohydrate Poly-
in measurements at both small and large deformation. We mers, 18(3), 199e208.
conclude that phase separation takes place at the concentration Pouzot, M., Durand, D., & Nicolai, T. (2004). Influence of the ionic strength on the
structure of heat-set globular protein gels at pH 7. b-lactoglobulin. Macromol-
range tested, and it leads to formation of a bicontinuous structure. ecules, 37(23), 8703e8708.
The phase separation could be enhanced by allowing for curing of Ridout, M. J., Garza, S., Brownsey, G. J., & Morris, V. J. (1996). Mixed iota-kappa
IC at temperatures above the coil-to-helix transition temperature of carrageenan gels. International Journal of Biological Macromolecules, 18(1e
2), 5e8.
KC, and this coarsening of the structure led to a slight increase in Rochas, C., Domard, A., & Rinaudo, M. (1980). Aqueous GPC of electrolytes and
the elastic modulus. Further investigations are required to provide polyelectrolytes. European Polymer Journal, 16(2), 135e140.
spectroscopic evidence of the phase separated bicontinuous Rochas, C., Rinaudo, M., & Landry, S. (1989). Relation between the molecular
structure and mechanical properties of carrageenan gels. Carbohydrate Poly-
structure and the structural transitions, and also to reproduce our mers, 10(2), 115e127.
rheological findings using pure forms of KC and IC. Ross-Murphy, S. B. (1995). Small deformation rheological behaviour of biopolymer
mixtures. In S. E. Harding, S. E. Hill, & J. R. Mitchell (Eds.), Biopolymer mixtures
(pp. 247e288). Nottingham: Nottingham University Press.
Acknowledgements
Stevenson, T. T., & Furneaux, R. H. (1991). Chemical methods for the analysis of
sulphated galactans from red algae. Carbohydrate Research, 210, 277e298.
The present work was financially supported by Research and Tuvikene, R., Truus, K., Robal, M., Pehk, T., Kailas, T., Vaher, M., et al. (2009).
Development Projects for Application in Promoting New Policy of Structure and thermal stability of pyruvated carrageenans from the red alga
Coccotylus truncatus. Carbohydrate Research, 344(6), 788e794.
Agriculture, Forestry and Fisheries (No. 22026). R.T. acknowledges Usov, A. I. (2011). Polysaccharides of the red algae. In H. Derek (Ed.), Advances in
receipt of financial support from the Estonian Ministry of Educa- carbohydrate chemistry and biochemistry (Vol. 65). London: Academic Press.
tion, targeted financing No. SF0690034s09. van den Berg, L., van Vliet, T., van der Linden, E., van Boekel, M., & van de Velde, F.
(2007a). Breakdown properties and sensory perception of whey proteins/
polysaccharide mixed gels as a function of microstructure. Food Hydrocolloids,
References 21(5e6), 961e976.
van den Berg, L., van Vliet, T., van der Linden, E., van Boekel, M., & van de Velde, F.
Brenner, T., Wang, Z., Achayuthakan, P., Nakajima, T., & Nishinari, K. (2013). (2007b). Serum release: the hidden quality in fracturing composites. Food Hy-
Rheology and synergy of k-carrageenan/locust bean gum/konjac glucomannan drocolloids, 21(3), 420e432.
gels. Carbohydrate Polymers, 98(1), 754e760. van de Velde, F., Peppelman, H. A., Rollema, H. S., & Tromp, R. H. (2001). On the
Chen, J. S. (2009). Food oral processing e a review. Food Hydrocolloids, 23(1), 1e25. structure of kappa/iota-hybrid carrageenans. Carbohydrate Research, 331(3),
Davies, W. E. A. (1971). The theory of elastic composite materials. Journal of Physics 271e283.
D-Applied Physics, 4(9), 1325e1339. Yaphe, W., & Arsenault, G. P. (1965). Improved resorcinol reagent for the determi-
De Kruif, C. G. (1993). The turbidity of renneted skim milk. Journal of Colloid and nation of fructose, and of 3,6-anhydrogalactose in polysaccharides. Analytical
Interface Science, 156(1), 38e42. Biochemistry, 13(1), 143e148.
Foegeding, E. A., Daubert, C. R., Drake, M. A., Essick, G., Trulsson, M., Vinyard, C. J., Zhang, J., & Rochas, C. (1990). Interactions between agarose and k-carrageenans in
et al. (2011). A comprehensive approach to understanding textural properties of aqueous solutions. Carbohydrate Polymers, 13(3), 257e271.
semi- and soft-solid foods. Journal of Texture Studies, 42(2), 103e129.

Anda mungkin juga menyukai