You are on page 1of 11

Food Hydrocolloids 121 (2021) 107033

Contents lists available at ScienceDirect

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

Carrageenan molecule conformations and electrokinetic properties in


electrolyte solutions: Modeling and experimental measurements
Aneta Michna *, Wojciech Płaziński, Dawid Lupa, Monika Wasilewska, Zbigniew Adamczyk
Jerzy Haber Institute of Catalysis and Surface Chemistry, Polish Academy of Sciences, Niezapominajek 8, PL-30239 Krakow, Poland

A R T I C L E I N F O A B S T R A C T

Keywords: Physicochemical properties of κ–carrageenan (KC) molecules and their aqueous solutions were determined using
Carrageenan molecule conformations molecular dynamics modeling and experimental techniques. Chain conformations, the end-to-end distance, the
Electrokinetic charge of carrageenan gyration radius and the density of the molecule were theoretically calculated. Experimentally, the diffusion
Hydrodynamic diameter of carrageenan
coefficient and electrophoretic mobility of the molecules were measured for ionic strength range 10− 4 to 0.15 M
Molecular dynamics of carrageenan
Viscosity of carrageenan solutions
and different pHs. Using these data, the hydrodynamic diameter, the electrokinetic charge and the effective
ionization degree of KC molecules were determined. The dynamic viscosity measurements for dilute KC solutions
enabled to determine of its molecule intrinsic viscosity for various ionic strengths. The viscosity attained 50,000
for dilute electrolyte solutions compared to the Einstein value for spheres equal to 2.5. This confirmed that KC
molecules assumed extended conformations in accordance with theoretical modeling. It is concluded that the
application of complementary theoretical and experimental methods furnishes reliable information about KC
molecule conformations in electrolyte solutions.

1. Introduction the position and the number of sulphate groups within the disaccharide
repeat units and in the content of 3,6-anhydrogalactose residues. The
Carrageenans are non-toxic, biocompatible macroions of large vis­ forms are mainly obtained by extraction from red seaweeds (Montolalu,
cosity and excellent solubility even in cold water. Thus, they are wildly Tashiro, Matsukawa, & Ogawa, 2008); (Bono, Anisuzzaman, & Ding,
applied in drug delivery systems (Bonferoni et al., 1993); (Bonferoni 2014). However, one should notice that the red seaweed extract is a
et al., 1994); (Hariharan, Wheatley, & Price, 1997), for controlled drug mixture of various forms of carrageenans, and the composition of the
release (Li, Ni, Shao, & Mao, 2014), as potent inhibitors of viruses mixture depends on the algal source, life stage and even extraction
(Gonzalez, Alarcón, & Carrasco, 1987); (Rodríguez et al., 2014); procedure (Reis et al., 2008). For example, κ and ι form is mainly
(Nakashima et al., 1987); (Rodríguez et al., 2014), and bacterial in­ extracted from Kappaphycus alvarezii and Eucheuma denticulatum, and
fections (Briones, Sato, & Bigol, 2014) as well as as potential therapeutic λ-carrageenan - from Gigartina skottsbergi and Sarcothalia crispate (Alba
agents against SARS-CoV-2 (Pereira & Critchley, 2020). It is also re­ & Kontogiorgos, 2019). The κ and λ forms of carrageenans are especially
ported that λ-carrageenan exhibits anti-tumour and immunomodulation interesting in terms of their structure and possible applications.
properties (Harding, 2017); (Zhou et al., 2004); (Luo et al., 2015). In Oppositely to λ-carrageenan (LC), the κ-carrageenan (KC) possesses
addition to these applications, they are used in cosmetics and food only one sulphate group per building block of disaccharide. The sulphate
products as efficient thickeners, gelling agents or stabilizers (Abd Karim content is in the range of 25%–30% (Harding, 2017). It is constructed
et al., 1999); (Saha & Bhattacharya, 2010), (Harding, 2017). from alternating α(1 → 3)-D-galactose-4-sulphate and β(1 → 4)-3,
These macroions belong to a group of linear polysaccharides pos­ 6-anhydro-D-galactose (Harding, 2017); (Reis et al., 2008). KC is soluble
sessing backbones formed by α(1 → 3) and β(1 → 4)-linked galactose in hot water, forms gels in the presence of K+ due to the helical form
residues with repeating sulphate half-ester groups and 3,6-anhydro- (Reis et al., 2008), which is additionally stabilized by hydrogen bonds
bridges (Imeson, 2009). There are six main forms of carrageenans: λ (Reis et al., 2008). It also belongs to a group of “reversibly
(lambda), κ (kappa), ι (iota), μ (mu), θ (theta) and ν (nu), which differ in soluble-insoluble polymers” because of precipitating with 0.2% KCl at

* Corresponding author.
E-mail addresses: aneta.michna@ikifp.edu.pl (A. Michna), wojciech.plazinski@ikifp.edu.pl, wojtek_plazinski@o2.pl (W. Płaziński), dawid.lupa@ikifp.edu.pl
(D. Lupa), monika.wasilewska@ikifp.edu.pl (M. Wasilewska), zbigniew.adamczyk@ikifp.edu.pl (Z. Adamczyk).

https://doi.org/10.1016/j.foodhyd.2021.107033
Received 3 March 2021; Received in revised form 17 June 2021; Accepted 11 July 2021
Available online 13 July 2021
0268-005X/© 2021 Published by Elsevier Ltd.
A. Michna et al. Food Hydrocolloids 121 (2021) 107033

38 ◦ C and dissolving the precipitate in distilled water (Reis et al., 2008); Analytical grade (pure p. a.) NaCl was purchased from Avantor
(Roy & Gupta, 2003). The molar mass of κ-carrageenan, determined by Performance Materials Poland S.A.
analytical centrifugation and light scattering, is of the order of 300 kg Eight generation of poly (amidoamine) dendrimers, PAMAM8,
mol− 1 (Slootmaekers, van Dijk, Varkevisser, van Treslong, & Reynaers, (8,33% w/w aqueous solution, dispersity index below 1.04 (Steche­
1991); (Slootmaekers, Mandel, & Reynaers, 1991); (Harding, 2017). KC messer & Eimer, 1997)), of the molar mass of 233 kg mol− 1, was pur­
is applied in the whole-cell and enzyme immobilization, for wastewater chased from Dendritech, Inc. (the USA). The PAMAM8 solutions were
treatment and beer and in ethanol production (Reis et al., 2008). applied for the formation of the anchoring layer, which was used to
Because of its significance, the properties of carrageenan solutions determine the carrageenan adsorption kinetics in OWLS experiments.
have been studied with the aim to evaluate their molar mass distribution The obtained results were described in Supplementary Material.
(Slootmaekers, van Dijk et al., 1991); (Thành et al., 2002); (Almutairi The solutions of carrageenan were prepared by dissolving a proper
et al., 2013); (Schuck et al., 2014), radii of gyration, contour lengths amount of the powder in NaCl solutions of a precisely controlled con­
(Slootmaekers, Mandel, & Reynaers, 1991), persistence lengths (Berth, centration and pH. Deionized water was obtained using Milli-Q Elix &
Vukovic, & Lechner, 2008), hydrodynamic diameters and second virial Simplicity 185 purification system from Millipore SAS Molsheim,
coefficients (Thành et al., 2002). France.
Several works have been also focused on the determination of the The temperature of all experiments was fixed of 298 K.
intrinsic viscosity of carrageenan solutions in various electrolytes
comprising multivalent ions (Zabik & Aldrich, 1965), (Zabik & Aldrich,
1967) for large ionic strength (Chronakis, Doublier, & Piculell, 2000). 2.2. Methods
Interesting results were obtained by Berth et al. who analysed LC and
KC solutions by static light scattering (multi-angle light scattering, The chemical composition of the sample, especially the sulfur con­
MALS) in 0.1 M NaNO3, which allowed to determine the molar mass of tent was determined by elemental analysis (Thermo Scientific FlashS­
the carrageenans (1400 kg mol− 1 for LC and 596 kg mol− 1 for KC), their mart Elemental Analyzer) and Fourier Transform Infrared Spectroscopy
radius of gyration (102 nm for LC and 104 nm for KC) and the second (FTIR) (FTIR Nicolet 6700 spectrometer, Thermo Scientific), whereas its
virial coefficient (10− 4 mol ml g− 2 for LC and 9.9 × 10− 4 mol ml g for molar mass was acquired using size exclusion chromatography coupled
KC) (Berth et al., 2008). The obtained data were interpreted in terms of to multi-angle light scattering (SEC-MALS) analysis. SEC-MALS analysis
the wormlike chain model using the Skolnik-Odijk-Fixman approach. was performed using high-performance liquid chromatography (HPLC)
The intrinsic persistence length of 3–4 nm and expansion factor of 1.5 instrument (1260 Infinity LC, Agilent Technologies) equipped with a UV
were also calculated. detector, MALS detector (DAWN HELEOS II, Wyatt Technology) and a
Physicochemical characteristics of LC involving the molar mass, differential refractometer (Optilab T-rEX, Wyatt Technology).
intrinsic viscosity and sedimentation coefficient (at pH 7.0 and ionic The electrophoretic mobility and diffusion coefficients of carra­
strength of 0.1 M) were acquired by Almutairi et al. (Almutairi et al., geenan molecules were measured for the broad ionic strength range
2013) using size exclusion chromatography coupled to MALS, capillary (10− 4 to 0.15 M at pH 5.8), using a combination of the micro-
viscometry, and analytical ultracentrifugation. An extended and flexible electrophoresis with Laser Doppler Velocimetry (LDV) and the dy­
conformation for LC molecules is confirmed by these investigations. namic light scattering (DLS) (Malvern Zetasizer Nano ZS apparatus). The
However, despite of extensive experimental effort, no comprehen­ Henry (Ohshima, 2012) and Stokes-Einstein (Einstein, 1908) equations
sive characteristics of KC molecules were acquired simultaneously were applied for determining the KC zeta potentials and hydrodynamic
applying theoretical modeling and experimental techniques, which can diameters for the above conditions. The carrageenan bulk concentra­
yield information about the electrokinetic properties and the molecule tions were equal to 300 and 500 mg L− 1 in DLS and 500 mg L− 1 in LDV
conformations under various ionic strengths and pHs. measurements.
Therefore, the goal of this work is to perform thorough characteris­ Mass loss measurements were performed by thermogravimetric
tics of KC solutions using the all-atom molecular dynamics (MD) analysis (TGA) and differential scanning calorimetry (DSC) measure­
modeling and complementary experimental methods such as the dy­ ments using Netzsch STA 409 PC microcalorimeter. Additionally, the
namic light scattering (DLS), micro-electrophoresis (LDV), the optical dry mass of the carrageenan powder was determined using static ther­
waveguide lightmode spectroscopy (OWLS), and the dynamic viscosity mogravimetry, where KC mass changes on time were monitored at
measurements. The previously unexplored range of dilute supporting constant temperature (378 K). The detailed protocol for static ther­
electrolyte (NaCl) concentration is studied in order to acquire reliable mogravimetry analysis and obtained TGA/DSC curves can be found in
information about the extended chain length. These investigations yield Supplementary Material.
quantitative information about the KC molecule properties, such as its The density of carrageenan solutions was determined using the high
effective charge, ionization degree, chain length and its cross-section precision Anton Paar DMA 5000 M densitometer.
area for a broad ionic strength range. This is significant for predicting The dynamic viscosity of the carrageenan solutions of defined mass
and controlling the adsorption of KC molecules at solid/electrolyte in­ concentrations was determined using the Anton Paar rolling-ball
terfaces, which is often applied to produce multilayer shells for targeted viscometer Lovis 2000 M/ME equipped with a short capillary tube.
drug delivery systems. Our results also facilitate a proper interpretation This allowed determining the viscosity of samples with a large precision
of experimental data derived from the quartz crystal microbalance or the (0.05%) using relatively small volumes of carrageenan solutions (0.1
streaming potential measurements and other experimental techniques mL).
that require calibration. Additionally, the flexibility of adjusting the Three sensors, situated in the viscometer capillary, measure a run­
viscosity of KC samples by ionic strength change has practical implica­ time of a steel ball in the capillary filled by the sample (pure electrolyte
tions for the regulation of the thickening ability of carrageenan or carrageenan solutions of various concentrations). From the measured
solutions. runtimes, the viscometer software yields the dynamic viscosity of the
solution using the following equation:
2. Materials and methods η = K(ρb − ρs )Δt (1)

2.1. Materials where: η is dynamic viscosity [mPa s], K [m2 s− 2] is the calibration
constant, ρb is the ball density [kg m− 3], ρs is sample density [kg m− 3],
Carrageenan sodium salt was supplied by Sigma Aldrich (Poland) in and Δt is the runtime [s].
the form of crystalline powder. The measurements were carried out for the KC volume fraction not

2
A. Michna et al. Food Hydrocolloids 121 (2021) 107033

exceeding 5 × 10− 4 (dilute macroion concentration limit) and at a fixed After equilibration, production simulations were carried out for a
NaCl concentration varied between 10− 4 and 0.15 M. duration of 200 ns and the data were saved to trajectory every 2 ps. The
The kinetics of carrageenan adsorption from dilute solutions on end-to-end distance and gyration radius values were calculated by using
supporting PAMAM8 layers was determined using the OWLS method the GROMACS routines gmx polystat and gmx mindist.
according to the procedure described elsewhere (Wasilewska, Adamc­
zyk, Pomorska, Nattich-Rak, & Sadowska, 2019); (Michna, Pomorska, 3. Results and discussion
Nattich-Rak, Wasilewska, & Adamczyk, 2020). The OWLS 210 instru­
ment (Microvacuum Ltd., Budapest, Hungary) was used. The apparatus 3.1. Theoretical modeling results
is equipped with a laminar slit shear flow cell comprising a silica-coated
waveguide (OW2400, Microvacuum). The adsorbing substrates were The calculation presented hereafter were performed for κ-carra­
planar optical waveguides made of a glass substrate (refractive index geenan (KC) molecules composed of 2, 5, 10 and 20 monomers, each
1.526) covered by a film of Si0.78Ti0.22O2 of the thickness 170 nm, and comprising one –OSO-3 group (see Fig. 1). The molar mass of fully
the refractive index 1.8. A grating embossed in the substrate enables the ionized monomer calculated from its chemical composition is equal to
light to be coupled into the waveguide layer. The sensor surface was 0.385 kg mol− 1, whereas the molar mass of sodium ion compensated
coated with an additional layer (10 nm) of pure SiO2 (Wasilewska et al., monomer is equal to 0.408 kg mol− 1. Let us note that in the accepted
2019). nomenclature ‘monomer’ is equivalent to the two consecutive residues
in the KC chain (i.e. disaccharide building block).
2.3. Theoretical modeling Primarily, in the MD simulations, the molecule conformation, the
end-to-end distance and the gyration radius were determined as a
A series of carrageenan chains of various lengths, composed of 4, 10, function of the monomer number, denoted by Nm. Exemplary snapshots
20 and 40 monosaccharide residues (equivalent to 2, 5, 10 and 20 of KC chain conformations obtained for NaCl concentration of 10− 2 M
‘monomers’, in accordance to the notation accepted in this work) was and different monomer numbers are shown in Fig. 2. Qualitatively, one
considered in the theoretical modeling. The initial configurations of the can observe that the chains are quasi-rigid, exhibiting a rod-like shape
systems, including the solvation of the KC chains as well as the addition with no sharp bending. The most frequently occurring pattern of solute-
of ions were created using the CHARMM-GUI online server (Park et al., solute hydrogen bonding included interactions between the –SO-3 and
2019); (Lee et al., 2016). The modeling system consisted of cubic boxes –OH groups of the two consecutive monosaccharide residues. Apart
of the initial edge dimensions varying from 4.2 to 14.7 nm, depending from the conformationally restricted mutual orientation of the neigh­
on the system. The corresponding number of water molecules varied bouring residues, no tendency to the formation of regular, helical shapes
from 2300 to 110000, respectively. The appropriate number of Na+ and in a larger dimensional scale was observed.
Cl− ions was added to each system, accounting for its neutral charge and The semi-rigid conformation of the KC chain was confirmed by the
the desired ionic strength value (10− 2 M). additional, enhanced sampling simulations and the resulting free energy
The all-atom molecular dynamics (MD) simulations were carried out maps (FEMs) calculated with respect to the glycosidic dihedral angle
using the GROMACS 2016.4 package (Abraham et al., 2015). The values (see details in Supplementary Material). Inspecting FEMs (Sup­
CHARMM36 force field (Guvench et al., 2011); (Guvench et al., 2008) plementary Material), one can observe that the FEM area corresponding
was used to describe the interactions involving KC molecules, accom­ to the free energy levels < 5 kJ/mol covers a relatively significant
panied by the CHARMM-compatible explicit TIP3P water model. The fraction of the map calculated for the β(1 → 4) type of linkage, on the
parameters describing the anhydro residues were prepared manually contrary to that calculated for the α(1 → 3) linkage. This indicates that
and relied on the parameters for unfunctionalized galactose units. The the semi-rigid conformation of the KC chain is distorted mainly by the
introduced changes include: (i) defining new covalent bond between O6 rearrangments within the β(1 → 4) linkage. Furthermore, the alternative
and C3; (ii) conversion of the O3 and H3 atoms into ‘dummy’ atoms, minima of the free energy, not associated with the shapes sampled
non-interacting with any other part of the system by non-bonded in­ during standard MD simulations, are located at approx. 12 kJ/mol (β(1
teractions; (iii) adjusting the partial charges on C6 and O6 to maintain → 4) linkage) above the most favorable conformation or essentially
the nautral charge of the residue; (iv) changing the atom types accord­ absent (α(1 → 3) linkage). Thus, the population of the corresponding
ingly to their altered chemical character in the case of O6, C6 and C3. chain geometries is negligible (<1%).
The modeling was carried out applying periodic boundary conditions Independent MD runs confirmed that the monomer length was
and in the isothermal-isobaric ensemble. The temperature was main­ practically independent of the chain length, adopting an average value
tained close to its reference value (298 K) by applying the V-rescale of 0.90 ± 0.03 nm (denoted later as lm). Interestingly, this agrees within
thermostat (Bussi, Donadio, & Parrinello, 2007), whereas for the con­ error bounds with the experimental value reported by (Slootmaekers,
stant pressure (1 bar, isotropic coordinate scaling) the Mandel, & Reynaers, 1991) for λ-carrageenan, which was equal to 0.89
Parrinello-Rahman barostat (Parrinello & Rahman, 1981) was used with nm.
a relaxation time of 0.4 ps. The equations of motion were integrated with The MD modeling also allowed to determine the average end-to-end
a time step of 2 fs using the leap-frog scheme (Hockney, 1970). The distance of the molecule Le and the average gyration radius Rg as a
hydrogen-containing solute bond lengths were constrained by the function of the number of monomers Nm. These dependencies are
application of the LINCS procedure with a relative geometric tolerance illustrated in Fig. 3. As seen, both the end-to-end distance and the gy­
of 10− 4 (Hess, Bekker, Berendsen, & Fraaije, 1997); (Hess, 2008). The ration radius were well approximated by linear functions with the slope
full rigidity of the water molecules was enforced by application of the slightly smaller than predicted for the fully extended chain where one
SETTLE procedure (Miyamoto & Kollman, 1992). The electrostatic in­ has:
teractions were modeled by using the particle-mesh Ewald method
Le = lm Nm (2)
(Darden, York, & Pedersen, 1993) with cut-off set to 1.2 nm, while van
der Waals interactions (LJ potentials) were switched off between 1.0 and Also, as seen in Fig. 3, the values of the gyration radius derived from
1.2 nm. The translational center-of-mass motion was removed every MD slightly deviate from the values of gyration radius predicted for a
timestep separately for the solute and the solvent. straight cylinder of the length equal to Le, i.e., Rg = lm Nm /121/2 . This
The systems were subjected to geometry minimization and MD-based confirms that the carrageenan molecules assume extended rod-like
equilibrations in the NPT ensemble, lasting 2–20 ns, depending on the conformations for Nm up to 20.
system size. After this step we checked that the geometry of the rings in The MD-originating descriptors were applied to determine the KC
the anhydro residues was in accordance to the expectations, i.e. 1C4.

3
A. Michna et al. Food Hydrocolloids 121 (2021) 107033

Fig. 1. A schematic representation of the chemical structure of the KC molecule with the marked monomer unit, corresponding to the two consecutive residues of
3,6-anhydrogalactose (3,6-AnGal) and galactopyranose-4-sulphate (Gal4S).

2
Fig. 2. Snapshots of KC chain conformations for 20 and 40 residues derived from MD modeling for 10− M NaCl concentration (equivalent to 10 and 20 monomers,
respectively). Hydrogen atoms and solvent molecules are omitted for clarity.

molecule density using the method developed in previous work molar mass of KC furnishes useful data that are inaccessible for direct
(Adamczyk, Morga, Kosior, & Batys, 2018). Accordingly, the size of the theoretical modeling because of excessive time of computations. In
simulation boxes, where a single KC molecule was confined, was sys­ order to assess the validity of this extrapolation, throughout experi­
tematically increased that resulted in a decrease in the KC mass fraction mental studies were performed to determine the diffusion coefficient,
from 0.01 to 0. The density of these systems ρs, as well as that of the pure the hydrodynamic diameter, the electrokinetic charge, the ionization
solvent ρe, were determined in the independent MD runs. Then, the degree, and intrinsic viscosity of KC molecules.
dependence of ρe/ρs on wp was plotted and fitted by a straight line
characterized by the slope sp and the density was calculated from the
formula: 3.2. Experimental characteristics of carrageenan solutions
ρe
ρp = (3) To evaluate the composition of carrageenan sample, a comprehen­
1 + sp
sive physicochemical analysis was performed using the FTIR and the
In this way, one obtains, at the temperature of 298 K, three different elemental analysis methods. The details of performed experiments and
ρp values, dependent on the assumption inherent in the calculation their results were discussed in Supplementary Material (sections: 1.2
scheme: (i) ρp = 2.24 × 103 kg m− 3 for bare, negatively charged chain and 1.3). Using these methods it is shown that the molar fraction of KC in
without either hydration or co-cations contributing to the KC mass; (ii) sample is equal to 0.944, whereas the second present component is LC.
ρp = 2.10 × 103 kg m− 3
when considering the sodium co-cations The number average molar mass of KC was equal to 475 ± 0.08 kg
contributing to the mass of the KC chain but no hydration; (iii) ρp = mol− 1, as found using SEC-MALS. This value is close to the mass average
2.01 × 103 kg m− 3 in the case where sodium co-cations contribute to the molar mass of KC, which was determined to be equal to 480 ± 0.08 kg
mass of the KC chain and, additionally, the chain hydration is assumed mol− 1. Therefore, one can conclude that KC sample exhibits low dis­
(1 water molecule per monomer). Using these values one can calculate persity (Mw/Mn = 1.01).
the volume of a monomer considering its molar mass M1 from the The dry mass of the carrageenan powder used for preparing its so­
dependence v1 = M1 /(ρp Av), which was equal to 0.29, 0.30 and 0.32 lutions used in the experiments was determined via the TGA and static
nm3 for the cases: (i), (ii) and (iii), respectively. Consequently, the thermogravimetry experiments described in Supplementary Informa­
equivalent monomer diameter assuming its cylindrical shape calculated tion. The static thermogravimetry revealed that the water content in the
as d1 = (4v1 /π lm )1/2 was equal to 0.64, 0.66 and 0.67 nm, respectively. carrageenan sample was equal to 10.5%, whereas the TG/DTA/DSC
For the sake of convenience, the theoretical data obtained from measurements showed 13.5%. For further calculations, a mean value of
modeling are collected in Table 1. water content equal to 12% was used. The determined water content is
It should be mentioned the extrapolation of these results to a larger comparable with the value of 14% reported by Dul et al. (Dul et al.,
2015).

4
A. Michna et al. Food Hydrocolloids 121 (2021) 107033

Fig. 3. Part (a): The average KC molecule end-to-


end distance Le vs. the number of monomers
derived from MD modeling. The solid line denotes
the linear fitting of theoretical data. The dashed line
shows the results predicted for the fully extended
chain, i.e., Le = Nm lm, Part (b): The gyration radius
of KC vs. the number of monomers. The solid line
denotes linear fitting of theoretical data derived
from MD simulations and the dashed line shows the
results predicted for the fully extended chain of a
cylindrical shape, i.e., Rg = Nmlm/121/2. Vertical
bars denote the fluctuations of the given quantity
found during MD simulations and expressed as
standard deviation values. The NaCl concentration
was equal to 10− 2 M.

Primarily, in the experiments, the diffusion coefficient of KC mole­ mg L− 1 were extrapolated to the infinite dilution using the correction
cules for various ionic strengths was determined by DLS as described function proposed by Slootmaekers et al. (Slootmaekers, Mandel, &
above. It was equal to (4.9 ± 0.4) × 10− 12 and (6.7 ± 0.3) × 10− 12 m2 Reynaers, 1991).
s− 1 at the ionic strength of 10− 4 and 0.15 M, respectively (see Table 2). It
should be mentioned, however, that the precision of these measure­ D0 = D(1 + kD cb )− 1
(4)
ments decreases for ionic strength below 10− 3 M due to volume exclu­
where: D is the diffusion coefficient determined by DLS, D0 is the cor­
sion effects enhanced by the long-range electrostatic interactions among
rected diffusion coefficient and kD is the constant equal to 4.3 × 10− 4 L
charged KC molecules. In order to account for this effect, the primary
mg− 1.
experimental data obtained for the KC bulk concentration of 300–500
The corrected values are listed in Table 2. Interestingly, for 0.15 M

5
A. Michna et al. Food Hydrocolloids 121 (2021) 107033

Table 1 coefficient D data one can determine the electrokinetic charge at the KC
Primary physicochemical characteristics of the KC molecule. molecule applying the Lorentz–Stokes relationship (Adamczyk, Bratek,
Quantity [unit], symbol Value Remarks Jachimska, Jasiński, & Warszyński, 2006); (Michna, Adamczyk, Sofiń­
ska, & Matusik, 2017):
Monomer molar mass [kg 0.385 fully ionized chain
mol − 1], M1 0.408 sodium salt kT
Monomer contour length 0.90 ± this work, MD modeling qe = μ = 3πηdH μe (5)
D e
[nm], lm 0.03
Molecule density [kg m− 3], 2.24 ± 0.2 this work, MD modeling, Consequently, the number of elementary charges Nc per one KC
ρp × 103 no hydration molecule can be calculated as:
2.10 ± 0.2 this work, MD modeling,
/
× 103 contributing mass of Na+ kT
2.01 ± 0.2 this work, MD modeling, contributing Nc = qe e = 6.25 × 1010 μe (6)
D
× 103 mass of Na+, with hydration
Monomer volume [nm3], ν1 0.286 no hydration, calculated as v1 = M1 /
(ρp Av) where e is the elementary charge (equal unity), D is expressed in m2 s− 1,
0.305 contributing mass of Na + kT is expressed in J (kg m2 s− 2) and μe is expressed in μm cm (V s)− 1.
0.318 contributing mass of Na+, with Eq. (6) is valid for an arbitrary charge distribution and the shape of
hydration molecules. However, its accuracy decreases for larger ionic strengths
Monomer equivalent 0.64 ± no hydration, calculated as where the double-layer thickness κ− 1 = (εkT/2e2I)1/2 (where ε is the
cylinder diameter [nm], 0.04 (4v1 /πlm )1/2 electric permittivity of the solvent, and I is the ionic strength) becomes
dc
0.66 ± contributing mass of Na+ comparable with the molecule diameter.
0.03
It should be emphasized that the electrokinetic charge which is
0.67 ± contributing mass of Na+, with
0.03 hydration located below the flow shear plane (Hunter, 1981), is an important
parameter controlling macromolecule interactions among themselves, i.
* hydration for KC: 0.5 H2O per residue.
e., their solution stability and their interactions with interfaces, i.e.,
their adsorption capacity.
NaCl the corrected value of the diffusion coefficient is equal to (5.6 ± Using the experimental diffusion coefficient and the electrophoretic
0.3) × 10− 12 m2 s− 1, which matches the value obtained by Slootmaekers mobility data one can calculate from Eq. (6) that KC molecule charge
et al. (Slootmaekers, Mandel, & Reynaers, 1991), albeit for the LC
sample of an average molar mass of 607 kg mol− 1 and ionic strength of
0.1 M.
As previously discussed (Adamczyk et al., 2018) the increase of the
diffusion coefficient with the supporting electrolyte concentration is
caused by the change in the macromolecule conformation, which as­
sumes a less extended form.
In order to determine the effective (compensated) charge of KC
molecules in NaCl solutions, the electrokinetic method exploiting the
LDV technique (Ohshima, 2012) was applied. This method yields the
absolute values of the ratio of the molecule migration velocity to the
applied electric field, defined as the electrophoretic mobility and
denoted byμe . The dependence of the electrophoretic mobility on the
NaCl concentration is shown in Fig. 4 and these data are also collected in
Table 2. As can be noticed, the mobility increases from − 6.3 to − 2.9 μm
cm (V s)− 1 for the NaCl concentration (ionic strength) equal to 10− 4 and
0.15 M, respectively (pH 5.8). These electrophoretic mobility values
correspond to the zeta potential of the molecule (calculated from the
Debye-Hückel-Henry model) equal to − 120 and − 56 mV, respectively.
Negative values of the electrophoretic mobility and the zeta potential
confirm that the electrokinetic charge of the KC molecule is negative for Fig. 4. The dependence of the electrophoretic mobility of the KC molecule
this range of NaCl concentrations. (bulk concentration of 500 mg L− 1) on the NaCl concentration, pH 5.8, the solid
Using the experimental electrophoretic mobilityμe and the diffusion line is the guide for the eyes.

Table 2
Experimental characteristics of KC solutions at pH 5.8 in NaCl electrolyte of various concentrations [M], T = 298 K.
1
cNaCl [M] κ− [nm] D [m2 s− 1] D0* [m2 s− 1] dH [nm] dH0 [nm] dH0*** [nm] μe [μm cm (Vs)− 1] ζ [mV] Nc [1] α** [1]
− 12 − 12
10–4
30.5 4.9 × 10 4.0 × 10 100 ± 20 120 ± 20 _ − 6.3 ± 0.3 − 120 ± 10 − 400 ± 80 0.34
12 12
10–3 9.63 5.6 × 10− 4.6 × 10− 88 ± 20 106 ± 20 110 ± 20 − 5.2 ± 0.3 − 100 ± 12 − 290 ± 60 0.25
12 12
10–2 3.05 5.8 × 10− 4.8 × 10− 85 ± 20 102 ± 20 110 ± 20 − 4.6 ± 0.3 − 89 ± 12 − 250 ± 70 0.21
12 12
0.05 1.36 5.9 × 10− 4.9 × 10− 83 ± 20 100 ± 20 _ − 4.4 ± 0.1 − 89 ± 6 − 230 ± 70 0.20
12 12
0.15 0.786 6.7 × 10− 5.6 × 10− 73 ± 20 88 ± 20 80 ± 10 − 2.9 ± 0.2 − 56 ± 6 − 130 ± 30 0.11
− 1
D0 = D(1 + kD cb ) ;
− 4 − 1
* kD = 4.3 × 10 L mg ;
dH = kT/3πηD;
dH0 = kT/3πηD0 ; T = 298K
kT kT
α = Nc /Nm ; Nc = μ = 6.25 × 1010 μe , Nm = Mp /M1
** D0 e e D0

***determined from OWLS measuremetns.

6
A. Michna et al. Food Hydrocolloids 121 (2021) 107033

varies between − 400 and − 130 for NaCl concentration varying between ( )− 1
10− 4 and 0.15 M, respectively, see Table 2. Given that the number of dH0 = dH 1 + CH σ 2M (9)
monomers in the KC molecule Nm = Mp/M1 is equal to 1160, its nominal
charge stemming from the dissociation of ionogenic groups should be where:dH0 is the hydrodynamic diameter for a monodisperse system, dH
equal to − 1160 e. This comparison indicates that the electrokinetic is the experimental value of the hydrodynamic diameter for a disperse
charge is only a small fraction of the nominal molecule charge, i.e., the system exhibiting the same average molar mass, σ M is the relative
effective dissociation degree is equal to 34 and 11% for NaCl concen­ standard deviation of the size distribution and CH is the dimensionless
tration equal to 10− 4 and 0.15 M, respectively. Physically, such a sig­ constant given by
nificant charge compensation is the result of the counterion
accumulation in the diffuse part of the electric double-layer caused by 0.5 c21
CH = − c 1 + (10)
the large electrostatic field generated by the presence of the surface c1 ln 2 λ + c2 (c1 ln 2 λ + c2 )2
charges. In the macroion oriented literature, it is referred to as the
For our case taking λ = 1460, c1 = 1, and c2 = − 0.11 (this corre­
Manning ion condensation (Manning, 1979). It is also interesting to
sponds to cylindrical shape of the molecule) one obtains CH = − 0.05.
mention that such behaviour was previously reported for PDADMAC
Thus, even for a disperse system where σM = 0.5, the correction is less
(Adamczyk, Jamroży, Batys, & Michna, 2014), and for PLL (Adamczyk
than 1%, which is practically negligible compared to other experimental
et al., 2018) macroions.
errors.
In the further analysis of the KC molecule conformations is useful to
However, given the low sensitivity of the hydrodynamic diameter to
define the hydrodynamic diameter, which in contrast to diffusion co­
the aspect ratio parameter predicted by Eq. (8) a more reliable infor­
efficient, is independent of the temperature and the solvent viscosity. It
mation about the KC conformation can be derived from the intrinsic
can be directly calculated using the Stokes-Einstein relationship (Ein­
viscosity data obtained via the dynamic viscosity measurements.
stein, 1908):

dH =
kT
(7) 3.3. Dynamic viscosity measurements
3πηD
A proper interpretation of the viscosity measurements in terms of the
where η is the dynamic viscosity of the solvent. volume fraction requires the density of the KC molecule to be known.
For the NaCl concentration of 10− 4 and 0.15 M one obtains from Eq. This parameter was experimentally determined by the solution dilution
(7) the value of dH (calculated using the corrected diffusion coefficients method as described in the Supplementary Material. At the temperature
listed in Table 2) equals to 120 ± 20 and 88 ± 20 nm, respectively (see of 298 K, the density was equal to ρp = (2.03 ± 0.13) × 103 kg m− 3 for
Table 2). One can notice that the dH values obtained from the OWLS the NaCl concentration of 10− 4 - 10− 2 M that agrees with error bounds
measurements for the NaCl concentration range of 10− 3 to 0.15 M agree with previously reported by Slootmaekers et al. (Slootmaekers, Mandel,
with the data derived from DLS. The method for calculating dH using the & Reynaers, 1991) and the above theoretical value of (2.01 ± 0.2) × 103
OWLS measurements is described in the Supplementary Material. kg m− 3.
It is interesting to compare hydrodynamic diameter derived from Initially, in these measurements, the dependencies of the dynamic
experiments with the theoretical model pertinent to slender body hy­ viscosity of dilute KC solutions (characterized by bulk concentration cb
drodynamics characterized by the length to width (aspect ratio) up to 1000 mg L− 1) was measured using the above-described method.
parameter, denoted by λ, much larger than unity (Brenner, 1974). For The zero shear viscosity of the solution denoted by ηs was calculated by
such a case the hydrodynamic diameter can be expressed in the an efficient extrapolation procedure. These primary experimental results
following form (Mansfield & Douglas, 2008); (Adamczyk et al., 2012): were expressed as the dependence of the normalized viscosity ηs/ηe
Le λ (where ηe is the supporting electrolyte viscosity) on the KC volume
d H0 = = dc = f1 (λ) (8) fraction Φv = cb/ρp for various NaCl concentrations in the range 10− 4 to
(c1 ln 2 λ + c2 ) c1 ln 2 λ + c2
0.15 M. The isotonic electrolyte solutions required for the lower NaCl
where Le can be treated as the contour length, and c1, c2 are the concentration were prepared considering the above-estimated KC
dimensionless constants depending on the shape of the body. molecule ionization degree using the following formula:
It should be mentioned that Eq. (8) is valid for monodisperse mac­ /
1
roion solutions. I * = × 10− 3 Nc cb Mp (11)
2
For prolate spheroids one has c1 = 1, c2 = 0; for blunt cylinders: c1 =
1, c2 = − 0.11 (Brenner, 1974); for linear chain of touching beads c1 = 1, where I* is the excess ionic strength due to the KC molecule ionization
c2 = − 0.25, for the chain of beads forming a semi-circles: c1 = 0.95, c2 = and cb is expressed in mg L− 1.
0.02 (Adamczyk et al., 2006). One can calculate from Eq. (11) that for cb = 100 mg L− 1 and Nc
In our case Le is equal to 1040 nm, which is calculated considering equal to − 400, I* = 4 × 10− 5 M.
that the number of monomers is equal to 1160 and that the length of the The dependencies of the relative viscosity, ηs/ηe on the volume
monomer derived from MD modeling is equal to 0.90 nm. This value fraction Φv of KC molecules acquired for various NaCl concentrations are
corresponds to the limiting length of the KC molecule in a fully expanded presented in Fig. 5. The slopes of these dependencies in the limit of low
state. Additionally, considering the chain diameters given in Table 1 one volume fractions give directly the intrinsic viscosity [η], which varied
obtains λ = 1460, for dc = 0.71. In consequence, one can calculate from between (2.2 ± 0.2) × 103 and (5.0 ± 0.3) × 104 for the NaCl concen­
Eq. (8) that the hydrodynamic diameter is equal to 130 nm, for the cy­ tration of 0.15 and 10− 4 M, respectively (at pH = 5.6), see Table 3. It
lindrical molecule shape. Interestingly, for the linear touching bead should be mentioned that the intrinsic viscosity for the PBS buffer so­
shape, one obtains almost identical values of 133 nm. As can be seen, lutions (ionic strength 0.15 M, pH 7.4) is equal to that determined for
these values agree within experimental error bounds with experimental pure 0.15 M NaCl solutions.
hydrodynamic diameter in the limit of low ionic strength, which was It is also interesting to mention that the value of the intrinsic vis­
equal to 120 ± 20 nm (see Table 2). cosity for 0.15 M NaCl obtained in this work is comparable with that
It is interesting to estimate the influence of the dispersity of the reported by (Almutairi et al., 2013) who determined 1080 mL g− 1 for
molecule molar mass distribution on the precision of the hydrodynamic λ-carrageenan of the average molar mass of 870 kg mol− 1 and NaCl
diameter determination. As shown in (Adamczyk et al., 2018) the concentration of 0.15 M. This corresponds to the dimensionless value of
correction to the hydrodynamic diameter is given by: the intrinsic viscosity used in our work [η] = 2160.

7
A. Michna et al. Food Hydrocolloids 121 (2021) 107033

Fig. 5. a) The dependence of the relative viscosity ηs/ηe on the volume fraction of KC Φv, determined for various concentrations of NaCl. Hollow squares show the
results obtained for 0.15 M solution of the PBS buffer (pH 7.4). The solid line denotes a linear fit of experimental data. b) Dependence of ηi/Φv (where ηi = ηs/ηe) on
the volume fraction of KC in its solutions, Φv. Dashed lines denote the linear fit of experimental data.

Given that the obtained intrinsic viscosity values exceed by orders of where c1v = 3/15, c2v = 1/15 and cv is equal to 14/15 for blunt cylinders.
magnitude the Einstein results for spherically shaped macromolecules Eq. (12) is valid for λ ≫ 1 and for monodisperse macroion solutions.
where [η] = 2.5, one can assume that KC molecules assume extended As shown in (Adamczyk et al., 2018), using Eq. (12) the following for­
conformation for the entire NaCl concentration range, which corre­ mula can be derived for converting the experimental intrinsic viscosity
sponds to the slender body regime. Therefore, the experimental data can data obtained for disperse system to the corresponding values pertinent
be quantitatively interpreted exploiting the theoretical results pertinent to monodisperse system:
to the zero shear rate intrinsic viscosity for slender bodies (Brenner,
( )− 1
1974): [η]0 = [η] 1 + Cv σ 2M (13)
λ2 λ2
[η]* = c1 v + c2v + cv (12)
ln 2 λ − 0.5 ln 2 λ − 1.5 where [η]0 is the corrected intrinsic viscosity for the monodisperse sys­
tem and the dimensionless Cv constant is given by

8
A. Michna et al. Food Hydrocolloids 121 (2021) 107033

Table 3 pertinent to the extended chain derived from MD modeling. For ionic
The intrinsic viscosity [η], the λ parameter, the equivalent cylinder chain strength 0.15 M, the predicted chain length of KC molecule is equal to
diameter dc, the cylinder lengths Lc and the cross-section area in the side-on 520 nm, which confirms its largely elongated shape. One can argue that
orientation Sc of the KC molecules for various ionic strengths derived from these data have significance for interpreting KC molecule adsorption,
DLS and viscometry. especially for producing macroion films in the layer-by-layer processes.
cNaCl [η][1] λ [1] dc [nm] Lc [nm] Sc [nm2]
[M]
4. Conclusions
0 7.9 × 104* 1460* 0.71 1040* 740
10–4 5.0 ± 0.4 × 1140 ± 0.81 ± 920 ± 750 ± 150
Physicochemical properties of κ–carrageenan (KC) were thoroughly
104 50 0.15 200
10–3 2.4 ± 0.2 × 760 ± 40 1.0 ± 0.2 760 ± 760 ± 150 determined by molecular dynamics (MD) modeling, the SEC-MALS,
104 150 FTIR, DLS, OWLS, LDV techniques and the dynamic viscosity measure­
10–2 1.0 ± 0.1 × 520 ± 30 1.3 ± 0.25 690 ± 970 ± 200 ments. The modeling confirmed that the KC molecule exhibits a flexible-
104 150 rod shape with no sharp bending. There appear intramolecular hydrogen
0.15 2.2 ± 0.2 × 210 ± 10 2.5 ± 0.4 520 ± 70 1300 ±
bonding between –SO3 and –OH groups within the monomer. No ten­
103 250
dency to the formation of a large-scale helical conformation was pre­
*Calculated from the MD modeling using the monomer length of 0.90 nm. dicted. Additionally, several parameters of primary significance were
determined such as the monomer length equal to 0.90 nm and the chain
3 β21 + 3β22 β31 + 3β32 diameter equal to 0.71 nm. This enabled to calculate the maximum
Cv = 1 − + (14)
2 β1 + 3β2 β1 + 3β2 extended (contour) length of a KC molecule of arbitrary molar mass.
From experimental measurements, the diffusion coefficient, the hy­
where β1 = ln 2λ−1 0.5β2 = ln 2λ−1 1.5. drodynamic diameter and the electrophoretic mobility of molecules
From Eq. (14) one can calculate that for λ = 1460 pertinent to hy­ were determined for the NaCl concentration of 10− 4 to 0.15 M (at pH 5.6
drated and fully extended chain, Cv = 0.85, which indicates that for and 7.4). This allowed to calculate the electrokinetic charge of KC
macroion molar mass distribution characterized by σM = 0.1 the molecules, which was considerably smaller (in absolute terms) than the
correction is below 1%. However, for samples of large dispersity typical nominal charge calculated assuming full dissociation of ionogenic
to LC (Slootmaekers, Mandel, & Reynaers, 1991) where σ M = 0.3 the groups. This resulted in the low effective ionization degree of the
correction is equal from 7 to 10%. molecule amounting to 31 and 11% for NaCl concentration range of
Using these data, one can calculate the aspect ratio parameter of an 10− 4 to 0.15 M, respectively.
equivalent cylinder by a numerical inversion of Eq. (12) Additionally, the viscometric measurements for dilute KC solutions
( ) confirmed that the intrinsic viscosity attains large values between (2.2
λ = fv− 1 [η]0 (15) ± 0.1) × 103 and (5.0 ± 0.3) × 104 for the NaCl concentration of 0.15
One can also calculate λ using the iterative scheme, which produces and 10− 4 M, respectively, which confirmed that the molecules assume
the following formula (Adamczyk et al., 2018) very extended conformations. Using the intrinsic viscosity and the hy­
( ) drodynamic diameter experimental data the equivalent cylinder diam­
λ=
15[η]0 1
2 (16) eter, the length of the KC chain and the geometrical cross-section area
f1 (λ1 ) were explicitly calculated. For NaCl concentration of 10− 4 M the
experimental equivalent cylinder lengths was equal to 920 nm. For NaCl
where concentration of 0.15 M, the experimental chain length of KC molecule
3 1 was still vary large and equal to 520 nm. These data unequivocally
f1 (λ1 ) = + (17) confirm that the KC molecule behaviour in electrolyte solutions can be
(ln 2λ1 − 0.5) (ln 2λ1 − 1.5)
adequately described in terms of the slender body hydrodynamic model.
( )12 One can argue that these newly acquired data have significance for
andλ1 = 15[η]0 . The precision of Eq. (17) for is ca. 1% for the
predicting and properly interpreting carrageenian molecule thickening
intrinsic viscosity exceeding 100. properties and for producing macroion films in the layer-by-layer pro­
Using Eq. (16) one can calculate that the λ parameter for KC mole­ cesses at various substrates, comprising carrier nano- and
cules and NaCl concentration of 10− 4 M is equal to 1140, which ap­ microparticles.
proaches the value pertinent to the fully extended chain of 1460 derived
from MD modeling. For NaCl concentration of 0.15 M, the λ parameter
significantly decreases to 210. These results unequivocally confirm that Declaration of competing interest
for this broad range of NaCl concentration the KC molecules exhibit an
extended conformation, hence the assumption of the slender body The authors declare no conflict of interest.
regime is fulfilled.
Additionally, knowing λ one can calculate the equivalent cylinder Acknowledgments
diameter of the KC chain using Eq. (8) transformed to the form
c1 ln 2 λ + c2 This work was financially supported by the National Science Centre,
dc = dH0 (18) Poland, Opus Project, UMO-2018/31/B/ST8/03277.
λ

where dH0 is previously determined hydrodynamic diameter of KC References


molecules listed in Table 2.
Consequently, the equivalent cylinder length can be calculated as Lc Abd Karim, A., Sulebele, G. A., Azhar, M. E., & Ping, C. Y. (1999). Effect of carrageenan
on yield and properties of tofu. Food Chemistry, 66(2), 159–165. https://doi.org/
= λdc and the geometrical cross-section area in the side-on orientation is
10.1016/S0308-8146(98)00258-1
given by Sc = dc Lc. Abraham, M. J., Murtola, T., Schulz, R., Páll, S., Smith, J. C., Hess, B., et al. (2015).
Values of dc and Lc calculated in this way are given in Table 3. It is GROMACS : High performance molecular simulations through multi-level
parallelism from laptops to supercomputers. SoftwareX, 1–2, 19–25. https://doi.org/
interesting to mention that the equivalent cylinder lengths for NaCl
10.1016/j.softx.2015.06.001
concentration of 10− 4 M is equal to 920 nm, which approaches the value Adamczyk, Z., Bratek, A., Jachimska, B., Jasiński, T., & Warszyński, P. (2006). Structure
of poly(acrylic acid) in electrolyte solutions determined from simulations and

9
A. Michna et al. Food Hydrocolloids 121 (2021) 107033

viscosity measurements Z. Journal of Physical Chemistry B, 110(45), 22426–22435. Hunter, R. J. (1981). In R. Ottewill, & R. Rowell (Eds.), Zeta potential in colloid science.
https://doi.org/10.1021/jp063981w Principles and applications. Academic Press.
Adamczyk, Z., Cichocki, B., Ekiel-Jezewska, M. L., Słowicka, A., Wajnryb, E., & Imeson, A. P. (2009). Carrageenan and furcellaran. In Handbook of hydrocolloids.
Wasilewska, M. (2012). Fibrinogen conformations and charge in electrolyte Woodhead Publishing Limited. https://doi.org/10.1533/9781845695873.164.
solutions derived from DLS and dynamic viscosity measurements. Journal of Colloid Lee, J., Cheng, X., Swails, J. M., Yeom, M. S., Eastman, P. K., Lemkul, J. A., et al. (2016).
and Interface Science, 385(1), 244–257. https://doi.org/10.1016/j.jcis.2012.07.010 CHARMM-GUI input generator for NAMD, GROMACS, AMBER, OpenMM, and
Adamczyk, Z., Jamroży, K., Batys, P., & Michna, A. (2014). Influence of ionic strength on CHARMM/OpenMM simulations using the CHARMM36 additive force field. Journal
poly(diallyldimethylammonium chloride) macromolecule conformations in of Chemical Theory and Computation, 12(1), 405–413. https://doi.org/10.1021/acs.
electrolyte solutions. Journal of Colloid and Interface Science, 435, 182–190. https:// jctc.5b00935
doi.org/10.1016/j.jcis.2014.07.037 Li, L., Ni, R., Shao, Y., & Mao, S. (2014). Carrageenan and its applications in drug
Adamczyk, Z., Morga, M., Kosior, D., & Batys, P. (2018). Conformations of poly- l -lysine delivery. Carbohydrate Polymers, 103, 1–11. https://doi.org/10.1016/j.
molecules in electrolyte solutions: Modeling and experimental measurements. carbpol.2013.12.008
Journal of Physical Chemistry C, 122(40), 23180–23190. https://doi.org/10.1021/ Luo, M., Shao, B., Nie, W., Wei, X. W., Li, Y. L., Wang, B. L., et al. (2015). Antitumor and
acs.jpcc.8b07606 adjuvant activity of λ-carrageenan by stimulating immune response in cancer
Alba, K., & Kontogiorgos, V. (2019). Seaweed polysaccharides ( agar , alginate immunotherapy. Scientific Reports, 5, 11062. https://doi.org/10.1038/srep11062
carrageenan ). In L. Melton, F. Shahidi, & P. Varelis (Eds.), Encyclopedia of food Manning, G. S. (1979). Counterion binding in polyelectrolyte theory. Accounts of
chemistry (pp. 240–250). Academic Press. https://doi.org/10.1016/B978-0-08- Chemical Research, 12(12), 443–449. https://doi.org/10.1021/ar50144a004
100596-5.21587-4. Mansfield, M. L., & Douglas, J. F. (2008). Transport properties of rodlike particles.
Almutairi, F. M., Adams, G. G., Kök, M. S., Lawson, C. J., Gahler, R., Wood, S., et al. Macromolecules, 41(14), 5422–5432. https://doi.org/10.1021/ma702839w
(2013). An analytical ultracentrifugation based study on the conformation of lambda Michna, A., Adamczyk, Z., Sofińska, K., & Matusik, K. (2017). Monolayers of poly(amido
carrageenan in aqueous solution. Carbohydrate Polymers, 97(1), 203–209. https:// amine) dendrimers on mica – in situ streaming potential measurements. Journal of
doi.org/10.1016/j.carbpol.2013.04.027 Colloid and Interface Science, 485, 232–241. https://doi.org/10.1016/j.
Berth, G., Vukovic, J., & Lechner, M. D. (2008). Physicochemical characterization of jcis.2016.09.007
Carrageenans— A critical reinvestigation. Journal of Applied Polymer Science, 110, Michna, A., Pomorska, A., Nattich-Rak, M., Wasilewska, M., & Adamczyk, Z. (2020).
3508–3524. https://doi.org/10.1002/app.28937 Hydrodynamic solvation of poly(amido amine) dendrimer monolayers on silica.
Bonferoni, M. C., Rossi, S., Tamayo, M., Pedraz, J. L., Dominguez-Gil, A., & Caramella, C. Journal of Physical Chemistry C, 124(32), 17684–17695. https://doi.org/10.1021/
(1993). On the employment of λ-carrageenan in a matrix system . I . Sensitivity to acs.jpcc.0c04638
dissolution medium and comparison with Na carboxymethylcellulose and xanthan Miyamoto, S., & Kollman, P. A. (1992). Settle: An analytical version of the SHAKE and
gum. Journal of Controlled Release, 26, 119–127. RATTLE algorithm for rigid water models. Journal of Computational Chemistry, 13(8),
Bonferoni, M. C., Rossi, S., Tamayo, M., Pedraz, J. L., Dominguez-Gil, A., & Caramella, C. 952–962. https://doi.org/10.1002/jcc.540130805
(1994). On the employment of λ-carrageenan in a matrix system . II . λ-Carrageenan Montolalu, R. I., Tashiro, Y., Matsukawa, S., & Ogawa, H. (2008). Effects of extraction
and hydroxypropylmethylcellulose mixtures. Journal of Controlled Release, 30, parameters on gel properties of carrageenan from Kappaphycus alvarezii
175–182. (Rhodophyta). Journal of Applied Phycology, 20(5), 521–526. https://doi.org/
Bono, A., Anisuzzaman, S. M., & Ding, O. W. (2014). Effect of process conditions on the 10.1007/s10811-007-9284-2
gel viscosity and gel strength of semi-refined carrageenan (SRC) produced from Nakashima, H., Kido, Y., Kobayashi, N., Motoki, Y., Neushul, M., & Yamamoto, N.
seaweed (Kappaphycus alvarezii). Journal of King Saud University - Engineering (1987). Purification and characterization of an avian myeloblastosis and human
Sciences, 26(1), 3–9. https://doi.org/10.1016/j.jksues.2012.06.001 immunodeficiency virus reverse transcriptase inhibitor , sulfated polysaccharides
Brenner, H. (1974). Rheology of a dilute suspension of axisymmetric Brownian particles. extracted from sea algae. Antimicrobial Agents and Chemotherapy, 31(10), 1524–1528.
International Journal of Multiphase Flow, 1(2), 195–341. https://doi.org/10.1016/ Ohshima, H. (Ed.). (2012). Electrical phenomena at interfaces and biointerfaces:
0301-9322(74)90018-4 Fundamentals and applications in nano-, bio-, and environmental sciences. John Wiley &
Briones, A. V., Sato, T., & Bigol, U. G. (2014). Antibacterial activity of polyethylenimine/ Sons, Inc. https://doi.org/10.1002/9781118135440.
carrageenan multilayer against pathogenic bacteria. Advances in Chemical Engineering Park, S., Lee, J., Qi, Y., Kern, N. R., Lee, H. S., Jo, S., et al. (2019). CHARMM-GUI Glycan
and Science, 4(2), 233–241. https://doi.org/10.4236/aces.2014.42026 Modeler for modeling and simulation of carbohydrates and glycoconjugates.
Bussi, G., Donadio, D., & Parrinello, M. (2007). Canonical sampling through velocity Glycobiology, 29(4), 320–331. https://doi.org/10.1093/glycob/cwz003
rescaling. The Journal of Chemical Physics, 126. https://doi.org/10.1063/1.2408420, Parrinello, M., & Rahman, A. (1981). Polymorphic transitions in single crystals: A new
014101. molecular dynamics method. Journal of Applied Physics, 52(12), 7182–7190. https://
Chronakis, I. S., Doublier, J., & Piculell, L. (2000). Viscoelastic properties for kappa- and doi.org/10.1063/1.328693
iota-carrageenan in aqueous NaI from the liquid-like to the solid-like behaviour. Pereira, L., & Critchley, A. T. (2020). The COVID 19 novel coronavirus pandemic 2020:
International Journal of Biological Macromolecules, 28(1), 1–14. https://doi.org/ Seaweeds to the rescue? Why does substantial, supporting research about the
10.1016/S0141-8130(00)00141-0 antiviral properties of seaweed polysaccharides seem to go unrecognized by the
Darden, T., York, D., & Pedersen, L. (1993). Particle mesh Ewald : An N ⋅ log ( N ) method pharmaceutical community in these desperate times? Journal of Applied Phycology,
for Ewald sums in large systems in large systems. The Journal of Chemical Physics, 98, 32(3), 1875–1877. https://doi.org/10.1007/s10811-020-02143-y
10089. https://doi.org/10.1063/1.464397 Reis, R. L., Neves, N., Mano, J. F., Gomes, M. E., Marques, A. P., & Azevedo, H. S. (Eds.).
Dul, M., Paluch, K. J., Kelly, H., Healy, A. M., Sasse, A., & Tajber, L. (2015). Self- (2008). Natural-based polymers for biomedical applications (1st ed.). CRC Press.
assembled carrageenan/protamine polyelectrolyte nanoplexes-investigation of Rodríguez, A., Kleinbeck, K., Mizenina, O., Kizima, L., Levendosky, K., Jean-Pierre, N.,
critical parameters governing their formation and characteristics. Carbohydrate et al. (2014). In vitro and in vivo evaluation of two carrageenan-based formulations
Polymers, 123, 339–349. https://doi.org/10.1016/j.carbpol.2015.01.066 to prevent HPV acquisition. Antiviral Research, 108, 88–93. https://doi.org/10.1016/
Einstein, A. (1908). Elementare theorie der brownschen 1) bewegung. Zeitschrift Für j.antiviral.2014.05.018
Elektrochemie Und Angewandte Physikalische Chemie, 14(17), 235–239. https://doi. Roy, I., & Gupta, M. N. (2003). Smart polymeric Materials : Emerging biochemical
org/10.1002/bbpc.19080141703 applications. Chemistry & Biology, 10, 1161–1171. https://doi.org/10.1016/j.
Gonzalez, M. E., Alarcón, B., & Carrasco, L. (1987). Polysaccharides as antiviral Agents : chembiol .2003.12.004
Antiviral activity of carrageenan. Antimicrobial Agents and Chemotherapy, 31(9), Saha, D., & Bhattacharya, S. (2010). Hydrocolloids as thickening and gelling agents in
1388–1393. food: A critical review. Journal of Food Science & Technology, 47(6), 587–597.
Guvench, O., Greene, S. N., Kamath, G., Brady, J. W., Venable, R. M., Pastor, R. W., et al. https://doi.org/10.1007/s13197-010-0162-6
(2008). Additive empirical force field for hexopyranose monosaccharides. Journal of Schuck, P., Gillis, R. B., Besong, T. M. D., Almutairi, F., Adams, G. G., Rowe, A. J., et al.
Computational Chemistry, 29(15), 2543–2564. https://doi.org/10.1002/jcc.21004 (2014). SEDFIT-MSTAR: Molecular weight and molecular weight distribution
Guvench, O., Mallajosyula, S. S., Raman, E. P., Hatcher, E., Vanommeslaeghe, K., analysis of polymers by sedimentation equilibrium in the ultracentrifuge. Analyst,
Foster, T. J., et al. (2011). CHARMM additive all-atom force field for carbohydrate 139, 79–92. https://doi.org/10.1039/c3an01507f
derivatives and its utility in polysaccharide and carbohydrate À protein modeling. Slootmaekers, D., Mandel, M., & Reynaers, H. (1991). Dynamic light scattering by κ- and
Journal of Chemical Theory and Computation, 7(10), 3162–3180. https://doi.org/ λ-carrageenan solutions. International Journal of Biological Macromolecules, 13(1),
10.1021/ct200328p 17–25. https://doi.org/10.1016/0141-8130(91)90005-F
Harding, S. E. (2017). In An introduction to polysaccharide biotechnology (2nd ed.). CRC Slootmaekers, D., van Dijk, J. A. P. P., Varkevisser, F. A., van Treslong, C. J. B., &
Press. https://doi.org/10.1016/s0300-9084(98)80079-5. Reynaers, H. (1991). Molecular characterisation of κ- and λ-carrageenan by gel
Hariharan, M., Wheatley, T. A., & Price, J. C. (1997). Controlled-release tablet matrices permeation chromatography, light scattering, sedimentation analysis and
from Carrageenans : Compression and dissolution studies. Pharmaceutical osmometry. Biophysical Chemistry, 41(1), 51–59. https://doi.org/10.1016/0301-
Development and Technology, 2(4), 383–393. 4622(91)87209-N
Hess, B. (2008). P-LINCS : A parallel linear constraint solver for molecular simulation. Stechemesser, S., & Eimer, W. (1997). Solvent-dependent swelling of poly(amido amine)
Journal of Chemical Theory and Computation, 4(1), 116–122. https://doi.org/ starburst dendrimers. Macromolecules, 30(7), 2204–2206. https://doi.org/10.1021/
10.1021/ct700200b ma9614914
Hess, B., Bekker, H., Berendsen, H. J. C., & Fraaije, J. G. E. M. (1997). LINCS : A linear Thành, T. T. T., Yuguchi, Y., Mimura, M., Yasunaga, H., Takano, R., Urakawa, H., et al.
constraint solver for molecular simulations. Journal Of Computational Chemistry, 18 (2002). Molecular characteristics and gelling properties of the carrageenan family, 1:
(12), 1463–1472. https://doi.org/10.1002/(SICI)1096-987X(199709)18. Preparation of novel carrageenans and their dilute solution properties.
:12<1463::AID-JCC4>3.0.CO;2-H. Macromolecular Chemistry and Physics, 203(1), 15–23. https://doi.org/10.1002/
Hockney, R. W. (1970). The potential calculation and some applications. Methods in 1521-3935(20020101)203:1<15::AID-MACP15>3.0.CO. ;2-1.
Computational Physics, 9, 135–211. Wasilewska, M., Adamczyk, Z., Pomorska, A., Nattich-Rak, M., & Sadowska, M. (2019).
Human serum albumin adsorption kinetics on silica: Influence of protein solution

10
A. Michna et al. Food Hydrocolloids 121 (2021) 107033

stability. Langmuir, 35(7), 2639–2648. https://doi.org/10.1021/acs. Zabik, M. E., & Aldrich, P. J. (1967). The effect of cations on the viscosity of Lambda-
langmuir.8b03266 Carrageenan. Journal of Food Science, 32(1), 91–97. https://doi.org/10.1111/j.1365-
Zabik, M. E., & Aldrich, P. J. (1965). The effect of selected anions of potassium salts on 2621.1967.tb01966.x
the viscosities of lambda-carrageenan dispersions. Journal of Food Science, 30(1), Zhou, G., Sun, Y., Xin, H., Zhang, Y., Li, Z., & Xu, Z. (2004). In vivo antitumor and
111–117. https://doi.org/10.1111/j.1365-2621.1965.tb00272.x immunomodulation activities of different molecular weight lambda-carrageenans
from Chondrus ocellatus. Pharmacological Research, 50(1), 47–53. https://doi.org/
10.1016/j.phrs.2003.12.002

11

You might also like