You are on page 1of 8

Journal of Molecular Liquids 287 (2019) 111008

Contents lists available at ScienceDirect

Journal of Molecular Liquids

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

Chain conformation of xanthan in solution as influenced by temperature


and salt addition
Cristina-Eliza Brunchi ⁎, Mihaela Avadanei, Maria Bercea, Simona Morariu
“Petru Poni” Institute of Macromolecular Chemistry, 41-A Grigore Ghica Voda Alley, 700487 Iasi, Romania

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

Article history: The capillary viscometry and Fourier transform infrared spectroscopy in attenuated total reflection (ATR-FTIR)
Received 23 April 2019 were used to investigate the conformation of xanthan gum (XG) chains in aqueous solutions, in the absence/pres-
Received in revised form 15 May 2019 ence of NaCl and in the temperature range from 25 to 90 °C. The content of acetate and pyruvate groups per side
Accepted 19 May 2019
of XG was determined by 1H NMR spectroscopy. The viscometric parameters (the intrinsic viscosity, [η], and the
Available online 22 May 2019
hydrodynamic interaction parameter, B) were determined by modeling the experimental data according to the
Keywords:
new Wolf model. The viscosity for XG solution in the absence/presence of NaCl as temperature rises has an ascen-
Xanthan gum dent evolution which takes place with slope variations. These changes were attributed to conformational transi-
Conformational change tion of the chains from the ordered to disordered form affecting the polymer-polymer and polymer-solvent
Viscosity interactions. Salt addition diminishes the electrostatic repulsions between COO− groups, the ordered conforma-
NMR tion is stabilized and the temperature at which conformation changes occur is shifted to higher values (above 55
FTIR °C for XG solution with 10−2 mol/L NaCl). The ordered-disordered transition of the chains was also observed in
ATR–FTIR spectra. Significant shifts in the regions 1750–1550 cm−1, corresponding to vibration of COO− groups,
indicate changes in the hydrogen bonding network.
© 2019 Elsevier B.V. All rights reserved.

1. Introduction degree of substitution with acetate and pyruvate groups depends on


the fermentation process and post-fermentation treatments. In aqueous
Xanthan gum (XG), an extracellular microbial anionic polysaccha- media, the XG chains can adopt a helical ordered conformation (rigid
ride produced by the bacterium Xanthomonas campestris, is continu- structure) or a disordered coil (flexible structure) as a function of tem-
ously in the attention of researchers looking for the improvements of perature, ionic strength, pH, shear rate [24–27], solvent quality (which
the existing materials properties or the obtaining others for new poten- is varied by using solvent/non-solvent mixtures) [28], or other environ-
tial applications. In this context, the researches on XG focus on its chem- mental conditions. In ordered conformation the side chains are folded
ical modification or on the finding of optimal conditions to obtain the and associated with the main chain by hydrogen bonds (i.e., methyl
materials with desired properties. The chemical structure can be group of acetate residues binds to hemiacetal oxygen atoms of alternate
changed even from the production process by controlling the tempera- D-glucose residues from the main chain) [29]. In disordered conforma-
ture [1,2], pH [1], stirring speed [2,3], air flow rate [4], fermentation time tion, the side chains are no longer associated as a result of the repulsion
[5], bacterial strain [6,7] or by subsequent degradation [8–16], electrostatic interactions between COO− groups from pyruvate resi-
functionalization [12,17–19] or oxidation [20,21]. dues. Since the difference in rigidity between the two conformations
The primary structure of this polysaccharide consists of repeating leads to distinct properties in solution, a variety of techniques have
glucose units as main chain and side chains glycosidically linked at been involved in studying the helix-coil transition of xanthan chains
each second glucose residue in C-3 position (Fig. 1) [22]. (such as nuclear magnetic resonance, circular dichroism, optical rota-
The side chains consist of one glucuronic acid residue located be- tion, electron microscopy, differential scanning calorimetry, viscometry,
tween two mannose units. Predominantly, the inner mannose is light scattering, potentiometric titration [30,31], atomic force micros-
substituted stoichiometrically with an acetyl group (at C-6) and the ter- copy [32], dynamic mechanical thermal analysis and phase transition
minal one has variable degrees of pyruvate substitution (at C-4 and C-6) analysis [33]). Another method able to provide valuable information
but recently, other substitution patterns have been identified [23]. The about chains conformation is FTIR spectroscopy. In literature there are
studies related to the chains conformation of some polysaccharide
⁎ Corresponding author. based on shifts in the absorption frequency and/or changes in the ab-
E-mail address: brunchic@icmpp.ro (C.-E. Brunchi). sorptivity of particular absorption bands [34–36] but for xanthan are

https://doi.org/10.1016/j.molliq.2019.111008
0167-7322/© 2019 Elsevier B.V. All rights reserved.
2 C.-E. Brunchi et al. / Journal of Molecular Liquids 287 (2019) 111008

Fig. 1. Primary structure of xanthan gum.

missing. To our knowledge, this method was used to identify the 2.3.2. Viscosity measurements
xanthan structure obtained from different carbon sources [37–40] or Viscosity measurements were performed in temperature range be-
modified by ultrasonic degradation [41] and interactions with other tween 27.5 and 75 °C by using an Ubbelohde capillary viscometer con-
(macro)molecules [17,42–45]. The aim of this study is to investigate nected to a LMV 830 (Lauda, Germany). The temperature was
how the characteristics of XG chains and intermolecular interactions controlled by a circulating-water bath (ECO ET 155). Polymer solutions
change with increasing the temperature or change of ionic strength, with increasingly low concentrations were prepared directly inside the
useful information for food, medicine, cosmetic, agriculture, water viscometer by sequentially dilution of the stock polymer solution. For
treatment or other applications. The conformations adopted of the XG each concentration of XG aqueous solution, the flow time taken into ac-
chains in aqueous solution mediated by salt and temperature were an- count represents an average value of five measurements.
alyzed through Fourier transform infrared spectroscopy (ATR–FTIR)
corroborated with viscometric data. 2.3.3. ATR–FTIR spectroscopy
The FTIR spectral measurements were performed by using the Ver-
2. Experimental tex 70 spectrometer (Bruker, Germany), in the attenuated total reflec-
tion configuration (ATR) on a Golden Gate accessory (single bounce
2.1. Materials diamond crystal, Specac Ltd.). The measurements were carried out in
the temperature range of 26–90 °C with a heating rate of 1 °C/min.
Commercial xanthan gum (XG) and sodium chloride (NaCl) from Millipore water and solutions with cNaCl of 10−5 and 10−2 mol/L were
Sigma Aldrich were used as received. According to relation proposed used as reference. The raw data and the reference were collected as
by Milas et al. [46] the viscometric molecular weight of XG was 128 co-added scans at a 4 cm−1 resolution. The spectra for analysis
determinated as being 1.165 × 106 g/mol [47] and by X-ray fluorescence were obtained by subtracting the reference from the raw spectra. Each
measurements potassium was identified as the metallic ion [48]. sample was subjected to temperature scanning in two separate mea-
surements, in the same experimental conditions. Processing and manip-
ulation of data were made with the OPUS 6.5 software (Bruker Optics,
2.2. Sample preparation Germany).

XG aqueous solutions, salt free and with different salt concentrations 3. Results and discussion
(cNaCl = 10−5 and 10−2 mol/L), were prepared in highly purified deion-
ized water from Milli-Q PF (Millipore, Switzerland) at room tempera- 3.1. 1H NMR analysis
ture under gentle stirring. For viscosity measurements, solutions with
polymer concentrations, c, of 2.1 × 10−2 g/dL and 1 g/dL for ATR–FTIR The content of acetate and pyruvate groups determines the con-
spectroscopy, respectively, were used. Proton nuclear magnetic reso- formation of XG chains in different media, influencing the solution
nance (1H NMR) measurements were performed for XG aqueous solu- properties (e.g., viscosity, rheological properties, phase behavior,
tion (with XG concentration of 0.5 g/dL) with sodium acetate (as etc.). The acetate groups favor the intramolecular attractive interac-
internal standard with concentration of 3 × 10−3 mol/L) prepared in tions stabilizing the ordered conformation of XG while the pyruvate
D2O. The concentration of the salt solution was selected taking into ac- groups prevent the formation of a compact helix due to intramolec-
count the conformational peculiarity and segments mobility of xanthan ular electrostatic repulsion interactions [29,49]. These groups, in or-
chains which have an important impact on the interactions in aqueous dered conformation, cannot be detected in 1H NMR spectra under
solution. Measurements were conducted on homogeneous solution high-resolution condition and room temperature due to the dipole-
equilibrated overnight at room temperature. dipole strong interactions between hydrogen or carbon nuclei
which determine the broadening of resonance line [50] (as exempli-
2.3. Methods fied in Fig. 2 for XG solution). At 25 °C the protons corresponding to
CH 3 groups from acetate and pyruvate units located at 2.16 and
2.3.1. 1H NMR spectroscopy 1.46 ppm, respectively, show very low signals due to the ordered
1
H NMR spectrum for XG aqueous solution was obtained with a and rigid conformation of XG chains in solution.
Bruker Avance DRX 400 spectrometer. The signal corresponding to Increasing temperature (at 50 and 80 °C, respectively) weakens
methyl protons in specific groups from XG (i.e., acetate and pyruvate the dipole-dipole interactions between proton nuclei and the
groups) were expressed in δ (ppm) relative to the protons from CH3 strength of the signals becomes stronger (Fig. 2). The best signals
of sodium acetate (AcNa). Spectrum acquisition was performed at 25, for protons corresponding to acetate and pyruvate groups were ob-
50 and 80 °C. tained at 80 °C when the XG chains have a random coil conformation.
C.-E. Brunchi et al. / Journal of Molecular Liquids 287 (2019) 111008 3

Fig. 2. 1H NMR spectra of XG in D2O (400 MHz) at different temperatures.

In addition, the signals for the protons from carboxyl (ionizable) or from 1H NMR spectrum were used to calculate the number of the ace-
hydroxyl groups overlap and change their position from 4.76 ppm tate (Nac) and pyruvate (Npyr) groups per gram of polysaccharide:
(at 25 °C) to 4.11 ppm (at 80 °C). This shift is due to the existence
of intramolecular H-bonds in which OH groups are implicated as Aac =AAcNa
donor to an alkoxy or hydroxy group [51]. A distinct peak at nac ¼ ð1Þ
c=cAcNa
5.3 ppm corresponding to the equatorial alpha anomeric proton of
mannopyranosic unit carbon and the peaks between 3.4 and
Apyr =AAcNa
3.9 ppm which can be attributed to the hydrogen from CH or CH2 npyr ¼ ð2Þ
c=cAcNa
groups are, also, shown in 1H NMR spectra.
A better quantification of the data provided by 1H NMR spectrum for
deuterated aqueous solution of xanthan in the presence of AcNa at 80 °C where c and cAcNa are the concentrations of XG (g/L) and AcNa (mol/L).
allowed us to add new information on the polysaccharide structure. So, The number of acetate and pyruvate groups (nac, npyr) corresponding
an acetate/pyruvate ratio of 6/5 was determined for the investigated XG to one side chain of XG was determined with the following relationships
considering the number of acetate and pyruvate groups per side chain of [17]:
xanthan.
For this purpose the integrals of the protons signals from substituent M∙Nac
nac ¼ ð3Þ
groups (Aac, Apyr) with reference to the protons from AcNa estimated 1−M pyr ∙N pyr −M ac ∙Nac
4 C.-E. Brunchi et al. / Journal of Molecular Liquids 287 (2019) 111008

M∙Npyr
npyr ¼ ð4Þ of state and, so, its changes can be described by a total differential
1−Mpyr ∙Npyr −Mac ∙Nac
as follows [53]:
     
∂lnη ∂lnη ∂lnη
d lnη ¼ dc þ dT þ dp
Here, M is the molar mass of repeating unit of XG excluding acetate ∂c T;p;γ_ ∂T c;p;γ_ ∂p c;T;γ_
 
and pyruvate groups, Mac and Mpyr are the molar masses of acetate and ∂lnη
þ dγ_ ð6Þ
pyruvate groups, respectively. ∂γ_ c;p;T

3.2. Viscometric investigations


If the first partial differential, Eq. (6), represents the specific hydro-
dynamic volume of polymer, {η}, at a certain polymer concentration
Viscometry is a simple method able to investigate the behavior of
(c), for a given value of temperature (T), pressure (p) and shear rate (γ_
macromolecular chains in dilute solution by means of different visco-
), in the limit of infinite dilution and absence of shear rate, {η} becomes
metric parameters. Between them, the intrinsic viscosity (related to
identical with the intrinsic viscosity.
the hydrodynamic dimensions, conformation and flexibility of polymer
chains) and a viscometric dimensionless parameter (quantifying the
lim fηg ¼ ½η
hydrodynamic interactions that occur between the polymer chain and c→0
_
γ→0
solvent molecules) are the most used.
A commonly used method in determination of the intrinsic viscosity
([η]) for uncharged polymers and polyelectrolites in solvents containing According to this model, [η] can be determined both from initial
an adequate amount of salt is that of extrapolation. For example, the slope of relative viscosity, lnηr, dependence on polyelectrolyte concen-
Huggins extrapolation method described in Eq. (5) can be applied tration (which is independent of any model assumption) and by using
when exists a linear dependence between reduced viscosity, ηsp/c, and the following equation:
polymer concentration, c.
c½η þ Bc2 ½η½η
ln ηr ¼ ð7Þ
ηsp 1 þ Bc½η
¼ ½η þ kH ½η2 c ð5Þ
c
where B, and [η]* are system specific constants representing viscometric
interaction parameters.
Fig. 3a exemplifies, for XG solution with 10−2 mol/L NaCl, the lin- In the present study, the results of viscometric measurements were
ear evolution of ηsp/c as a function of c at 70 °C; the same allure of ηsp/ modeled with Eq. (7).
c - c curve was observed for all temperatures. So, [η] can be obtained The dependence of relative viscosity, lnηr, on the polymer concentra-
as intercept of the linear fit of ηsp/c vs. c. As the salt concentration de- tion for XG solution in the absence/presence of NaCl is presented in
creases, XG solution exhibits the characteristics of polyelectrolyte Fig. 3b only for 70 °C in order to avoid cluttering the chart; for all studied
due to the repulsive Coulombic interactions between charged groups temperatures, the evolution of lnηr - c dependences for XG solution in
located along the main chains. The upturn of reduced viscosity in the the absence/presence of NaCl are the same as at 70 °C.
field of low polymer concentrations describes the expanded confor- In Fig. 3b it can be observed the agreement between experimental
mation of the chains and prevents the determination of [η] by the data and the curves (line) which demonstrates that Eq. (7) is appropri-
above mentioned method. Therefore, we have chosen to evaluate ate to calculate the [η], hydrodynamic interaction parameter B and char-
the experimental viscosity data by a more adequate model [53] acteristic specific hydrodynamic volume [η]*; in all cases, the errors
which demonstrates its ability to describe the viscosity behavior were b10 %. The values of [η] and B obtained as a function of tempera-
for different polyelectrolytes [47,52–54], uncharged polymers ture and cNaCl are presented in Fig. 4.
[55–57], polymer mixtures [44,58], polyelectrolyte complexes [59]. Fig. 4a shows that increasing temperature from 25 to 75 °C leads to
This model was proposed by Wolf and starts from the assumption high values of intrinsic viscosity for XG solution both in the absence
that the viscosity of the dilute polyelectrolyte solutions is a variable and presence of NaCl; the low values of [η] indicate an ordered

Fig. 3. Dependences of ηsp/c (a) and lnηr (b) on the polymer concentration, c, for XG aqueous solutions in the absence/presence of NaCl at 70 °C. The lines in the plot (a) are only to guide the
eye, while in the plot (b) they result from modeling the experimental data with Eq. (7).
C.-E. Brunchi et al. / Journal of Molecular Liquids 287 (2019) 111008 5

Fig. 4. Variation of [η] (a) and B parameter (b) determined with Eq. (7) as a function of temperature for XG aqueous solutions in the absence/presence of NaCl. The plot (a) includes
previous data on [η] for aqueous solutions of XG in the absence/presence of NaCl at 25 °C [47] and in salt free solutions at 37 °C [44]. The lines in both plots are only to guide the eye.

conformation of the chains while the high ones reveal a disordered As can be seen in Fig. 4b, the values of B both in absence/pres-
conformation. ence of NaCl are positive on the entire temperature range typical
Starting from the assertion that at 25 °C in aqueous solution the for the behavior of polymers in thermodynamically good solvents;
xanthan chains exist in disordered and partly ordered conformation in unfavorable thermodynamic conditions the of B are negative
[24], we consider that up to 37 °C the intermolecular associations [47,52]. In the absence of salt and for low concentration NaCl, the
based on hydrogen bonds are destroyed and so, the solution viscosity positive values of B are higher than those obtained in solution
decreases. As the temperature increases (from 37 to 50 °C), the intramo- with cNaCl = 10−2 mol/L. In the range of temperatures between
lecular hydrogen bonds between the side chains and backbone are bro- 25 and 55 °C, the B values in these solvents are close and reach a
ken, the side chains are free to move and initiates the transition of the maximum value at the same temperature (about 50 °C). Up to
main chain conformation (but also of theirs) to a disordered form. The ≈40 °C, the Coulombic and non-Coulombic interactions counter-
disordered conformation is extended due to the repulsion between car- balance each other, but above it the polymer-polymer interactions
boxylate groups [24,60] and explains the high viscosity of the XG solu- became strong, the chain conformation expands and the macro-
tion. Fitzpatrick and co-workers [27] have found that the xanthan molecule responds different to temperature rise. By adding more
chains (in native state) in salt free solution undergo a conformational salt (cNaCl = 10−2 mol/L), the value of B decreases and approaches
transition over a wide range of temperatures (35–55 °C) while, a pre- to zero as a result of shielding repulsion electrostatic interactions
transition attributed to the aggregates disruption, located around 45 and macromolecular chains adopt a shrink coiled conformation
°C, was evidenced by rheological investigations [61]. The high reactivity similar those adopted in poor thermodynamic solvents. In determi-
of XG chains in disordered form as compared with the ordered ones is nation of B values, for XG solution with cNaCl = 10−2 mol/L, for the
responsible for the development of new intermolecular interactions best fit with Eq. (7), [η]* was set zero. [η]* is the characteristic spe-
(such as the repulsion electrostatic interactions as a results of the kinetic cific hydrodynamic volume only required of polyelectrolytes to
energy growth of the side chains or non-specific interactions favored by model the maxima of η sp/c - c plot in the domain of extremely
the temperature increase or energetic and geometric conditions) low c and in the absence of electrostatic repulsion interactions it
[51,60,62] which increase the viscosity at temperatures above 60 °C. becomes zero.
The presence of salt counteracts the repulsion electrostatic interactions
between charged groups and helps to stabilize the ordered conforma- 3.3. Infrared spectral analysis
tion and strengthens the intermolecular interactions. Consequently,
the xanthan chains with such conformation need a higher temperature In order to improve the information resulted from the viscometric
to generate remarkable changes of viscosity (i.e., around 55 °C for solu- behavior of XG chains in aqueous solution, the infrared spectroscopy
tion with cNaCl = 10−2 mol/L). was used. The xanthan gum aqueous solutions in the absence/presence
Conformational modification of polymer chains in solution can of salt were subjected to temperature increase between 26 and 90 °C,
influence the thermodynamic preference of contacts between poly- with an increment of 1 °C/min.
mer segments against those of the polymer chains and solvent mol- Fig. 5 shows the modification at temperature increasing from 26 to
ecules or vice versa. In Eq. (7), B parameter (as the Huggins 90 °C of the FTIR spectra, for the XG solution at three salt concentrations.
constant) quantifies the interactions between polymer segments The asymmetric stretching vibrations of carboxylate groups in the glu-
and gives the shape of lnηr - c plot. For uncharged polymers (i.e., curonic acid and pyruvate units are observed in the 1750–1550 cm−1
XG solution with cNaCl = 10−2 mol/L, Fig. 3b) the lnηr dependence region [64]. The ν(CO) vibrations of the alcoholic COH groups, the
of c is linear while for polyelectrolytes the Coulombic interactions rings skeletal ones and those characteristic to the glycosidic COC link-
between charged groups lead to a curved dependence (i.e., XG solu- ages are all covered in the broad 1170–970 cm−1 band.
tion in the absence of NaCl and with cNaCl = 10−5 mol/L, Fig. 3b). In At room temperature (26 °C), the profile of the bands is influenced
the case of polyelectrolytes, B counts the effect of Coulombic and by the ionic strength of the solution. The νasym(COO−)glu vibration is
non-Coulombic interactions (B = BCoulomb + Bnon-Coulomb); BCoulomb quite stable, being positioned at 1586 cm−1 in water and 1587 cm−1
represents the positive contribution of Coulombic interaction and in 10−2 mol/L NaCl, and dominates almost the entire spectrum. The
Bnon-Coulomb is the negative contribution of non-Coulombic interac- νasym(COO−)pyr, with a finer vibronic structure than νasym(COO−)glu,
tions to the hydrodynamic interaction parameter [63]. have the maximum around 1673 cm−1 in water and around
6 C.-E. Brunchi et al. / Journal of Molecular Liquids 287 (2019) 111008

Fig. 5. Representative temperature–dependent ATR-FTIR in the fingerprint region spectra for XG solution in the absence/presence of NaCl. ΔT = 26–90 °C.

1670 cm−1 in the salt solutions, and the most important shoulder is 1691 cm−1, when cNaCl = 10−2 mol/L. In parallel, the temperature in-
around 1690 cm−1. At room temperature, the νasym(COO−)pyr band be- crease led to the development of the bands around 1024, 1050–1070,
comes lower as the ionic strength of solution increased. When the tem- 1100–1110, 1350 or 1468 cm−1, belonging to ν(CO), δ(COH), δ(CH)
perature was gradually increased, the most spectacular spectral changes and δ(CH2) vibrations in new structures.
occurred at the glucuronic and pyruvate carboxylates, as expected, and As representative features of the movement of the lateral trisaccha-
the effect of temperature was stronger in the saline solutions. In the so- ride chains, the νasym(COO−)glu and νasym(COO−)pyr bands can give a
lutions with no salt or with trace of salt, the position of the νasym(COO−) hint about the onset temperature of the main phase transition. The

glu shifted to the blue, while νasym(COO )pyr shifted to the red, accom- changes in the bands profile with temperature is the sum of the growths
panied by the simultaneous switching in their whole intensity. At 90 and/or decays of some of their components, which shifts the maximum
°C, νasym(COO−)glu is greatly diminished in intensity and is observed of the whole band. As the νasym(COO−)pyr band is isolated in compari-
at 1598 cm−1 in water, and at 1611 cm−1 when cNaCl = 10−5 mol/L. son with νasym(COO−)glu, its evolution under the impact of temperature
In the solution with 10−2 mol/L NaCl, at 90 °C the νasym(COO−)glu prac- increasing was easiest to follow and interpret. The temperature depen-
tically became buried in the more intense band of νsym(COO−)pyr that is dence of the position of νasym(COO−)pyr maximum, exemplified in
centered on 1570 cm−1. Fig. 6a, shows that the temperature of the main transition increases
The two starting components of the νasym(COO−)pyr band, around from ≈46 °C in pure water to ≈52 °C in the low salt solution. When
1670 and 1690 cm−1, were selectively growing up upon temperature cNaCl = 10−2 mol/L, the sub-band at 1690 cm−1 blueshifts constantly
increase as a function of the ionic strength: the sub-band at up to ≈48 °C (reaching 1693 cm−1), where it remains until the temper-
1670 cm−1 increased in pure water, and that at 1690 cm−1 increased ature is around 70 °C. After 70 °C, νasym(COO−)pyr shifts clearly to lower
when cNaCl was 10−2 mol/L. For solution with cNaCl = 10−5 mol/L, the wavenumbers.
two sub-bands increased in a concerted way. As a results, at 90 °C the From the intensity variations presented in Fig. 6b (intensity that
maximum of νsym(COO−)pyr was red-shifted to 1664 cm−1 for XG solu- was normalized to a reference band, 968 cm−1 , a skeletal C-O-C
tion in the absence of NaCl, but greatly displaced to the blue, at mode [65]), the enhancement of the 1690 cm−1 component when
C.-E. Brunchi et al. / Journal of Molecular Liquids 287 (2019) 111008 7

Fig. 6. Temperature – dependent variation of the spectral characteristics of νsym(COO−)pyr: (a) position, given by wavenumber (cm−1); (b) ATR absorbance.

cNaCl = 10−2 mol/L, started after 45 °C and is extended over approx. strength. Remarkable changes noticed in ATR–FTIR spectra of XG aque-
25 °C. This temperature domain roughly corresponds to the plateau ous solutions with temperature and salt concentration are closely re-
domain in the wavenumber plot in Fig. 6a and the ascending part lated to structural organization of its macromolecule and indicate the
of the solution viscosity from Fig. 4a. For the low salt concentration complexity of this process. The movement of the lateral trisaccharide
or no-salt solutions, the dependences of the normalized intensity of chains, by means of the carboxylate groups from pyruvate as terminal
the whole νasym(COO−)pyr band on temperature are very similar to units, indicated the values of the 46 and 52 °C as main temperature tran-
each other (Fig. 6b). They present the same small plateau region be- sition points in the no-salt and low-salt solutions.
tween 30 and 42 °C and identified above with the reorganization of There is a good agreement between the spectral behavior and the
hydrogen bonds in the chains entanglements and disordered re- observations from viscometry measurements, especially for the salt
gions. The increase in intensity of νasym(COO−)pyr took place up to free XG aqueous solution.
≈55 °C in water solution and ≈50 °C for 10−2 mol/L NaCl, followed
by a relative decrease, which was counterbalanced by an increase Acknowledgement
in the bandwidth.
The sudden redshift of νasym(COO−)pyr at a certain temperature This work was supported by a grant of the Ministry of National Edu-
(Fig. 6a) is an indicative of a change in the micro-environment of the py- cation, CNCS–UEFISCDI, project number PN–III–P2–2.1–PED–2016-
ruvate (and glucuronic acids) units, in terms of the strength and nature 1536 (contract 130PED/2017).
of the hydrogen bonds they establish. The two main types of pyruvate
carboxylates (corresponding to the 1690 and 1670 cm−1 components) References
that were selectively evidenced in the higher salt concentration and in
[1] F. Callet, M. Milas, M. Rinaudo, On the role of thermal treatments on the properties
the no-salt solution, respectively, are engaged in hydrogen bonding of of xanthan solutions, Carbohydr. Polym. 11 (1989) 127–137.
different strengths. They exist from the beginning, as the xanthan in so- [2] J.A. Casas, V.E. Santos, F. García-Ochoa, Xanthan gum production under several oper-
lution has the chains in both the folded and unfolded structures, where ational conditions: molecular structure and rheological properties, Enzym. Microb.
Technol. 26 (2000) 282–291.
the terminal carboxylates form intra- and intermolecular hydrogen [3] M. Papagianni, S.K. Psomas, L. Batsilas, S.V. Paras, D.A. Kyriakidis, M. Liakopoulou-
bonds. The detaching of the trisaccharide chains from the main cellulose Kyriakides, Xanthan production by Xanthomonas campestris in batch cultures, Pro-
backbone is connected with an increase in viscosity [66] and with the cess Biochem. 37 (2001) 73–80.
[4] F. Garcıa-Ochoa, E.G. Castro, V.E. Santos, Oxygen transfer and uptake rates during
sudden shift and intensity of the carboxylate groups' bands in the termi-
xanthan gum production, Enzym. Microb. Technol. 27 (2000) 680–690.
nal pyruvate and glucuronic units. [5] F. García-Ochoa, V.E. Santos, J.A. Casas, E. Gómez, Xanthan gum: production, recov-
ery and properties, Biotechnol. Adv. 18 (2000) 549–579.
[6] T. Gumus, A. Sukru Demirci, M. Mirik, M. Arici, Y. Aysan, Xanthan gum production of
4. Conclusions Xanthomonas spp. isolated from different plants, Food Sci. Biotechnol. 19 (2010)
201–206.
[7] S. Rosalam, R. England, Review of xanthan gum production from unmodified
Xanthan gum chain conformation in aqueous solution was discussed starches by Xanthomonas campestris sp, Enzym. Microb. Technol. 39 (2006)
by using capillary viscometry as well as 1H NMR and ATR–FTIR spectros- 197–207.
[8] N.V. Binh, T.B. Diep, H.D. Sang, H.P. Thao, N.T. Thom, T.M. Quynh, Low molecular
copy at different temperatures and salt contents. According to 1H NMR weight xanthan prepared by gamma irradiation and its effects on development of
spectra of XG solution in D2O and AcNa, the xanthan chains exist in or- seedlings, RAD Conf. Proc. 1 (2016) 95–98.
dered conformation at 25 °C (with low signals of protons corresponding [9] B.E. Christensen, O. Smidsrød, A. Elgsaeter, B.T. Stokke, Depolymerization of double-
stranded xanthan by acid hydrolysis: characterization of partially degraded double
to CH3 groups from acetate and pyruvate units), while above 50 °C the
stranded and single-stranded oligomers released from the ordered structures, Mac-
random coils prevail (the signals of protons from the two groups be- romolecules 26 (1993) 6111–6120.
come stronger). By heating the XG solutions, the order-disorder confor- [10] B.E. Christensen, M.H. Myhr, O. Smidsrød, Degradation of double-stranded xanthan
by hydrogen peroxide in the presence of ferrous ions: comparison to acid hydroly-
mational transitions occur and led changes in the balance between
sis, Carbohydr. Res. 280 (1996) 85–99.
Coulombic and non-Coulombic electrostatic interactions and, conse- [11] S.K.H. Gulrez, S. Al-Assafa, Y. Fanga, G.O. Phillips, A.P. Gunning, Revisiting the confor-
quence, the viscosity increases. The viscometric parameters (intrinsic mation of xanthan and the effect of industrially relevant treatments, Carbohydr.
viscosity and hydrodynamic interaction parameter) were determined Polym. 90 (2012) 1235–1243.
[12] S.N. Jampala, S. Manolache, S. Gunasekaran, F.S. Denes, Plasma-enhanced modifica-
according to Wolf model which describes the viscometric behavior of tion of xanthan gum and its effect on rheological properties, J. Agric. Food Chem. 53
macromolecules in various conditions of temperature and ionic (2005) 3618–3625.
8 C.-E. Brunchi et al. / Journal of Molecular Liquids 287 (2019) 111008

[13] R. Li, D.L. Feke, Rheological and kinetic study of the ultrasonic degradation of [40] M.H. Jazini, E. Fereydouni, K. Karimi, Microbial xanthan gum production from alkali
xanthan gum in aqueous solutions, Food Chem. 172 (2015) 808–813. pretreated rice straw, RSC Adv. 7 (2017) 3507–3514.
[14] R. Li, D.L. Feke, Rheological and kinetic study of the ultrasonic degradation of [41] H.M. Saleh, M.S.M. Annuara, K. Simarania, Ultrasound degradation of xanthan poly-
xanthan gum in aqueous solution: effects of pyruvate group, Carbohydr. Polym. mer in aqueous solution: its scission mechanism and the effect of NaCl incorpora-
124 (2015) 216–221. tion, Ultrason. Sonochem. 39 (2017) 250–261.
[15] M. Rinaudo, M. Milas, Enzymic hydrolysis of the bacterial polysaccharide xanthan by [42] A. Shalviri, Q. Liu, M.J. Abdekhodaie, X.Y. Wu, Novel modified starch–xanthan gum
cellulose, Int. J. Biol. Macromol. 2 (1980) 45–48. hydrogels for controlled drug delivery: synthesis and characterization, Carbohydr.
[16] M. Şen, H. Hayrabolulu, P. Taşkın, M. Torun, M. Demeter, M. Cutrubinis, O. Güven, Polym. 79 (2010) 898–907.
Radiation induced degradation of xanthan gum in the solid state, Rad. Phys. [43] Y. Tao, R. Zhang, W. Xu, Z. Bai, Y. Zhou, S. Zhao, Y. Xu, D. Yu, Rheological behavior
Chem. 124 (2016) 225–229. and microstructure of release-controlled hydrogels based on xanthan gum
[17] M. Hamcerencu, J. Desbrieres, M. Popa, A. Khoukh, G. Riess, New unsaturated deriv- crosslinked with sodium trimetaphosphate, Food Hydrocoll. 52 (2016) 923–933.
atives of xanthan gum: synthesis and characterization, Polymer 48 (2007) [44] C.-E. Brunchi, M. Bercea, S. Morariu, M. Avadanei, Investigations on the interactions
1921–1929. between xanthan gum and poly(vinyl alcohol) in solid state and aqueous solutions,
[18] A. Roy, S. Comesse, M. Grisel, N. Hucher, Z. Souguir, F. Renou, Hydrophobically mod- Eur. Polym. J. 84 (2016) 161–172.
ified xanthan: an amphiphilic but not associative polymer, Biomacromolecules 15 [45] I.-P. Merlusca, P. Plamadeala, C. Gîrbea, I.M. Popa, Xanthan-chitosan complex as a
(2014) 1160–1170. potential protector against injurious effects of neomycin, Cell. Chem. Technol. 50
[19] A. Sand, M. Yadav, K. Behari, Graft copolymerization of 2-acrylamidoglycolic acid on (2016) 577–583.
to xanthan gum and study of its physicochemical properties, Carbohydr. Polym. 81 [46] M. Milas, M. Rinaudo, B. Tinland, The viscosity dependence on concentration, molec-
(2010) 626–632. ular weight and shear rate of xanthan solutions, Polym. Bull. 14 (1985) 157–164.
[20] J. Guo, L. Ge, X. Li, C. Mu, D. Li, Periodate oxidation of xanthan gum and its [47] C.-E. Brunchi, S. Morariu, M. Bercea, Intrinsic viscosity and conformational parame-
crosslinking effects on gelatin-based edible films, Food Hydrocoll. 39 (2014) ters of xanthan in aqueous solutions: salt addition effect, Colloids Surf. B:
243–250. Biointerfaces 122 (2014) 512–519.
[21] D. Paiva, C. Gonçalves, I. Vale, M.S.M. Bastos, F.D. Magalhães, Oxidized xanthan gum [48] C.-E. Brunchi, M. Bercea, S. Morariu, M. Dascalu, Some properties of xanthan gum in
and chitosan as natural adhesives for cork, Polymers 8 (2016) 259. aqueous solutions: effect of temperature and pH, J. Polym. Res. 23 (2016) 123.
[22] P.E. Jansson, L. Kenne, B. Lindberg, Structure of extracellular polysaccharide from [49] I.H. Smith, K.C. Symes, C.J. Lawson, E.R. Morris, Influence of the pyruvate content of
Xanthomonas campestris, Carbohydr. Res. 45 (1975) 275–282. xanthan on macromolecular association in solution, Int. J. Biol. Macromol. 3 (1981)
[23] M.M. Kool, H. Gruppen, G. Sworn, H.A. Schols, The influence of the six constituent 129–134.
xanthan repeating units on theorder–disorder transition of xanthan, Carbohydr. [50] M. Rinaudo, M. Milas, F. Lambert, M. Vincendon, lH and 13C NMR investigation of
Polym. 104 (2014) 94–100. xanthan gum, Macromolecules 16 (1983) 816–819.
[24] E. Rochefort, S. Middleman, Rheology of xanthan gum: salt, temperature, and strain [51] B. Bernet, A. Vasella, 1H-NMR analysis of intra- and intermolecular H-bonds of alco-
effects in oscillatory and steady shear experiments, J. Rheol. 31 (1987) 337–369. hols in DMSO: chemical shift of hydroxy groups and aspects of conformational anal-
[25] G. Muller, M. Anrhourrache, J. Lecourtier, G. Chauveteau, Salt dependence of the ysis of selected monosaccharides, inositols, and ginkgolides, Helv. Chim. Acta 83
conformation of a single-stranded xanthan, Int. J. Biol. Macromol. 8 (1989) 167–172. (2000) 995–1021.
[26] Y. Matsuda, Y. Biyajima, T. Sato, Thermal denaturation, renaturation, and aggrega- [52] S. Morariu, C.-E. Brunchi, M. Bercea, The behavior of chitosan in solvents with differ-
tion of a double-helical polysaccharide xanthan in aqueous solution, Polym. J. 41 ent ionic strengths, Ind. Eng. Chem. Res. 51 (2012) 12959–12966.
(2009) 526–532. [53] B.A. Wolf, Polyelectrolytes revised: reliable determination of intrinsic viscosities,
[27] P. Fitzpatrick, J. Meadows, I. Ratcliffe, P.A. Williams, Control of the properties of Macromol. Rapid Commun. 28 (2007) 164−170.
xanthan/glucomannan mixed gels by varying xanthan fine structure, Carbohydr. [54] L. Ghimici, M. Nechifor, B.A. Wolf, Ionic polymers based on dextran: hydrodynamic
Polym. 92 (2013) 1018–1025. properties in aqueous solution and solvent mixtures, J. Phys. Chem. B 113 (2009)
[28] S. Morariu, M. Bercea, C.-E. Brunchi, Phase separation in xanthan solutions, Cell. 8020–8025.
Chem. Technol. 52 (2018) 569–576. [55] C.-E. Brunchi, S. Morariu, M. Cazacu, M. Bercea, Influence of the solvent quality on
[29] F. Callet, M. Milas, M. Rinaudo, Influence of acetyl and pyruvate contents on rheolog- the thermodynamic behavior of polymethylphenylsiloxane solutions, Ind. Eng.
ical properties of xanthan in dilute solution, Int. J. Biol. Macromol. 9 (1987) Chem. Res. 49 (2010) 12740–12746.
291–293. [56] L. Ghimici, C.-E. Brunchi, Titanium dioxide separation from water by PEG and
[30] P. Foss, B.T. Stokke, O. Smidsrød, Thermal stability and chain conformational studies Pluronic type polymers, Sep. Purif. Technol. 103 (2013) 306–312.
of xanthan at different ionic strengths, Carbohydr. Polym. 7 (1987) 421–433. [57] S. Morariu, C.-E. Brunchi, M. Cazacu, M. Bercea, The behavior of poly
[31] L. Bezemer, J.B. Ubbink, J.A. Dekooker, M.E. Kuil, J.C. Leyte, On the conformational (dimethylsiloxane-co-diphenylsiloxane)s in good and theta solvents, J. Chem. Eng.
transitions of native xanthan, Macromolecules 26 (1993) 6436–6446. Data 56 (2011) 1468–1475.
[32] J. Moffat, V.J. Morris, S. Al-Assaf, A.P. Gunning, Visualisation of xanthan conformation [58] C.-E. Brunchi, M. Bercea, S. Morariu, Hydrodynamic properties of polymer mixtures
by atomic force microscopy, Carbohydr. Polym. 148 (2016) 380–389. in solution, J. Chem. Eng. Data 55 (2010) 4399–4405.
[33] I.E. Raschip, I. Yakimets, C.P. Martin, S.S. Paes, C. Vasile, J.R. Mitchell, Effect of water [59] M. Bercea, L.E. Nita, J. Eckelt, B.A. Wolf, Polyelectrolyte complexes: phase diagram
content on thermal and dynamic mechanical properties of xanthan powder: a com- and intrinsic viscosities of the system water/poly(2-vinylpyridinium-Br)/poly(sty-
parison between standard and novel techniques, Powder Technol. 182 (2008) rene sulfonate-Na), Macromol. Chem. Phys. 213 (2012) 2504–2513.
436–443. [60] M. Tako, S. Nakamura, Evidence for intramolecular associations in xanthan mole-
[34] M. Mecozzi, S.Y. Oh, D. Yoo, Y. Shin, H.C. Kim, H.Y. Kim, Y.S. Chung, W.H. Park, J.H. cules, Agric. Biol. Chem. 53 (1989) 1941–1946.
Youk, Crystalline structure analysis of cellulose treated with sodium hydroxide [61] D.M.Q. Craing, A. Kee, S. Tamburic, D. Barnes, An investigation into the temperature
and carbon dioxide by means of X-ray diffraction and FTIR spectroscopy, Carbohydr. dependence of the rheological synergy between xanthan gum and locust bean gum
Res. 340 (2005) 2376–2391. mixtures, J. Biomater. Sci. Polym. Ed. 8 (1997) 377–389.
[35] J. Kong, S. Yu, Fourier transform infrared spectroscopic analysis of protein secondary [62] E.R. Morris, Ordered conformation of xanthan in solutions and “weak gels”: single
structures, Acta Biochim. Biophys. Sin. 39 (2007) 549–559. helix, double helix–or both? Food Hydrocoll. 86 (2019) 18–25.
[36] Y. Hishikawa, E. Togawa, T. Kondo, Molecular orientation in the nematic ordered cel- [63] A. Eich, B.A. Wolf, Intrinsic viscosities of polyelectrolytes: determination and model-
lulose film using polarized FTIR accompanied with a vapor-phase deuteration ing of the effects of extra salt, Chem. Phys. Chem. 12 (2011) 2786–2790.
method, Cellulose 17 (2010) 539–545. [64] H. Li, W. Hou, X. Li, Interaction between xanthan gum and cationic cellulose JR400 in
[37] P. Li, T. Li, Y. Zeng, X. Li, X. Jiang, Y. Wang, T. Xie, Y. Zhang, Biosynthesis of xanthan aqueous solution, Carbohydr. Polym. 89 (2012) 24–30.
gum by Xanthomonas campestris LRELP-1using kitchen waste as the sole substrate, [65] J.J. Cael, J.L. Koenig, J. Blackwell, Infrared and Raman spectroscopy of carbohydrates.
Carbohydr. Polym. 151 (2016) 684–691. Part VI: normal coordinate analysis of V-amylose, Biopolymers 14 (1975)
[38] S.V. Niknezhad, M.A. Asadollahi, A. Zamani, D. Biri, Production of xanthan gum by 1885–1903.
free and immobilized cells of Xanthomonas campestris and Xanthomonas [66] M. Tako, S. Nakamura, Rheological properties of Ca salt of xanthan in aqueous
pelargonii, Int. J. Biol. Macromol. 82 (2016) 751–756. media, Agric. Biol. Chem. 51 (1987) 2919–2923.
[39] Z. Wang, J. Wu, L. Zhu, X.i Zhan, Characterization of xanthan gum produced from
glycerol by a mutantstrain Xanthomonas campestris CCTCC M2015714, Carbohydr.
Polym. 157 (2017) 521–552.

You might also like