Goma Xantana

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Food Hydrocolloids 132 (2022) 107838

Contents lists available at ScienceDirect

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

Effect of downstream processing on the structure and rheological properties


of xanthan gum generated by fermentation of Melaleuca alternifolia
residue hydrolysate
Zhi-Xuan Li a, 1, Jia-Yu Chen a, 1, Yi Wu a, Zhong-Ying Huang a, Shu-Ting Wu a, Yun Chen a, b,
Jing Gao a, b, Yong Hu a, b, **, Chao Huang a, b, *
a
School of Food Science, Guangdong Pharmaceutical University, Zhongshan, 528458, People’s Republic of China
b
GDPU-HKU Zhongshan Biomedical Innovation Platform, Zhongshan, 528437, People’s Republic of China

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

Keywords: Low-cost Melaleuca alternifolia residue hydrolysate has great potential as a promising substrate for xanthan gum
Xanthan gum production but further downstream processing in addition to traditional alcohol-precipitation is necessary to
Downstream processing obtain purified xanthan gum from this substrate. In this study, the effect of downstream processing on the
Structure analysis
structure and rheological properties of xanthan gum generated by fermentation of Melaleuca alternifolia residue
Rheological characterization
hydrolysate was evaluated for the first time. Compared with acid-precipitation and alkali-precipitation, dialysis
Melaleuca alternifolia residue hydrolysate
Fermentation was much more suitable for xanthan gum purification as demonstrated by the energy dispersive spectrometer
(EDS) and thermogravimetric analysis (TGA) results, and the xanthan gum with relatively high purity could be
obtained after alcohol-precipitation and dialysis. Dialysis after alcohol-precipitation showed little influence on
the functional groups of xanthan gum but the microstructure of xanthan gum changed greatly in that a fila­
mentous structure was formed after this treatment. The fermentation performance could be evaluated accurately
based on the suitable downstream processing. The purified xanthan gum had similar rheological properties
(pseudoplasticity and solid-like/gel-like behavior) to the typical xanthan gum according to its flow curves and
linear viscoelastic analysis. However, some rheological properties (existence of Newtonian plateau in flow
curves, and crossover of storage modulus (G′ ) and loss modulus (G′′ ) in frequency sweep test) of the purified
xanthan gum solution were different from that of the typical xanthan gum. Overall, this study can offer some
important information for efficient and effective purification of xanthan gum with a good rheological perfor­
mance from low-cost Melaleuca alternifolia residue hydrolysate.

1. Introduction etc. (Blok et al., 2021; Du et al., 2022; Matsuyama et al., 2021). To make
xanthan gum production more suitable for industrialization, exploiting
Xanthan gum, a microbial polysaccharide mainly produced by Xan­ low-cost substrates for fermentation is always an important research
thomonas campestris, has many attractive characteristics such as special topic. To date, many low-cost feedstocks such as kitchen waste (Li et al.,
rheological properties, good water solubility and compatibility to 2016), glycerol (Wang, Wu, Zhu, & Zhan, 2017), hydrolyzed starch
various salts, good stability to heat, acid, and alkali, and therefore it has (Niknezhad, Asadollahi, Zamani, & Biria, 2016), etc. have been proven
many applications in food, petroleum, tissue engineering, and pharma­ to be suitable for xanthan gum production.
ceutical industry (Andrew, 1977; Huang, Zhong, & Yang, 2020; Kumar, For a long time, essential oils have attracted much attention and
Rao, & Han, 2018; Riaz, Iqbal, Jiang, & Chen, 2021; Sandford, Cottrell, profitable market for their important functions such as antioxidant ac­
& Pettitt, 1984). In particular, xanthan gum is an important food addi­ tivity, antibacterial and antifungal activity, food packaging and pre­
tive that has been widely applied as the thickener, stabilizer, emulsifier, servative, anxiolytic effect, etc. (Falleh, BenJemaa, Saada, & Ksouri,

* Corresponding author. No. 9 Changmingshui Avenue, Wuguishan Street, Zhongshan, 528458, People’s Republic of China.
** Corresponding author. No. 9 Changmingshui Avenue, Wuguishan Street, Zhongshan, 528458, People’s Republic of China.
E-mail addresses: lidhyong@126.com (Y. Hu), huangc@gdpu.edu.cn (C. Huang).
1
Both authors contributed equally to this work.

https://doi.org/10.1016/j.foodhyd.2022.107838
Received 3 January 2022; Received in revised form 1 April 2022; Accepted 27 May 2022
Available online 31 May 2022
0268-005X/© 2022 Elsevier Ltd. All rights reserved.
Z.-X. Li et al. Food Hydrocolloids 132 (2022) 107838

2020; Ju et al., 2019; Tariq et al., 2019; Zhang & Yao, 2019). Among 10, NaCl 5, initial pH 7.0) at 30 ◦ C and 200 rpm for 48 h. Subsequently
various plants, Melaleuca alternifolia (Australian tea tree) plays an 10% (v/v) seed culture was translated into 30 mL fermentation medium
important role in essential oils production for its many advantages such (Melaleuca alternifolia residue hydrolysate, pH 7.0) in 250 mL conical
as high essential oils extraction yield, ease of cultivation, and broad flask, and the fermentation was carried out at 30 ◦ C and 200 rpm. The
activities of its essential oils (Davies, Larkman, Marriott, & Khan, 2016; fermentation broth was taken out periodically for the measurement of
Pereira, de Sant’Anna, Silva, Pinheiro, & de Castro-Prado, 2014; Shep­ biomass, xanthan gum yield, and residual sugars concentration. All
herd, Wood, Raymond, Ablett, & Rose, 2015). Many solid residues are fermentation was carried out in duplicate and the results were expressed
generated after essential oils extraction, e.g., about 2–3 kg Melaleuca as the averages.
alternifolia can generate about 20–30 mL essential oils, and the
remaining solid residues are considered as wastes. Obviously, it is wise 2.3. Biomass and residual sugars concentration determination, and
and necessary to explore some high value-added utilization of the bio­ alcohol-precipitation of crude xanthan gum
resource of plant residues after essential oils extraction (Saha & Basak,
2020). The cells of X. campestris were separated from the fermentation broth
In this work, the hydrolysate of Melaleuca alternifolia residue after by centrifugation at 10610×g for 10 min. After that, the precipitate was
essential oils extraction was used as the substrate for xanthan gum washed by deionised water and recentrifuged at 10610×g for 10 min.
production by Xanthomonas campestris. Generally, xanthan gum in the Then, the biomass was determined by the dry cell weight after lyophi­
fermentation broth can be simply recovered by alcohol-precipitation lization. The residual sugars concentration in the cell-free supernatant of
with high efficiency and easy operation (Li et al., 2016; Nejadman­ fermentation broth after centrifugation was measured by the DNS
souri et al., 2020; Palaniraj & Jayaraman, 2011). However, it is worth method (Miller, 1959). To obtain xanthan gum, the cell-free supernatant
noting that various alcohol insoluble impurities exist in the Melaleuca was added with three volumes (v/v) of ethanol and maintained at
alternifolia residue hydrolysate as found in our pre-research (unpub­ − 20 ◦ C for 1 h and then centrifuged at 10610×g for 15 min. The pre­
lished), and therefore merely alcohol-precipitation is not enough for cipitation of crude xanthan gum, including both pure xanthan gum and
obtaining xanthan gum from this substrate. Obviously, the structure and other alcohol insoluble impurities present in the Melaleuca alternifolia
rheological properties of xanthan gum will be affected greatly due to the residue hydrolysate, was then further dried by lyophilization.
low purity.
In this situation, exploration of suitable downstream processing 2.4. Further xanthan gum separation and purification
(separation and purification) is very important for xanthan gum pro­
duction from Melaleuca alternifolia residue hydrolysate. On the one After alcohol-precipitation and drying, the crude xanthan gum was
hand, the suitable downstream processing can be found by investigating further separated and purified by three different methods. The first
the effect of different downstream processing on the performance of method was acid-precipitation. In detail, the dry alcohol-precipitate
purified xanthan gum. On the other hand, based on the suitable down­ (crude xanthan gum) was suspended in hydrochloric acid solution (pH
stream processing, the accurate evaluation on the efficiency of this 2.0) and shaken at 800 rpm at room temperature for 0.5 h to remove acid
bioconversion could be performed to show its potential for large-scale soluble impurities. Subsequently the acid insoluble precipitate (purified
application. Considering the above discussion, various downstream xanthan gum) was recovered by centrifugation at 10610×g for 15 min
processing steps were applied to separate and purify xanthan gum from and then lyophilized. The second method was alkali-precipitation. In
Melaleuca alternifolia residue hydrolysate, and the effect of downstream detail, the dry alcohol-precipitate (crude xanthan gum) was suspended
processing on the structure and rheological properties of xanthan gum in 0.5 M sodium hydroxide solution and shaken at 800 rpm at room
were evaluated systematically for the first time in this study. Overall, temperature for 0.5 h to remove alkali soluble impurities. Subsequently
this study can offer some important information for efficient and effec­ the alkali insoluble precipitate (purified xanthan gum) was recovered by
tive purification of xanthan gum with a good rheological performance centrifugation at 10610×g for 15 min and then lyophilized. The third
from low-cost Melaleuca alternifolia residue hydrolysate. method was dialysis. In detail, the crude xanthan gum was dissolved in
deionised water and then dialyzed with an 8000 Da dialysis membrane
2. Methods and materials against deionised water for 2 d to remove the small molecule impurities.
The remaining xanthan gum solution was then lyophilized to obtained
2.1. Fermentation substrate preparation dry purified xanthan gum. For each treatment, the polysaccharide
(xanthan gum) content was measured by the phenol-sulphuric acid
Melaleuca alternifolia residue after essential oils extraction by steam method (DuBois, Gilles, Hamilton, Rebers, & Smith, 1956). After further
distillation was smashed (particle size≤0.45 mm) and dried before separation and purification, the xanthan gum yield was expressed as the
usage. Then, the residue was hydrolyzed by 2% (w/v) sulphuric acid at amount of dry purified xanthan gum in per liter fermentation broth.
121 ◦ C for 1 h, and the hydrolysate was recovered by vacuum filtration.
Then, the hydrolysate was neutralized by calcium oxide to pH around 2.5. Structure and rheological characterization
9–11 and recovered by vacuum filtration. After that, the pH of the hy­
drolysate was adjusted to around 5–6 by sulphuric acid and recovered by The morphology of xanthan gum samples was observed by a field
vacuum filtration. Subsequently 1% (w/v) activated carbon was added emission scanning electron microscope (SEM, JSM-7610F PLUS, JEOL,
into the hydrolysate and stirred in 80 ◦ C water bath for 1 h, and then the Japan) operated at a voltage of 15 kV after spray-gold treatment. The
hydrolysate was recovered by vacuum filtration. After the above treat­ elemental analysis of xanthan gum samples was carried out by an energy
ments, the Melaleuca alternifolia residue hydrolysate was used for xan­ dispersive spectrometer (EDS, Ultim Max 40, Oxford Instruments, UK).
than gum production. The sugars concentration of this substrate was The functional groups of xanthan gum samples were measured by a
15.7 g/L. Fourier transform infrared spectroscopy (FT-IR, Thermo Scientific
Nicolet IS5, Thermo Fisher Scientific, USA) in a wavenumber ranging
2.2. Xanthan gum production from 500 to 4000 cm− 1. The thermogravimetric analysis (TGA) was
carried out by a TA SDT 650 (TA Instruments, USA) heated from room
Xanthomonas campestris HC003 (Laboratory of Food Science, temperature to 600 ◦ C with a heating rate of 10 C/min under nitrogen

Guangdong Pharmaceutical University) was used for xanthan gum atmosphere (flow rate of 100 mL/min).
production in this study. Firstly, the pre-culture was performed on pre- The rheological characterization of xanthan gum solutions was
cultivation medium (g/L, glucose 5, xylose 5, yeast extract 3, peptone measured by an MCR-102 rheometer (Anton Paar, Austria). The steady-

2
Z.-X. Li et al. Food Hydrocolloids 132 (2022) 107838

3. Results and discussion

3.1. Effect of downstream processing on the purity of xanthan gum

After fermentation, xanthan gum generated by X. campestris needs to


be recovered from the fermentation broth. Although the technologies of
xanthan gum separation and purification have been studied for many
decades, alcohol-precipitation is usually applied in most situations for
its advantages such as high efficiency, and ease of operation (Borges,
Moreira, Vendruscolo, & Ayub, 2008; Li et al., 2016; Nejadmansouri
et al., 2020; Palaniraj et al., 2011; Yoo & Harcum, 1999; Zhang & Chen,
2010).
However, it is worth noting that alcohol-precipitation might not be
completely suitable for xanthan gum separation and purification from
the fermentation substrate that also contains various alcohol insoluble
impurities because both xanthan gum and alcohol insoluble impurities
will be precipitated after adding alcohol into the fermentation broth. As
Fig. 1. Polysaccharide (xanthan gum) content and residual content of xanthan for the Melaleuca alternifolia residue hydrolysate after detoxification
gum products after different treatments. (a) Crude xanthan gum; (b) xanthan without adding microorganisms for fermentation, various alcohol
gum after acid-precipitation; (c) xanthan gum after alkali-precipitation (d) insoluble compounds were generated after adding three volumes (v/v)
xanthan gum after dialysis.
of ethanol (the ratio of ethanol addition was the same as the alcohol-
precipitation of crude xanthan gum from fermentation broth), suggest­
state viscosity curves were obtained at a temperature ranging from 25 to ing that merely using alcohol-precipitation for xanthan gum separation
45 ◦ C with a shear rate ramp from 0.01 to 1000 1/s. A strain sweep test and purification might not be enough, and more procedures were
was applied to measure the linear viscoelastic properties at a frequency necessary for further xanthan gum separation and purification from
of 10 rad/s in a strain range of 0.01–100%. A frequency sweep test was alcohol insoluble impurities. In addition to alcohol-precipitation, acid-
used to determine the dynamic viscoelastic properties at a constant precipitation, alkali-precipitation, and dialysis also have the potential
strain in the linear viscoelastic region (LVE) obtained from the strain for xanthan gum separation and purification (Bilanovic, Shelef, & Green,
sweep test with frequencies from 0.1 to 100 rad/s. At the same time, a 1994; Smith, 1984; Towle, 1977). Therefore, to evaluate their possibility
complex viscosity measurement was carried out in LVE. All the rheo­ for further xanthan gum separation and purification, all these three
logical characterization was carried out in triplicate and the results were procedures were applied to treat the crude xanthan gum (contained both
expressed as the averages. xanthan gum and alcohol insoluble impurities) after
alcohol-precipitation.
To evaluate the performance of further xanthan gum separation and
purification, the effect of different treatments on polysaccharide

Fig. 2. SEM micrographs of xanthan gum products after different treatments. (A) Crude xanthan gum; (B) xanthan gum after acid-precipitation; (C) xanthan gum
after alkali-precipitation (D) xanthan gum after dialysis.

3
Z.-X. Li et al. Food Hydrocolloids 132 (2022) 107838

Fig. 3. EDS analysis of xanthan gum products after different treatments. (A) Crude xanthan gum; (B) xanthan gum after acid-precipitation; (C) xanthan gum after
alkali-precipitation (D) xanthan gum after dialysis.

(xanthan gum) content of xanthan gum products should be measured 3.2. Effect of downstream processing on the structure of xanthan gum
firstly and it is the most important consideration for choosing the suit­
able method. As shown in Fig. 1, without further treatment, the poly­ 3.2.1. SEM analysis
saccharide content of crude xanthan gum after initial alcohol- The morphology of xanthan gum has been extensively reported in
precipitation was very low (48.8%). Obviously, this crude xanthan many previous studies that the native xanthan gum usually has an
gum was not suitable as the final product and further separation and irregular sheet-like or granular structure (Ahmad & Hasan, 2017;
purification was necessary. After different treatments (acid-precipita­ Kumar, Deepak, Sharma, Srivastava, & Kumar, 2017; Lei et al., 2021;
tion, alkali-precipitation, and dialysis), the polysaccharide content Nejadmansouri et al., 2020; Wang, Xiang, Li, Zhang, & Bai, 2021). It is
(88.2%) after dialysis was much higher than those (66.2% and 62.7%, worth noting that many modifications including the physical and
respectively) after acid-precipitation, and alkali-precipitation, showing chemical methods will alter the morphological structure of xanthan gum
that dialysis was the better choice for further xanthan gum separation and therefore ameliorate the properties of xanthan gum (Bhatia, Ahuja,
and purification. Overall, after dialysis, the high polysaccharide content & Mehta, 2015; He et al., 2022; Wang et al., 2021). Besides improving
(close to 90%) of purified xanthan gum made it much more suitable to be the xanthan gum purity, like these physical and chemical methods,
the final product. further downstream processing might also change the morphological
To be an ideal method for xanthan gum separation and purification, structure of xanthan gum.
it should keep xanthan gum and remove impurities as much as possible. The SEM morphologies of xanthan gum after different downstream
Therefore, in addition to polysaccharide content, the residual content of processing are shown in Fig. 2. As it depicted, the microstructure of
crude xanthan gum products after its further separation and purification crude xanthan gum was irregular granules and many tiny particles
was also measured. As shown in Fig. 1, the residual content after alkali- (possible impurities) covered the surface of these granules (Fig. 2A).
precipitation was higher than that after acid-precipitation, indicating After acid-precipitation, the irregular granules of purified xanthan gum
that the amount of acid soluble impurities was more than that of alkali as shown in the SEM image were much smaller than those of crude
soluble impurities. Compared with acid-precipitation, the residual xanthan gum (Fig. 2B), possibly due to the removal of acid soluble im­
content after dialysis was even lower, suggesting both acid soluble and purities. While for alkali-precipitation, some porous microstructure with
alkali soluble impurities could be removed by dialysis. rougher surface possibly caused by the removal of alkali soluble impu­
rities was observed, but the irregular granules of purified xanthan gum
did not become smaller (Fig. 2C). Overall, the morphological structure

4
Z.-X. Li et al. Food Hydrocolloids 132 (2022) 107838

(Fig. 3A). Among these other elements, the contents of S (7.29%), Ca


(7.17%), and K (3.69%) were relatively high. The elements S and Ca
mainly come from the existence of CaSO4 generated after the neutrali­
zation of the Melaleuca alternifolia residue hydrolysate, while the
element K might come from the hydrolysis of the Melaleuca alternifolia
residue. After acid-precipitation, the contents of K, Na, Mg, Si, and Cl
reduced totally but the contents of S and Ca increased (Fig. 3B) possibly
because HCl cannot react with CaSO4. As for alkali-precipitation,
although the S content reduced greatly, the contents of O and Na
increased greatly due to the adding of NaOH, and the alkali-
precipitation showed little effect on reducing Ca content (Fig. 3C).
Thus, alkali-precipitation seems not suitable for xanthan gum further
separation and purification. As shown in Fig. 3D, dialysis showed a good
performance on the removal of the impurities that the content of all the
elements including S, Ca, K, Na, Mg, Si, and Cl became much lower than
those of crude xanthan gum. The EDS results also proved that dialysis
was suitable for further xanthan gum separation and purification.

3.2.3. FT-IR analysis


The effect of different downstream processing (acid-precipitation,
Fig. 4. FT-IR spectrum of xanthan gum products after different treatments. (a)
Crude xanthan gum; (b) xanthan gum after acid-precipitation; (c) xanthan gum
alkali-precipitation, and dialysis) on the functional groups of xanthan
after alkali-precipitation (d) xanthan gum after dialysis. gum was measured by FT-IR in the wavenumber of 4000–500 cm− 1. As
shown in Fig. 4, the characteristic peaks of xanthan gum which were
reported in many previous studies (Kang et al., 2019; Liu et al., 2021;
of xanthan gum after both acid-precipitation and alkali-precipitation
Nejadmansouri et al., 2020; Sujithra, Deepika, Akshaya, & Ponnusami,
was irregular granules, which was similar to the crude xanthan gum,
2019) could be observed in the FT-IR image of the xanthan gum samples
suggesting that both treatments had merely a small influence on the
after different treatments. In detail, the peaks appeared from 3405 to
morphological structure of xanthan gum. However, the acid-
3449 cm− 1 corresponded to the hydroxyl (-OH) stretching, the peaks of
precipitation and alkali-precipitation might chemically modify (e.g.,
2921–2929 cm− 1 corresponded to the C–H stretching from aldehyde,
deacetylation, depyruvation, etc.) xanthan gum as systematically
CH2 or CH3 groups, the peaks of 1618–1621 cm− 1 corresponded to the
introduced in the recent review article (Riaz et al., 2021), and thus more – O), the peaks of 1405–1413 cm− 1
stretching of the carbonyl group (C–
attention should be paid to their effect on the molecular architecture of
corresponded to the C–H deflection angle, and the peaks of 1062–1116
xanthan gum.
cm− 1 corresponded to the stretching of C–O and acetal groups. It is
Interestingly, compared with acid-precipitation and alkali-
worth noting that the peaks that appeared from 1726 to 1729 cm− 1
precipitation, the dialysis treatment had a much greater influence on
corresponding to acetyl C– – O stretching vibration (Wang, Zhu, Yang, &
the morphological structure of xanthan gum. As shown in Fig. 2D, the
Lu, 2013) could be found in the FT-IR spectra of crude xanthan gum,
microstructure of purified xanthan gum after dialysis changed from the
xanthan gum after acid-precipitation, and xanthan gum after dialysis,
irregular granules to the intertwined filaments with smooth surfaces. In
but disappeared in the FT-IR spectrum of xanthan gum after
the previous study, some special rheological properties such as crossover
alkali-precipitation. In the previous study, alkali (NaOH) treatment was
of storage modulus (G′ ) and loss modulus (G′′ ), higher dynamic viscosity
applied for xanthan gum modification (deacetylation) and the FT-IR
than steady-state shear viscosity (η<|η*|, departure from Cox-Merz
results showed that the peak value of acetyl C– – O stretching vibration
rule), etc. could be found for the semiflexible wormlike xanthan gum
absorption decreased gradually as the NaOH concentration was higher,
sample, which was different from the rheological properties of rodlike
and even disappeared due to the full deacetylation of xanthan gum
xanthan gum sample (Lee & Brant, 2002b). Recently, it was found that
(Wang et al., 2013). Similarly, the peak disappearance from 1726 to
industry-scale microfluidization treatment, an innovative superfine
1729 cm− 1 in the FT-IR spectrum of xanthan gum after
pulverization technology, could form the similar filamentous structure
alkali-precipitation was possibly caused by the deacetylation of xanthan
of xanthan gum as shown in Fig. 2D, and this structure was related to the
gum. Although strong acid treatment with high temperature and long
changing rheological properties of xanthan gum from non-Newtonian
period can hydrolyze xanthan gum and then the composition of xanthan
flow to Newtonian flow (He et al., 2022). As a result, the downstream
gum can be analyzed by high-performance liquid chromatography (Tait,
processing of dialysis after alcohol-precipitation had potential to modify
Sutherland, & Clarke-Sturman, 1990), weak/mild acidic environment
xanthan gum and alter its rheological properties. On the other hand, the
(e.g., decreasing pH caused by cold plasma treatment) might not change
smooth surfaces of intertwined filaments as shown in Fig. 2D demon­
the basic infrared spectra of xanthan gum (Bulbul, Bhushette, Zambare,
strated that most impurities existing in crude xanthan gum have been
Deshmukh, & Annapure, 2019). This can partly explain the little influ­
removed after dialysis. According to the above results, dialysis after
ence of acid-precipitation with relatively mild conditions (diluted hy­
alcohol-precipitation can not only separate and purify xanthan gum
drochloric acid treatment at room temperature for a short period) on the
from the Melaleuca alternifolia residue hydrolysate, but also modify the
FT-IR spectra of xanthan gum in the present study. Compared with
microstructure of xanthan gum, therefore it is an attractive downstream
alkali-precipitation, dialysis mainly removed the small molecule impu­
processing for xanthan gum production from the Melaleuca alternifolia
rities, and thus the peaks that appeared from 1726 to 1729 cm− 1 could
residue hydrolysate.
be observed in the FT-IR spectrum of xanthan gum after dialysis (Pinto,
Furlan, & Vendruscolo, 2011). Overall, although the downstream pro­
3.2.2. EDS analysis
cessing of dialysis after alcohol-precipitation had a great influence on
To further evaluate the performance of different downstream pro­
the microstructure of xanthan gum, it affected little on the functional
cessing, the element composition of crude and purified xanthan gum was
groups of xanthan gum.
analyzed by EDS (Fig. 3). Without further separation and purification, in
addition to its main elements (C and O), the crude xanthan gum con­
3.2.4. TGA analysis
tained various other elements such as S, Ca, K, Na, Mg, Si, and Cl
The thermal degradation behaviors of xanthan gum after different

5
Z.-X. Li et al. Food Hydrocolloids 132 (2022) 107838

Fig. 5. TGA and DTG spectra of xanthan gum products after different treatments. (a) Crude xanthan gum; (b) xanthan gum after acid-precipitation; (c) xanthan gum
after alkali-precipitation (d) xanthan gum after dialysis.

downstream processing were evaluated (Fig. 5). Typically, there are


three stages in the TGA curves of xanthan gum (Bulbul et al., 2019). In
this study, this “three stages” thermal degradation could also be
observed for the crude xanthan gum after alcohol-precipitation. For the
first stage (room temperature to 180 ◦ C), the weight loss was mainly due
to dehydration and degradation of some volatile compounds (impu­
rities). For the second stage (180–385 ◦ C), a fast weight loss happened
caused by the pyrolytic depolymerization. The maximum degradation
rate of crude xanthan gum was 282.49 ◦ C. Finally, after the third stage
(385–600 ◦ C), the remaining percentages of crude xanthan gum sample
was 53.21%, indicating that many impurities existed in the crude xan­
than gum. For the xanthan gum obtained after acid-precipitation, more
stages rather “three stages” thermal degradation was observed. It was
reported that polysaccharides would be degraded after acid treatment
(Trygg, Beltrame, & Yang, 2019). Similarly, it was possible that xanthan
gum was ruptured after acid-precipitation, and therefore its TGA curve
was more complex and contained various stages. For the xanthan gum
obtained after alkali-precipitation, the “three stages” thermal degrada­
tion was not obvious in the TGA curve, and a gradual thermal degra­
dation was observed. In the previous study, acetyl groups had a great Fig. 6. Evolution of biomass, xanthan gum yield, and residual sugars concen­
influence on the thermal degradation behavior of xanthan gum that tration during the fermentation process on Melaleuca alternifolia residue
acetylation on xanthan gum chain could improve its thermal stability hydrolysate.
(Endo, Setoyama, Yamamoto, & Kadokawa, 2015) while breaking of
acetyl groups would make the decomposition temperature of xanthan gum in the following study.
gum to be lower (Sara, Yahoum, Lefnaoui, Abdelkader, &
Moulai-Mostefa, 2020). As introduced above, alkali-precipitation
caused deacetylation of xanthan gum (Wang et al., 2013), and there­ 3.3. Evaluation on fermentation performance based on the suitable
fore reduced the thermal stability of xanthan gum, and decreased the downstream processing for xanthan gum production
temperature of its depolymerization. It is worth noting that the
remaining percentages after thermal degradation of xanthan gum after Although the cost of Melaleuca alternifolia residue hydrolysate is low
acid-precipitation and alkali-precipitation (53.29% and 49.76%, enough to be a fermentation substrate for xanthan gum production, but
respectively) were still close to that of crude xanthan gum, suggesting its potential negative effect on the fermentation should not be ignored.
that many impurities still existed in the xanthan gum. For the xanthan On the one hand, antimicrobial activity is an important application of
gum obtained after dialysis, the “three stages” thermal degradation was Melaleuca alternifolia essential oils (Battisti, Caon, & Machado de Cam­
even more obvious than that of crude xanthan gum. During the first pos, 2021). Although most Melaleuca alternifolia essential oils have been
stage, the weight loss mainly happened from room temperature to extracted after steam distillation, some antimicrobial compounds might
100 ◦ C due to dehydration. For the second stage (pyrolytic depolymer­ still exist in the Melaleuca alternifolia residue. On the other hand, after
ization of xanthan gum), the maximum xanthan gum degradation rate acid hydrolysis of lignocellulosic biomass like the Melaleuca alternifolia
was greater (301.33 ◦ C vs. 282.49 ◦ C). And finally, the remaining per­ residue in this study, various complex and toxic components such as
centage of xanthan gum sample was much lower (22.94% vs. 53.21%). furan and aromatic compounds will also be generated in the lignocel­
These results indicated that many impurities have been removed after lulosic hydrolysates, which will inhibit the growth and products syn­
dialysis. Again, the TGA results showed that dialysis was a promising thesis of microorganisms (Palmqvist & Hahn-Hägerdal, 2000). Thus, it is
method for further separation and purification of xanthan gum. necessary to evaluate the fermentation performance of bioconversion
Considering the above results and discussion, dialysis showed much from Melaleuca alternifolia residue hydrolysate to xanthan gum. How­
better performance for xanthan gum separation and purification than ever, because the crude xanthan gum contains many impurities after
acid-precipitation and alkali-precipitation. Additionally, dialysis is a traditional alcohol-precipitation from Melaleuca alternifolia residue hy­
green technology without the potential pollution from acid and alkali. drolysate. In this situation, it is difficult to learn the fermentation per­
As a result, dialysis after alcohol-precipitation was chosen as the formance of xanthan gum production due to the difficulty in accurate
downstream processing for the separation and purification of xanthan measurement of xanthan gum yield.
As introduced in the previous sections, xanthan gum with high purity

6
Z.-X. Li et al. Food Hydrocolloids 132 (2022) 107838

Table 1 (cheese whey, crude glycerol, sugarcane molasses, and kitchen waste)
Comparison of xanthan gum yield on carbon sources consumption. can be used as media for xanthan gum production. Although the xanthan
Carbon sources Xanthan gum yield on References gum yield on carbon sources (sugars) consumption in this study was
carbon sources lower than that of xanthan gum production from glucose, xylose, and
consumption (%) kitchen waste, it was higher than that of xanthan gum production from
Glucose 56 Mohsin et al. (2021) cheese whey, crude glycerol, and sugarcane molasses. By the compari­
Xylose 53 Mohsin et al. (2021) son and considering the low cost of Melaleuca alternifolia residue hy­
Cheese whey 42 Niknezhad, Asadollahi, Zamani, drolysate and the relatively high xanthan gum yield on sugars
(lactose) Biria, and Doostmohammadi
(2015)
consumption, this bioprocess is efficient and promising for large-scale
Crude glycerol 27.5 Wang et al. (2016) application.
Sugarcane molasses 22.4 Khosravi-Darani et al. (2009)
Kitchen waste 67.07 Li et al. (2016) 3.4. Effect of downstream processing on the rheological properties of
Melaleuca 44.2 This study
xanthan gum
alternifolia
residue
hydrolysate 3.4.1. Flow curves
As a classical microbial polysaccharide that has been widely applied
as a thickener and stabilizer in the food industry, xanthan gum is famous
can be obtained with the downstream processing of dialysis after for its rheological properties in its solution. Generally, xanthan gum
alcohol-precipitation. Based on this technology, the xanthan gum yield solution has high viscosity in low shear and low concentration, and high
can be measured accurately, and therefore the fermentation process of stability in the environment of heat, acid, and alkali (Faria et al., 2011;
xanthan gum production is evaluated in this part (Fig. 6). In this study, Tian et al., 2015; Wu et al., 2021). As the pseudoplastic behavior is an
simply treated by calcium oxide neutralization and activated carbon important characteristic of xanthan gum solution, the flow curves of the
adsorption, the Melaleuca alternifolia residue hydrolysate was used for xanthan gum solution (1%) were evaluated firstly at 25 ◦ C for both the
xanthan gum production. As shown in Fig. 6, there was almost no lag crude xanthan gum without further purification and the xanthan gum
phase existing during this bioprocess that during the first 12 h of purified by dialysis after alcohol-precipitation (Fig. 7A). As shown in
fermentation, the synthesis of xanthan gum was quickly, suggesting that Fig. 7A, both the solutions of crude xanthan gum and purified xanthan
the fermentation environment of the Melaleuca alternifolia residue hy­ gum showed an obvious pseudoplasticity. Namely, the xanthan gum
drolysate showed little inhibition on the biosynthesis of xanthan gum by solutions had high viscosity at low shear rates possibly caused by
X. campestris. Similarly, the sugars utilization was very fast during the polymer entanglements and form a partial self-associated conformation;
first 12 h of fermentation, showing that X. campestris had adapted the however, as the shear rate increased, the weak structure of polymer
fermentation environment of the Melaleuca alternifolia residue hydro­ entanglements was changed to orienting polymer chains, and therefore
lysate quickly. The bioprocess of xanthan gum production has also been an obvious shear-thinning behavior (the viscosity of xanthan gum so­
evaluated in some previous studies with different fermentation sub­ lution decreased with the higher shear rate) was observed (Jang, Zhang,
strates and in many cases, the xanthan gum synthesis was usually later Chon, & Choi, 2015; Xu, Xu, Liu, Chen, & Gong, 2013). Generally, the
than that of cell growth of X. campestris (Li et al., 2016; Wang, Wu, Zhu, characteristic of shear-thinning in the solution of polysaccharides is
& Zhan, 2016). Interestingly, in this study, the cell growth and xanthan beneficial as dysphagia diets (Wei, Guo, Li, Ma, & Zhang, 2021). It is
gum synthesis of X. campestris were simultaneous from the first 12 h of worth noting that the viscosity of the solution of crude xanthan gum was
fermentation as the consumption of fermentable sugars. After 12 h of much lower than that of xanthan gum purified by dialysis after
fermentation, the biomass and xanthan gum yield continued to increase alcohol-precipitation especially in low shear rate (Fig. 7A), indicating
and the xanthan gum synthesis became even faster during 12–24 h of that the impurities in the crude xanthan gum affected the viscosity of its
fermentation. After 24 h of fermentation, the cell growth and xanthan solution greatly.
gum synthesis of X. campestris were slower and the biomass (1.7–2.2 Typically, a Newtonian plateau exists for the solutions of many food
g/L) and xanthan gum yield (4.8–5.0 g/L) were stable after 36 h and 48 gums/hydrocolloids such as guar gum and tara gum under low shear
h of fermentation, respectively due to the exhaustion of fermentable rate (e.g., <1 s− 1) as shown in the flow curves (Kongjaroen, Meth­
sugars in the medium. Overall, the xanthan gum yield was 5.0 g/L after acanon, & Gamonpilas, 2022). In contrast, this Newtonian plateau is
48 h of fermentation, and the xanthan gum yield on sugars consumption hardly observed in the solution of xanthan gum (He et al., 2022) and
was 44.2% (g/g). The performance of this bioprocess was compared even in the solutions of hybrid containing xanthan gum with the same
with other xanthan gum productions from various substrates. As shown shear rate (Kongjaroen et al., 2022), and usually, the Newtonian plateau
in Table 1, both the typical lignocellulose-derived fermentable sugars can be only found for the xanthan gum solution under extremely low
(glucose, xylose) (Zhou, Liu, & Zhao, 2021) and low-cost substrates shear rate (e.g., <10− 2 s− 1) (Yahoum, Moulai-Mostefa, & Le Cerf, 2016).

Fig. 7. (A) Flow curves of solutions (1%) of (a) purified xanthan gum and (b) crude xanthan gum; (B) Flow curves of solutions (1%) of purified xanthan gum under
different temperatures.

7
Z.-X. Li et al. Food Hydrocolloids 132 (2022) 107838

Table 2
The rheological parameters obtained by Cross model for the solutions of both
crude and purified xanthan gum.
T/oC η0 /mPa⋅s η∞ /mPa⋅s k n R2

Crude xanthan gum 25 803 2.092 0.483 0.672 0.998


Purified xanthan 25 23298 8.714 3.710 0.769 0.992
gum 35 12002 6.742 2.314 0.754 0.994
45 6785 3.357 1.448 0.733 0.998

As shown in Fig. 7A, a Newtonian plateau existed for the solution of


crude xanthan gum under a low shear rate (1 s− 1), which was similar to
the laws of the solutions of guar gum and tara gum (Kongjaroen et al.,
2022), indicating that the impurities in the crude xanthan gum changed
the laws of flow curves of its solution. Interestingly, a Newtonian plateau
was also found for the solution of purified xanthan gum with a relatively
high shear rate (10− 1 s− 1) when compared with the previous study
(Yahoum et al., 2016). Generally, no Newtonian plateau was observed
for xanthan gum solution under this shear rate. The relationship be­
tween the structure and rheological properties of xanthan gum have
systematically learned in many previous studies, and the structure of Fig. 8. Strain sweep dependency of storage modulus (G′ ) and loss modulus (G′′ )
xanthan gum can be reflected from its rheological properties (Ross-­ of solution (1%) of (a) purified xanthan gum and (b) crude xanthan gum.
Murphy, 1995; Shatwell, Sutherland, Ross-Murphy, & Dea, 1990).
Usually, xanthan gum has a typical weak-gel network in a salt solution the charge shielding effect (Alquraishi & Alsewailem, 2012). In addition
(Ross-Murphy, 1995), but this situation might be different in a solution to the flow curves measured at 25 ◦ C, the flow curves of solutions of
without salt. In this study, the xanthan gum solution was prepared by purified xanthan gum were also evaluated at higher temperatures (35
deionised water (in absence of salt), and the Newtonian plateau existing and 45 ◦ C) and fitted by the Cross model (Fig. 7B and Table 1). The
in the flow curves indicated that the purified xanthan gum had an temperature had an obvious influence on the pseudoplasticity of xan­
entanglement network (Shatwell et al., 1990). Obviously, the alteration than gum solutions (Fig. 7). Typically, increasing temperature will cause
of rheological properties was due to the variation of xanthan gum order-disorder transition of xanthan gum from an ordered state (helix)
structure, and this could also be found from the research on the visco­ to a disordered conformation (random coil) which involves the changes
elasticity (discussed in section 3.4.2). It is worth noting that some of polymer network and dissociation of the side chains from the back­
modifications such as microfluidization treatment and high-pressure bone (Bercea & Morariu, 2020; Pelletier et al., 2001). Thus, the viscosity
homogenization can change xanthan gum solution from and pseudoplasticity of xanthan gum solution decrease with the higher
non-Newtonian flow to Newtonian flow (Eren, Santos, & Campanella, temperature. As shown in Table 1, all the n values in Table 1 were less
2015; He et al., 2022). Obviously, the downstream processing of dialysis than 1, showing that all the samples tested were pseudoplastic fluids. It
after alcohol-precipitation could fulfill a similar effect as these advanced is worth noting that the values of η0 , η∞ , k, and n of the solutions of
modification technologies. purified xanthan gum became lower as the temperature increased, again
To date, various models have been developed to fit the rheological showing that the viscosity and pseudoplasticity of xanthan gum solution
results, and the Cross model was used widely to fit the flow curves of decreased with the higher temperature.
non-Newtonian fluids (Cross, 1965; Eren et al., 2015; Kongjaroen et al.,
2022; Xu et al., 2013; Yahoum et al., 2016). To depict the rheological 3.4.2. Viscoelasticity and complex viscosity
laws of crude and purified xanthan gum more systematically, the Besides flow curves, viscoelasticity is also an important rheological
rheological result of flow curves was further fitted to the classical Cross property of xanthan gum solution that usually exhibits gel-like behavior
model (Cross, 1965; Pelletier, Viebke, Meadows, & Williams, 2001; (Milas, Rinaudo, Knipper, & Schuppiser, 1990). The linear viscoelastic
Yahoum et al., 2016): behavior of the solutions (1%) of both crude and purified xanthan gum
was learned by a strain sweep measurement firstly. As shown in Fig. 8,
η0 − η∞
η= + η∞ both the storage modulus (G′ ) and the loss modulus (G′′ ) changed little
1 + kγ̇n
with the variation of strain sweep. Interestingly, the downstream pro­
where η0 is the zero-shear viscosity, η∞ is the infinite-shear viscosity, γ̇ is cessing of dialysis after alcohol-precipitation exhibited a great influence
the shear rate, k is the consistency factor, and n is the flow index. on the G′ and G′′ of xanthan gum solutions that the G′ and the G′′ of
The results of flow curves exhibited a good fit to the Cross model with solution of purified xanthan gum were much higher than that of solution
the correlation coefficients (R2) totally higher than 0.99 and the rheo­ of crude xanthan gum. Moreover, the G′′ was higher than G′ within the
logical parameters acquired from the Cross model are summarized in strain range tested for the solution of crude xanthan gum, suggesting
Table 2. As shown in Table 2, the η0 of the solution of purified xanthan that the solution of crude xanthan gum exhibited a typical viscous-like
gum was much higher than that of the solution of crude xanthan gum behavior. Obviously, this linear viscoelastic behavior is quite different
(23298 mPa s vs. 803 mPa s), again showing that the viscosity of xan­ from the typical xanthan gum solution which usually exhibits a
than gum was affected greatly by the impurities in the crude xanthan solid-like/gel-like behavior (Eren et al., 2015; Sara et al., 2020; Wei
gum. As discussed in the previous sections, the impurities in the xanthan et al., 2021). In contrast, after downstream processing of dialysis after
gum reduced the purity of xanthan gum, and thus the actual xanthan alcohol-precipitation, the G′ was totally higher than G′′ within the strain
gum concentration was lower in the solution (1%) of crude xanthan gum range tested for the solution of purified xanthan gum, showing that the
than that in the solution (1%) of purified xanthan gum. This can partly elastic property was much more predominant than the viscous one
explain the different η0 of the solutions of purified xanthan gum and within the linear viscoelastic region (LVE). As a result, suitable down­
crude xanthan. Moreover, as the crude xanthan gum might contain stream processing was necessary to obtain purified xanthan gum whose
various salts according to the EDS analysis (Fig. 3A), the existence of solution has a solid-like/gel-like behavior.
these salts in the xanthan gum solution also decreased the viscosity by For deeper learning the rheological property of the solution of

8
Z.-X. Li et al. Food Hydrocolloids 132 (2022) 107838

Fig. 9. (A) Frequency sweep dependency of storage modulus (G′ ) and loss modulus (G′′ ) of solution (1%) of purified xanthan gum; (B) Dependency of steady-state
shear viscosity (η) and complex viscosity (|η*|) of solution (1%) of purified xanthan gum versus shear rate and frequency respectively.

purified xanthan gum, the dynamic viscoelastic behavior of the xanthan observed for the semiflexible wormlike xanthan gum sample rather than
gum solution (1%) was measured in a constant strain in the LVE. As the rodlike xanthan gum sample (Lee et al., 2002b), again showing that
shown in Fig. 9A, the slope of storage modulus (G′ ) was higher than that the downstream processing of dialysis after alcohol-precipitation could
of loss modulus (G′′ ). With low frequency, G′′ was higher than G′ , and as modify the structure of xanthan gum and thus alter its rheological
the frequency became higher, G′ was higher than G′′ with high fre­ properties.
quency. Therefore, two regions were separated by the crossover point,
and the critical frequency (ωc) and relaxation time (tR = 2π/ωc) were 4. Conclusions
0.734 rad/s, and 8.560 s, respectively. For the region with the frequency
lower than the crossover point, the xanthan gum solution had viscous- To fulfill and learn the bioconversion from low-cost Melaleuca
like behaviors (G’’> G′ ) while for the region with the frequency alternifolia residue after essential oils extraction to high value-added
higher than the crossover point, the xanthan gum solution had solid- food additive xanthan gum, exploration of suitable downstream pro­
like/gel-like behavior (G’> G′′ ). The relationship between the dynamic cessing is necessary. After systematical comparison, dialysis after
viscoelastic properties and structure of xanthan gum has been elucidated alcohol-precipitation was found to be suitable for xanthan gum sepa­
in many previous studies (Lee & Brant, 2002a; Lee et al., 2002b). ration and purification, and based on this method, the bioconversion
Different from our results, no crossover of G′′ and G′ was observed for process could be evaluated accurately. The effect of downstream pro­
rodlike xanthan gum sample at various salt concentrations (Lee et al., cessing on the structure and rheological properties of xanthan gum was
2002a). Moreover, it was reported in many studies that G′ was always systematically studied with a series of characterization including SEM,
higher than G′′ for a native xanthan gum over the entire frequency range EDS, FT-IR, TGA, and rheological analysis. In addition to xanthan gum
in dynamic viscoelasticity measurement (Eren et al., 2015; Sara et al., separation and purification, downstream processing of dialysis after
2020; Yahoum et al., 2016). To obtain the similar rheological property alcohol-precipitation also modified the structure of xanthan gum and
of dynamic viscoelasticity like this study (existence of the critical fre­ therefore affect the rheological properties of xanthan gum. Overall, the
quency), suitable physical or chemical modifications (e.g., high-pressure purified xanthan gum with a good rheological performance could be
homogenization, carboxymethylation, and alkylation) were carried out obtained efficiently and effectively from the low-cost Melaleuca alter­
in these studies (Eren et al., 2015; Sara et al., 2020; Yahoum et al., nifolia residue hydrolysate, and therefore this substrate is attractive for
2016). In contrast to the rodlike xanthan gum sample, the crossover of industrial xanthan gum production. Future studies should focus on the
G′′ and G’ could be observed for the semiflexible wormlike xanthan gum scale-up of the downstream processing especially for exploring prom­
sample (Lee et al., 2002b), and this dynamic viscoelastic property was ising (efficient, cost-effective, reusable, and durable) material, equip­
similar to our results. The dynamic viscoelastic properties indicated that ment, and technology for large-scale dialysis considering its relatively
in addition to purification, the structure of xanthan gum could also be high cost. Moreover, the application of xanthan gum obtained from the
modified by the downstream processing of dialysis after low-cost Melaleuca alternifolia residue hydrolysate should be learned to
alcohol-precipitation, and the structure alteration might result in the promote its industrialization.
change of rheological properties of xanthan gum solution.
Besides the frequency sweep test, the complex viscosity (dynamic CRediT authorship contribution statement
viscosity) of xanthan gum solution was also analyzed. Fig. 9B depicts the
relationship between steady-state shear viscosity (η) and complex vis­ Zhi-Xuan Li: Conceptualization, Methodology, Validation, Formal
cosity (|η*|) versus shear rate and frequency respectively. As can be analysis, Investigation, Data curation, Writing – original draft, Writing –
observed, |η*| was higher than η with the shear rates and frequencies review & editing. Jia-Yu Chen: Conceptualization, Validation, Formal
tested, and the deviation between |η*| and η became greater as the shear analysis, Investigation, Data curation, Writing – original draft, Visuali­
rates and frequencies were higher. Obviously, the solution of purified zation. Yi Wu: Investigation, Formal analysis, Visualization. Zhong-
xanthan gum did not obey the Cox-Merz rules, where η≈|η*| at a given Ying Huang: Investigation, Formal analysis. Shu-Ting Wu: Validation,
shear rate and frequency (Cox & Merz, 1958). The deviation between |η Investigation. Yun Chen: Supervision, Funding acquisition. Jing Gao:
*| and η (η<|η*|) was mainly caused by the elastic gel-like structure Supervision, Funding acquisition. Yong Hu: Writing – review & editing,
which is broken under steady-state shear but affected little under Supervision, Project administration, Funding acquisition. Chao Huang:
oscillatory measurements (Martín-Alfonso, Cuadri, Berta, & Stading, Writing – original draft, Writing – review & editing, Methodology, Su­
2018). Thus, the departure from Cox-Merz rule confirmed that the pu­ pervision, Resources, Project administration, Funding acquisition.
rified xanthan gum sample had a weak-gel structure, which was in good
accordance with other xanthan gum solution reported in the previous
study (Abson et al., 2014). As introduced previously, this special rheo­ Declaration of competing interest
logical property (η<|η*|, departure from Cox-Merz rule) could be
The authors declare that they have no known competing financial

9
Z.-X. Li et al. Food Hydrocolloids 132 (2022) 107838

interests or personal relationships that could have appeared to influence Kang, Y., Li, P., Zeng, X., Chen, X., Xie, Y., Zeng, Y., et al. (2019). Biosynthesis, structure
and antioxidant activities of xanthan gum from Xanthomonas campestris with
the work reported in this paper.
additional furfural. Carbohydrate Polymers, 216, 369–375.
Khosravi-Darani, K., Farhadi, G., Mohammadifar, M., Hadian, Z., Seyed Ahmadian, F.,
Acknowledgments Komeili, R., et al. (2009). Comparison of bench-scale production of xanthan by
Xanthomonas campestris in solid state and submerged fermentation. Iranian Journal
of Nutrition Sciences & Food Technology, 4(1), 49–56.
This study was supported by Guangdong Basic and Applied Basic Kongjaroen, A., Methacanon, P., & Gamonpilas, C. (2022). On the assessment of shear
Research Foundation (2019A1515011843, 2019A1515010640, and extensional rheology of thickened liquids from commercial gum-based
2021A1515110999), Special project in key fields of colleges and uni­ thickeners used in dysphagia management. Journal of Food Engineering, 316, Article
110820.
versities in Guangdong Province (2021ZDZX4001), Special Support Kumar, A., Deepak, Sharma, S., Srivastava, A., & Kumar, R. (2017). Synthesis of xanthan
Project of Guangdong Province (2016TQ03N881), National Natural gum graft copolymer and its application for controlled release of highly water
Science Foundation of China (22078070), and the Science and Tech­ soluble Levofloxacin drug in aqueous medium. Carbohydrate Polymers, 171,
211–219.
nology Program of Zhongshan (2021B2001, 2020B2068, 2021B2048). Kumar, A., Rao, K. M., & Han, S. S. (2018). Application of xanthan gum as polysaccharide
in tissue engineering: A review. Carbohydrate Polymers, 180, 128–144.
References Lee, H.-C., & Brant, D. A. (2002a). Rheology of concentrated isotropic and anisotropic
xanthan solutions. 1. A rodlike low molecular weight sample. Macromolecules, 35(6),
2212–2222.
Abson, R., Gaddipati, S. R., Hort, J., Mitchell, J. R., Wolf, B., & Hill, S. E. (2014).
Lee, H.-C., & Brant, D. A. (2002b). Rheology of concentrated isotropic and anisotropic
A comparison of the sensory and rheological properties of molecular and particulate
xanthan solutions. 2. A semiflexible wormlike intermediate molecular weight
forms of xanthan gum. Food Hydrocolloids, 35, 85–90.
sample. Macromolecules, 35(6), 2223–2234.
Ahmad, R., & Hasan, I. (2017). Efficient remediation of an aquatic environment
Lei, Y., Gao, S., Xiang, X., Li, X., Yu, X., & Li, S. (2021). Physicochemical, structural and
contaminated by Cr(VI) and 2,4-dinitrophenol by XG-g-polyaniline@ZnO
adhesion properties of walnut protein isolate-xanthan gum composite adhesives
nanocomposite. Journal of Chemical & Engineering Data, 62(5), 1594–1607.
using walnut protein modified by ethanol. International Journal of Biological
Alquraishi, A. A., & Alsewailem, F. D. (2012). Xanthan and guar polymer solutions for
Macromolecules, 192, 644–653.
water shut off in high salinity reservoirs. Carbohydrate Polymers, 88(3), 859–863.
Li, P., Li, T., Zeng, Y., Li, X., Jiang, X., Wang, Y., et al. (2016). Biosynthesis of xanthan
Andrew, T. R. (1977). Applications of xanthan gum in foods and related products.
gum by Xanthomonas campestris LRELP-1 using kitchen waste as the sole substrate.
Extracellular Microbial Polysaccharides, 45, 231–241. AMERICAN CHEMICAL
Carbohydrate Polymers, 151, 684–691.
SOCIETY.
Liu, Y., Zhu, Y., Wang, Y., Quan, Z., Zong, L., & Wang, A. (2021). Synthesis and
Battisti, M. A., Caon, T., & Machado de Campos, A. (2021). A short review on the
application of eco-friendly superabsorbent composites based on xanthan gum and
antimicrobial micro- and nanoparticles loaded with Melaleuca alternifolia essential
semi-coke. International Journal of Biological Macromolecules, 179, 230–238.
oil. Journal of Drug Delivery Science and Technology, 63, Article 102283.
Martín-Alfonso, J. E., Cuadri, A. A., Berta, M., & Stading, M. (2018). Relation between
Bercea, M., & Morariu, S. (2020). Real-time monitoring the order-disorder
concentration and shear-extensional rheology properties of xanthan and guar gum
conformational transition of xanthan gum. Journal of Molecular Liquids, 309, Article
solutions. Carbohydrate Polymers, 181, 63–70.
113168.
Matsuyama, S., Kazuhiro, M., Nakauma, M., Funami, T., Nambu, Y., Matsumiya, K., et al.
Bhatia, M., Ahuja, M., & Mehta, H. (2015). Thiol derivatization of Xanthan gum and its
(2021). Stabilization of whey protein isolate-based emulsions via complexation with
evaluation as a mucoadhesive polymer. Carbohydrate Polymers, 131, 119–124.
xanthan gum under acidic conditions. Food Hydrocolloids, 111, Article 106365.
Bilanovic, D., Shelef, G., & Green, M. (1994). Xanthan fermentation of citrus waste.
Milas, M., Rinaudo, M., Knipper, M., & Schuppiser, J. L. (1990). Flow and viscoelastic
Bioresource Technology, 48(2), 169–172.
properties of xanthan gum solutions. Macromolecules, 23(9), 2506–2511.
Blok, A. E., Bolhuis, D. P., Kibbelaar, H. V. M., Bonn, D., Velikov, K. P., & Stieger, M.
Miller, G. L. (1959). Use of dinitrosalicylic acid reagent for determination of reducing
(2021). Comparing rheological, tribological and sensory properties of
sugar. Analytical Chemistry, 31(3), 426–428.
microfibrillated cellulose dispersions and xanthan gum solutions. Food Hydrocolloids,
Mohsin, A., Akyliyaevna, K. A., Zaman, W. Q., Hussain, M. H., Mohsin, M. Z., Al-
121, Article 107052.
Rashed, S., et al. (2021). Kinetically modelled approach of xanthan production using
Borges, C. D., Moreira, A. D. S., Vendruscolo, C. T., & Ayub, M. A. Z. (2008). Influence of
different carbon sources: A study on molecular weight and rheological properties of
agitation and aeration in xanthan production by Xanthomonas campestris pv pruni
xanthan. International Journal of Biological Macromolecules, 193, 1226–1236.
strain 101. Revista Argentina de Microbiología, 40(2), 81–85.
Nejadmansouri, M., Shad, E., Razmjooei, M., Safdarianghomsheh, R., Delvigne, F., &
Bulbul, V. J., Bhushette, P. R., Zambare, R. S., Deshmukh, R. R., & Annapure, U. S.
Khalesi, M. (2020). Production of xanthan gum using immobilized Xanthomonas
(2019). Effect of cold plasma treatment on Xanthan gum properties. Polymer Testing,
campestris cells: Effects of support type. Biochemical Engineering Journal, 157, Article
79, Article 106056.
107554.
Cox, W. P., & Merz, E. H. (1958). Correlation of dynamic and steady flow viscosities.
Niknezhad, S. V., Asadollahi, M. A., Zamani, A., & Biria, D. (2016). Production of
Journal of Polymer Science, 28(118), 619–622.
xanthan gum by free and immobilized cells of Xanthomonas campestris and
Cross, M. M. (1965). Rheology of non-Newtonian fluids: A new flow equation for
Xanthomonas pelargonii. International Journal of Biological Macromolecules, 82,
pseudoplastic systems. Journal of Colloid Science, 20(5), 417–437.
751–756.
Davies, N. W., Larkman, T., Marriott, P. J., & Khan, I. A. (2016). Determination of
Niknezhad, S. V., Asadollahi, M. A., Zamani, A., Biria, D., & Doostmohammadi, M.
enantiomeric distribution of terpenes for quality assessment of Australian tea tree
(2015). Optimization of xanthan gum production using cheese whey and response
oil. Journal of Agricultural and Food Chemistry, 64(23), 4817–4819.
surface methodology. Food Science and Biotechnology, 24(2), 453–460.
DuBois, M., Gilles, K. A., Hamilton, J. K., Rebers, P. A., & Smith, F. (1956). Colorimetric
Palaniraj, A., & Jayaraman, V. (2011). Production, recovery and applications of xanthan
method for determination of sugars and related substances. Analytical Chemistry, 28
gum by Xanthomonas campestris. Journal of Food Engineering, 106(1), 1–12.
(3), 350–356.
Palmqvist, E., & Hahn-Hägerdal, B. (2000). Fermentation of lignocellulosic hydrolysates.
Du, F., Qi, Y., Huang, H., Wang, P., Xu, X., & Yang, Z. (2022). Stabilization of O/W
II: Inhibitors and mechanisms of inhibition. Bioresource Technology, 74(1), 25–33.
emulsions via interfacial protein concentrating induced by thermodynamic
Pelletier, E., Viebke, C., Meadows, J., & Williams, P. A. (2001). A rheological study of the
incompatibility between sarcoplasmic proteins and xanthan gum. Food Hydrocolloids,
order–disorder conformational transition of xanthan gum. Biopolymers, 59(5),
124, Article 107242.
339–346.
Endo, R., Setoyama, M., Yamamoto, K., & Kadokawa, J.-i. (2015). Acetylation of xanthan
Pereira, T. S., de Sant’Anna, J. R., Silva, E. L., Pinheiro, A. L., & de Castro-Prado, M. A. A.
gum in ionic liquid. Journal of Polymers and the Environment, 23(2), 199–205.
(2014). In vitro genotoxicity of Melaleuca alternifolia essential oil in human
Eren, N. M., Santos, P. H. S., & Campanella, O. (2015). Mechanically modified xanthan
lymphocytes. Journal of Ethnopharmacology, 151(2), 852–857.
gum: Rheology and polydispersity aspects. Carbohydrate Polymers, 134, 475–484.
Pinto, E. P., Furlan, L., & Vendruscolo, C. T. (2011). Chemical deacetylation natural
Falleh, H., Ben Jemaa, M., Saada, M., & Ksouri, R. (2020). Essential oils: A promising eco-
xanthan (Jungbunzlauer®). Polímeros, 21(1), 47–52.
friendly food preservative. Food Chemistry, 330, Article 127268.
Riaz, T., Iqbal, M. W., Jiang, B., & Chen, J. (2021). A review of the enzymatic, physical,
Faria, S., de Oliveira Petkowicz, C. L., de Morais, S. A. L., Terrones, M. G. H., de
and chemical modification techniques of xanthan gum. International Journal of
Resende, M. M., de França, F. P., et al. (2011). Characterization of xanthan gum
Biological Macromolecules, 186, 472–489.
produced from sugar cane broth. Carbohydrate Polymers, 86(2), 469–476.
Ross-Murphy, S. B. (1995). Structure–property relationships in food biopolymer gels and
He, X., Dai, T., Sun, J., Liang, R., Liu, W., Chen, M., et al. (2022). Effective change on
solutions. Journal of Rheology, 39(6), 1451–1463.
rheology and structure properties of xanthan gum by industry-scale
Saha, A., & Basak, B. B. (2020). Scope of value addition and utilization of residual
microfluidization treatment. Food Hydrocolloids, 124, Article 107319.
biomass from medicinal and aromatic plants. Industrial Crops and Products, 145,
Huang, J., Zhong, C., & Yang, Y. (2020). Aggregating thermodynamic behavior of
Article 111979.
amphiphilic modified Xanthan gum in aqueous solution and oil-flooding properties
Sandford, P. A., Cottrell, I. W., & Pettitt, D. J. (1984). Microbial polysaccharides: New
for enhanced oil recovery. Chemical Engineering Science, 216, Article 115476.
products and their commercial applications. Pure and Applied Chemistry, 56(7),
Jang, H. Y., Zhang, K., Chon, B. H., & Choi, H. J. (2015). Enhanced oil recovery
879–892.
performance and viscosity characteristics of polysaccharide xanthan gum solution.
Sara, H., Yahoum, M. M., Lefnaoui, S., Abdelkader, H., & Moulai-Mostefa, N. (2020).
Journal of Industrial and Engineering Chemistry, 21, 741–745.
New alkylated xanthan gum as amphiphilic derivatives: Synthesis, physicochemical
Ju, J., Chen, X., Xie, Y., Yu, H., Guo, Y., Cheng, Y., et al. (2019). Application of essential
and rheological studies. Journal of Molecular Structure, 1207, Article 127768.
oil as a sustained release preparation in food packaging. Trends in Food Science &
Technology, 92, 22–32.

10
Z.-X. Li et al. Food Hydrocolloids 132 (2022) 107838

Shatwell, K. P., Sutherland, I. W., Ross-Murphy, S. B., & Dea, I. C. M. (1990). Influence of Wang, L., Xiang, D., Li, C., Zhang, W., & Bai, X. (2021). Effects of deacetylation on
the acetyl substituent on the interaction of xanthan with plant polysaccharides — II. properties and conformation of xanthan gum. Journal of Molecular Liquids, 117009.
Xanthan-guar gum systems. Carbohydrate Polymers, 14(2), 115–130. Wang, L., Zhu, F., Yang, Q., & Lu, D. (2013). Rheological properties of modified xanthan
Shepherd, M., Wood, R., Raymond, C., Ablett, G., & Rose, T. (2015). Ecotype variation in and their influence on printing performances on cotton with reactive dyes in screen
early growth, coppicing, and shoot architecture of tea tree (Melaleuca alternifolia). printing. Cellulose, 20(4), 2125–2135.
Industrial Crops and Products, 76, 844–856. Wei, Y., Guo, Y., Li, R., Ma, A., & Zhang, H. (2021). Rheological characterization of
Smith, I. H. (1984). Precipitation of xanthan gum. EP. Patent 0068706B1. polysaccharide thickeners oriented for dysphagia management: Carboxymethylated
Sujithra, B., Deepika, S., Akshaya, K., & Ponnusami, V. (2019). Production and curdlan, konjac glucomannan and their mixtures compared to xanthan gum. Food
optimization of xanthan gum from three-step sequential enzyme treated cassava Hydrocolloids, 110, Article 106198.
bagasse hydrolysate. Biocatalysis and Agricultural Biotechnology, 21, Article 101294. Wu, M., Shi, Z., Ming, Y., Wang, C., Qiu, X., Li, G., et al. (2021). Thermostable and
Tait, M. I., Sutherland, I. W., & Clarke-Sturman, A. J. (1990). Acid hydrolysis and high- rheological properties of natural and genetically engineered xanthan gums in
performance liquid chromatography of xanthan. Carbohydrate Polymers, 13(2), different solutions at high temperature. International Journal of Biological
133–148. Macromolecules, 182, 1208–1217.
Tariq, S., Wani, S., Rasool, W., Shafi, K., Bhat, M. A., Prabhakar, A., et al. (2019). Xu, L., Xu, G., Liu, T., Chen, Y., & Gong, H. (2013). The comparison of rheological
A comprehensive review of the antibacterial, antifungal and antiviral potential of properties of aqueous welan gum and xanthan gum solutions. Carbohydrate Polymers,
essential oils and their chemical constituents against drug-resistant microbial 92(1), 516–522.
pathogens. Microbial Pathogenesis, 134, Article 103580. Yahoum, M. M., Moulai-Mostefa, N., & Le Cerf, D. (2016). Synthesis, physicochemical,
Tian, M., Fang, B., Jin, L., Lu, Y., Qiu, X., Jin, H., et al. (2015). Rheological and drag structural and rheological characterizations of carboxymethyl xanthan derivatives.
reduction properties of hydroxypropyl xanthan gum solutions. Chinese Journal of Carbohydrate Polymers, 154, 267–275.
Chemical Engineering, 23(9), 1440–1446. Yoo, S. D., & Harcum, S. W. (1999). Xanthan gum production from waste sugar beet pulp.
Towle, G. A. (1977). Xanthan recovery process. U.S. Patent, Article 4051317. Bioresource Technology, 70(1), 105–109.
Trygg, J., Beltrame, G., & Yang, B. (2019). Rupturing fungal cell walls for higher yield of Zhang, Z., & Chen, H. (2010). Fermentation performance and structure characteristics of
polysaccharides: Acid treatment of the basidiomycete prior to extraction. Innovative xanthan produced by Xanthomonas campestris with a glucose/xylose mixture.
Food Science & Emerging Technologies, 57, Article 102206. Applied Biochemistry and Biotechnology, 160(6), 1653–1663.
Wang, Z., Wu, J., Zhu, L., & Zhan, X. (2016). Activation of glycerol metabolism in Zhang, N., & Yao, L. (2019). Anxiolytic effect of essential oils and their constituents: A
Xanthomonas campestris by adaptive evolution to produce a high-transparency and review. Journal of Agricultural and Food Chemistry, 67(50), 13790–13808.
low-viscosity xanthan gum from glycerol. Bioresource Technology, 211, 390–397. Zhou, Z., Liu, D., & Zhao, X. (2021). Conversion of lignocellulose to biofuels and
Wang, Z., Wu, J., Zhu, L., & Zhan, X. (2017). Characterization of xanthan gum produced chemicals via sugar platform: An updated review on chemistry and mechanisms of
from glycerol by a mutant strain Xanthomonas campestris CCTCC M2015714. acid hydrolysis of lignocellulose. Renewable and Sustainable Energy Reviews, 146,
Carbohydrate Polymers, 157, 521–526. Article 111169.

11

You might also like