You are on page 1of 14

Industrial Crops & Products 178 (2022) 114591

Contents lists available at ScienceDirect

Industrial Crops & Products


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

Bioprocess development for production of xylooligosaccharides prebiotics


from sugarcane bagasse with high bioactivity potential
Mahak Gupta a, Ridhika Bangotra a, Surbhi Sharma a, Surbhi Vaid a, Nisha Kapoor a,
Harish Chander Dutt b, Bijender Kumar Bajaj a, *
a
School of Biotechnology, University of Jammu, Jammu, India
b
Department of Botany, University of Jammu, Jammu, India

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

Keywords: The xylooligosaccharides (XOS, prebiotics) production from xylan extracted from sugarcane bagasse (SB)
Aspergillus flavus MG-7 biomass was accomplished using an in-house produced xylosidase-free endoxylanase from an isolate Aspergillus
L. plantarum M-13 flavus MG-7. Delignification of SB by sodium hypochlorite, and xylan extraction by coupled sodium hydroxide
Prebiotics
and thermal treatment were optimized for various process variables based on Design of Experiments (DoE) to
Sugarcane bagasse
achieve enhanced xylan yield (96.11%, w/w). The extracted xylan was analysed by Fouriertransform infrared
Xylooligosaccharides
spectroscopy and thermogravimetric analysis to evaluate its structural morphology, purity and thermal stability.
Xylan hydrolysate obtained by endoxylanase was quantified for xylobiose (69.57%, w/w), xylotriose (13.17%)
and xylotetrose (8.377%), and XOS were examined by 1H NMR. Growth-promoting potential of XOS was
established for a probiotic Lactobacillus plantarum M-13 isolated from kalarei. XOS exhibited excellent bioactivity
potential in terms of high positive prebiotic activity score, significant antioxidant activity and antibacterial
activity. Efficient XOS production promises value-addition for biorefining of SB biomass.

1. Introduction 2021; Forsan et al., 2021).


Prebiotics are the non-digestible but fermentable oligosaccharides
Lignocellulosic biomass (LB) represented by agroindustrial/forestry which may enhance the growth, vigour, and metabolic potential of some
residues is an inexpensive, plentiful and renewable resource that can be beneficial bacteria in the gastrointestinal tract (GIT) called as ‘pro­
exploited for production of energy, biofuels, chemicals and materials of biotics’ (Holscher, 2017). Prebiotics must be resistant to acidic gut
high commercial value (Khaire et al., 2021; Meghana and Shastri, 2020). conditions and digestive enzymes so that these reach the colon, and get
However, hemicellulose, especially the xylan, major hemicellulosic fermented by the resident bacteria (Ghosh et al., 2021). Recently, there
fraction of LB, is underutilized during bioconversion process, and thus, has been tremendous research interest in gut microbiota composition
undermines the overall process. Therefore, new/novel technologies are and health, and favourable interventions by probiotics and prebiotics
being investigated for complete utilization of LB-polysaccharides for mainly due to their multifaceted health advantages in various systemic
successful biorefining (Alvarez et al., 2017). Sugarcane is one of the disorders such as gastro-intestinal, cardiovascular, neurological, in­
largest conventional crops grown in various parts of the world including flammatory, oncological and endocrine systems (Oniszczuk et al., 2021).
India (Yang et al., 2021). The sugarcane bagasse (SB) left after the Beneficial gut microbiota (probiotics) in the host’s colon selectively
extraction of sugarcane juice is a very rich source of carbohydrates utilizes the prebiotics as substrates, and increases their numbers. Thus,
(40–45% of cellulose, 20–35% of hemicelluloses), and may be used for prebiotics enhance the growth and number of beneficial gut bacteria
the production of various commercial commodities (Meghana and which not only exclude various pathogenic bacteria but confer a variety
Shastri, 2020; Sharma et al., 2021a). Furthermore, extraction of xylan, of other health benefits. Furthermore, the short chain fatty acids (SCFAs)
and its utilization for the production of various high value industrial generated due to prebiotics fermentation possesses immunomodulatory
products such as xylitol, xylose and xylooligosaccharides (XOS) pre­ properties, and enhances the gut barrier integrity (Oniszczuk et al.,
biotics promises value-addition in the biorefining of SB (de Freitas et al., 2021). Various commonly used prebiotics include inulin,

* Corresponding author.
E-mail address: bajajbijenderk@gmail.com (B.K. Bajaj).

https://doi.org/10.1016/j.indcrop.2022.114591
Received 30 October 2021; Received in revised form 1 January 2022; Accepted 18 January 2022
Available online 29 January 2022
0926-6690/© 2022 Elsevier B.V. All rights reserved.
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

fructooligosaccharides (FOS), galactooligosaccharides, lactulose, poly­ A probiotic bacterial isolate L. plantarum M-13 (LPM-13) from kalarie
dextrose, isomaltooligosaccharides, xylooligosaccharides, lactitol, and (Gupta and Bajaj, 2017) that possessed several probiotic attributes
others (Guarino et al., 2020). (Gupta and Bajaj, 2018, 2017) was used for analysis of growth pro­
Xylooligosaccharides (XOS), the oligomers of β-1, 4-linked xylose moting ability of XOS. LPM-13 was preserved in glycerol at –20 ◦ C,
with numerous substitutions have recently emerged as a potential class revived in De Man, Rogosa and Sharpe (MRS) broth (at 37◦ C, 180 rpm,
of prebiotics (Yang and Xu, 2018). XOS are produced by the hydrolysis 18 h) and used for various experiments. The fungal isolate MG-7, an
of xylan, and may differ with respect to degree of polymerization (DP) i. isolate from Himalayan Shivalik foothills, was procured from Fermen­
e. the occurrence of monomeric units, degree of substitution (ratio of tation Biotechnology Laboratory, School of Biotechnology, University of
arabinose to xylose), and the linkages between them (Khaleghipour Jammu, Jammu. The fungal strain MG-7 was capable of producing
et al., 2021). Normally, DP of XOS in the range of 2–10 is preferable. Due xylosidase-free endoxylanase. The isolate MG-7 was identified as
to the presence of β-1, 4-xylosidic bonds, the XOS are resistant to gastric Aspergillus flavus based on culture morphology, microscopic, and
enzymes of the host’s gut, and are successfully fermented by beneficial sequence analysis of PCR-amplified internal transcribed spacer (ITS)
gut bacteria like Bifidobacterium species (B. adolescentis, B. infantis, B. region (Sharma et al., 2021b). The ITS sequence showed high resem­
longum, B. bifidum, and others), and several Lactobacillus species which blance (99%) with that of Aspergillus flavus strains available at National
result in increased growth and number of these bacteria (Holscher, Centre for Biotechnology Information (NCBI) database, thus, the isolate
2017). XOS prebiotics have been reported to promote the selective was identified and designated as A. flavus MG-7. The ITS sequence was
growth of a variety of other probiotics (Zhang et al., 2018). Further­ submitted to GenBank (https://www.ncbi.nlm.nih.gov/nuccor­
more, a wide bioactivity spectrum of XOS prebiotics viz. immunomod­ e/MG736928) with an accession number MG736928. Xylosidase-free
ulatory, anticancerous, antimicrobial, growth regulatory, antioxidant, endoxylanase was in-house produced from A. flavus MG-7 under sub­
antimicrobial and other bioactive properties promote the host’s health merged fermentation, and used for XOS production from xylan extracted
in multifaceted ways (Avila et al., 2020; Khaleghipour et al., 2021). from SB.
Additional health promoting benefits of XOS include gastrointestinal The bacterial pathogens (Staphylococcus aureus, Escherichia coli,
function improvement, blood cholesterol lowering, reducing stomach Micrococcus luteus, Enterococcus faecalis, Bacillus cereus, Pseudomonas
lesions, and others, thus, indicating a huge application potential of XOS aeruginosa, Streptococcus pneumoniae, Pseudomonas alcaligenes, Bacillus
for food and pharmaceutical industries (Zhang et al., 2018a). subtilis and Klebsiella sp.) were procured from the culture repository of
The enzymatic production of XOS is preferred due to several ad­ the Fermentation Biotechnology Laboratory of the School of Biotech­
vantages such as no or low production of undesirable by-products, easy nology, University of Jammu. All the media, chemicals and reagents
recovery of pure XOS, and the environmental safety of the process (Avila used were of high analytical grade.
et al., 2020). Endoxylanases produce XOS from xylan, however,
β-xylosidase degrades the XOS to produce xylose. Therefore, endox­ 2.2. Polysaccharide composition of biomass
ylanases intended for XOS production must be β-xylosidase-free (Surek
and Buyukkileci, 2017). Owing to numerous applications of XOS, in­ The polysaccharide composition (%, w/w) of raw and delignified SB
dustries are attempting to develop new technologies for production of was determined (Sluiter et al., 2008) by HPLC (Shimadzu RID-20A)
XOS with high purity, and with a desired degree of polymerization. using Aminex HPX 87 H column (temperature 80ºC) using water as
However, the production of XOS still remains an expensive and the mobile phase. The samples were run at an isocratic mode with a flow
complicated process despite high market value and demand (Forsan rate of 0.6 ml/min and an end time of 20 min. The quantitative analysis
et al., 2021). of various monosaccharides (glucose, xylose, mannose and arabinose)
Considering the importance of XOS for food and pharmaceutical was done using appropriate reference standards.
industries, the present study aimed at producing XOS from sugarcane
bagasse xylan by employing an in-house produced process-apt xylosi­ 2.3. Pretreatment of sugarcane bagasse for delignification
dase-free endoxylanase from an indigenously isolated Aspergillus flavus
MG-7. Design of experiments (DoE) based optimization of delignifica­ Pretreatment of sugarcane bagasse was executed by four different
tion, and xylan extraction from sugarcane bagasse is being reported for delignifying reagents i.e. sodium hypochlorite, sodium chlorite, aqueous
the first time to obtain high xylan yield. Furthermore, the enzymatic ammonia, and combination of hydrogen peroxide and acetic acid
hydrolysis of xylan using A. flavus MG-7 endoxylanase was optimized for (H2O2–Hac) at a biomass loading of 1:10 (%, w/v) under different set of
production of XOS. Also, the XOS were investigated for desired func­ conditions (Reddy and Krishnan, 2016). Reaction conditions (reagent
tional attributes with respect to their growth promoting effect on a concentration, temperature, time) for delignification were: sodium hy­
probiotic L. plantarum M-13 (LPM-13), prebiotic activity score, antiox­ pochlorite (1% w/v, 40ºC, 1 h), sodium chlorite (1% w/v, 75ºC, 2 h),
idant, and antibacterial activity. LPM-13 was previously isolated from aqueous ammonia (15% w/v, 40ºC, 16 h), H2O2–Hac (1:1 hydrogen
an indigenously fermented milk product kalarei, and characterized for peroxide 21.6 M and glacial acetic acid 8.74 M, 60ºC, 7 h). After
functional probiotic attributes, and used for development of fermented delignification, the reaction contents were vacuum filtered and lignin
foods (Gupta and Bajaj, 2018, 2017). Additionally, the extracted xylan, rich filtrate was examined for sugar loss (Dubios et al., 1956). The
and the XOS were studied by nuclear magnetic resonance (1H NMR), sample was appropriately diluted with distilled water, and used for
Fourier transform infrared spectroscopy (FTIR) and thermogravimetric sugar assay. To 1.0 ml of diluted sample, 1.0 ml of phenol (5% v/v) and
analysis (TGA) to elucidate their structural and functional 5.0 ml of sulphuric acid (96% v/v) was added, and reaction content was
characteristics. incubated at room temperature for 20 min with intermittent shaking.
Sugar content was examined spectrophotometrically (A490) using xylose
2. Material and methods standard and expressed as mg/g. The distilled water was used as blank.
The most efficient delignification reagent (sodium hypochlorite) which
2.1. Lignocellulosic biomass, process organisms, chemicals and reagents incurred minimum sugar loss was selected, and the delignification
process was optimized for various process parameters.
Sugarcane bagasse (SB) was collected from a local sugarcane juice
vendor of Jammu, India. The biomass was finely chopped, washed with 2.4. Optimization of delignification of sugarcane bagasse by sodium
hot distilled water and dried in oven at 60◦ C for 24 h. SB was grounded, hypochlorite
sieved to a uniform size (<4 mm) and stored in an air tight container at
room temperature. Central composite design (CCD) of response surface methodology

2
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

(Design-Expert 6.0 software, Stat Ease, Inc., Minneapolis, Minnesota, enzyme preparation so obtained was lyophilized. For ascertaining the
USA) was applied for optimization of sodium hypochlorite based xylosidase-free mode of action of the xylanase, enzymatic hydrolysis of
delignification of SB. The four process variables at minimum and the standard xylan (beech wood xylan, 0.5% w/v) was carried out by
maximum level (indicated in parentheses) were: biomass loading using xylanase at different loadings (20 − 100 IU) at 50ºC under shaking
(5–10%, w/v), temperature (40–60◦ C), time (60–90 min) and sodium (150 rpm, 16 h). The endoxylanolytic nature of the xylanase was
hypochlorite concentration (0.50–1%, v/v). Thirty experiments devel­ examined (Bajaj et al., 2011) by thin layer chromatography (TLC) on
oped by DoE were performed, and the results were examined by analysis pre-coated silica gel plates (Macherey-Nagel, Alugram extra SIL
of variance (ANOVA). The significance of process variables, and the G/UV254).
model was determined (p < 0.05). Three dimensional response surface
plots generated by the model were used to determine the influence of
variables on the response (xylan yield %, w/w) individually and 2.8. Enzymatic production of xylooligosaccharides using A. flavus MG-7
interactively. endoxylanase
dry weight of extracted xylan (g)
Xylan (%, w/w) = × 100 Enzymatic production of XOS from xylan extracted from SB was
weight of the sample (g)
accomplished using in-house generated xylosidase-free endoxylanase
The validation of the process model was done based on the point from fungal isolate A. flavus MG-7. Optimization of various process
prediction tool of the software. The experiments were performed at variables was done in a sequential manner using one variable at a time
optimal level of variables predicted by the software, and the experi­ (OVAT) approach. Enzymatic hydrolysis was carried out at 50ºC with
mental and predicted responses were analysed. mild shaking (50 rpm). The reaction without enzyme was set as a con­
trol. Effect of enzyme dosage (xylanase loading 5 IU to 120 IU) on XOS
2.5. Optimization of xylan extraction from delignified sugarcane bagasse production was studied using xylan at 0.5% (w/v) as the substrate.
Similarly, the effect of xylan concentration (0.5–3.0%, w/v) at a selected
Three different approaches i.e. alkali hydrolysis, acid hydrolysis, and ‘xylanase dosage’, was examined, and the effect of reaction time (16, 24
coupled alkali (sodium hydroxide) and thermal treatment (121 ºC for and 48 h) was determined using selected optimal values of ‘xylanase
1 h), were examined for xylan extraction from delignified SB at biomass loading’ and ‘xylan concentration’. The total reducing sugar content
loading of 1:10 (%, w/v) (Chapla et al., 2012). The coupled alkali (so­ (expressed as xylose equivalent mg/g) was examined for different ex­
dium hydroxide) and thermal treatment which yielded the maximum periments, and the reaction conditions which yielded maximum xylose
xylan (%, w/w), was further optimized by DoE for four process variables equivalent were considered as the optimal ones for enzymatic hydrolysis
(minimum and maximum level in parentheses) i.e. extraction tempera­ of xylan.
ture (65–100ºC), biomass loading (5–10%, w/v), extraction time The qualitative analysis of XOS was executed by TLC (Bajaj et al.,
(10–20 min) and sodium hydroxide concentration (10–14%, v/v). 2011) while the quantitative analysis (XOS %, w/w) was done by HPLC
Thirty experiments generated by the software were performed, and the under operating conditions described in Section 2.2, at a column tem­
results were examined by ANOVA. The significance of the variables and perature of 40ºC. Xylose and xylooligosaccharides (xylobiose X2, xylo­
the model was ascertained. The individual and the interactive influences triose X3 and xylotetrose X4) were used as the reference standards
of the process variables on xylan yield were studied by the 3-D response (Megazyme, Prolab). The XOS hydrolysate was lyophilized and stored at
surface plots and perturbation plot. The optimum level of the variables –20ºC for further use.
predicted by the point prediction tool, was used to execute the experi­
ments, and the observed (experimental), and predicted responses were
analysed, to validate the process model. 2.9. Nuclear magnetic resonance analysis of xylooligosaccharides

2.6. Structural characterization of the xylan extracted from sugarcane For obtaining the NMR spectrum (Bruker Avance III, 400 MHz), a
bagasse biomass 10 mg of lyophilized XOS was dissolved in 0.5 ml of D2O at the following
operating conditions i.e. 25ºC, 125 scans, proton 90 R at 10.65 ms, and
Xylan extracted from the delignified SB was characterized by Fourier the recycle delay at 3 s. NMR analysis was done at Sophisticated Test
transform infrared (FTIR) spectroscopy, and thermogravimetric analysis and Instrumentation Centre (STIC), Cochin University of Science and
(TGA) (Yiin et al., 2017; Li et al., 2016). FTIR spectrum of the extracted Technology (CUSAT), Kerala, India.
xylan was obtained using Thermo Nicolet, Avatar 370 spectrophotom­
eter ( 4000–400 cm− 1, resolution 4 cm− 1, KBr beam splitter). TGA of
xylan was done by heating the sample (20.0 mg) at 50–750 ◦ C with 2.10. Prebiotic potential of xylooligosaccharides for growth of L.
10 ◦ C increase/min in continuous nitrogen gas flow (50 ml/min) (Perkin plantarum M-13
Elmer, Diamond TG/DTA). Furthermore, the surface topography of the
residual SB biomass after delignification and xylan extraction, was Effect of XOS on the selected probiotic bacterium LPM-13 was
examined by scanning electron microscopy, SEM at 1000X to 3000X determined by using modified MRS (% w/v, proteose peptone 1.0, beef
(JEOL Model JSM - 6390LV instrument, accelerating voltage 30 kV) extract 1.0, yeast extract 0.5, polysorbate 0.1, ammonium citrate 0.2,
(Sharma et al., 2021a). All the biomass samples were analysed at So­ sodium acetate 0.5, magnesium sulphate 0.01, manganese 0.005,
phisticated Test and Instrumentation Centre (STIC), Cochin University dipotassium phosphate 0.2, pH 6.5 ± 0.2) in which the normal carbon
of Science and Technology (CUSAT), Kerala, India. source (glucose) was replaced either with XOS or with any of the com­
mercial prebiotics (inulin, fructooligosaccharide FOS or lactulose), as
2.7. Xylosidase-free endoxylanase from fungal isolate Aspergillus flavus the sole carbon source. MRS with glucose was used as the standard for
MG-7 the growth analysis of LPM-13 (Noori et al., 2017).
The filter sterilized prebiotic or glucose solution was added to the
The xylanolytic activity of the A. flavus MG-7 was ascertained by pre-autoclaved medium broth (1%, w/v), and inoculated with the pro­
qualitative and quantitative assay (Bajaj et al., 2011). Xylanase pro­ biotic LPM-13 (1% v/v, 108–109 cells/ml) and incubated (37◦ C,
duced under submerged fermentation was partially purified and 180 rpm, 144 h). Samples withdrawn at different time intervals were
concentrated by ammonium sulphate fractionation (60–80%), followed examined for growth of LPM-13 (log cfu/ml) based on viable plate count
by dialysis for 8 h with intermittent change of buffer after every 4 h. The on MRS agar.

3
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

2.11. Stability of xylooligosaccharides towards human gastric juice supernatant, and 50 µl of normal saline. The inhibition in the growth of
the pathogen by XOS was calculated using the following equation:
The stability analysis of XOS in artificial human gastric juice was
A1 – A0 = ACal
evaluated using FOS as the control (Wang et al., 2015). Artificial human
gastric juice was prepared (% w/v, Na2HPO4⋅H2O 0.825; NaH2PO4 where A1 is the absorbance of sample; A0 is the absorbance of the
1.435; NaCl 0.8; KCl 0.02; CaCl2.2 H2O 0.01 and MgCl2.6 H2O 0.018, in blank, and ACal is the absorbance calculated for each pathogen.
distilled water), and set at different pH (1− 5) with 1.0 M HCl. XOS or The inhibition of the pathogen by XOS was calculated as follows:
FOS were suspended (1%, w/v) in artificial gastric juice, and incubated
at 37ºC for 6 h. Samples withdrawn at different time intervals were Inhibition(%) =
ACal
× 100
examined for the total and reducing sugar content (Sharma et al., 2021a, ANB
2021b). The stability of xylooligosaccharides (degree of hydrolysis) was
where ANB is absorbance of each pathogen suspension grown in NB.
calculated according to the formula (Wang et al., 2015):

Hydrolysis degree (%) =


Reducing sugar released
× 100 2.14. Antioxidant activity of xylooligosaccharides
Initial total sugar − initial reducing sugar
Antioxidant activity of XOS was analyzed spectrophotometrically
where reducing sugar released was the difference between the reducing (A517) based on 2, 2-diphenyl-1-picrylhydrazyl (DPPH) radical scav­
sugar content at specified time and the reducing sugar content present at enging activity (Avila et al., 2020). XOS-SB was used at different con­
0 h. centrations (0.5–5.0 mg/ml in acetate buffer 50 mM, pH 6), and DPPH
radical scavenging activity was determined using the following
2.12. Prebiotic activity score of xylooligosaccharides equation:
DPPH scavenging activity (%) = 1 − absorbance of sample
absorbance of control × 100
Prebiotic activity score of XOS was examined (Zhang et al., 2018)
using probiotic LPM-13, and E. coli or S. aureus as pathogenic bacteria. 3. Results and discussion
Activated culture of LPM-13 or bacterial pathogen was inoculated (1%
v/v, 108-109 cells/ml) into modified MRS broth containing XOS-SB or 3.1. Polysaccharide composition of sugarcane bagasse biomass
glucose (1%, w/v), as the sole carbon source, and incubated at 37 ◦ C for
48 h under shaking (180 rpm). The growth of bacterial cultural was The holocellulose content of SB biomass was observed to be 57.87%
examined by viable plate count, and expressed as log cfu/ml. The pre­ (w/w) that included cellulose (38.25%, w/w) and hemicelluloses
biotic activity score of XOS-SB was determined using the following (19.62%, w/w), and three principle monomeric sugars i.e. glucose
equation (Zhang et al., 2018): (42.51% w/w), xylose (15.05% w/w) and arabinose (7.26% w/w).
ProbX48 − ProbX0 PathX48 − PathX0 Sharma et al. (2019) reported holocellulosic content of 58.6% in raw
Prebiotic activity score = −
ProbG48 − ProbG0 PathG48 − PathG0 sugarcane bagasse which represented cellulose and hemicellulose frac­
tions of 38.6% and 20.0%, respectively. Similarly, a detailed analysis of
where, ProbX48 and ProbX0 indicate the growth of probiotic LPM-13 sugarcane bagasse (Hesam et al., 2020) showed the following constitu­
in XOS medium at 48 h, and 0 h, respectively; PathX48 and PathX0 ents (%, w/w): cellulose 33.36, hemicelluloses 25.01, lignin 23.27, ash
represent the growth of bacterial pathogen in XOS medium at 48 h, and 2.91, and crude protein 1.16. Composition of sugarcane bagasse biomass
0 h, respectively; ProbG48 and ProbG0 indicate the growth of probiotic may vary due to diversity of sugarcane varieties, climatic and cultivation
LPM-13 in glucose medium at 48 h, and 0 h, respectively; PathG48 and conditions, and other local factors (Sharma et al., 2019).
PathG0 indicate growth of bacterial pathogen in glucose medium at 48 h,
and 0 h, respectively.
3.2. Delignification of sugarcane bagasse biomass

2.13. Impact of xylooligosaccharides on antibacterial activity of Delignification of SB aimed to remove lignin while maintaining the
L. plantarum M-13 integrity of carbohydrates in the substrate. Among the four pretreatment
methods, minimum sugar loss (0.75 ± 0.04 mg/g) was realized by so­
LPM-13 was inoculated (1% v/v, 108-109 cells/ml) into modified dium hypochlorite based pretreatment as compared to other pre­
MRS broth containing XOS (1%, w/v) as the sole carbon source, and treatments i.e. liquid ammonia (0.82 ± 0.02 mg/g), hydrogen peroxide-
incubated at 37 ◦ C for 18 h under shaking (180 rpm). The cultural broth acetic acid (H2O2–Hac) (0.95 ± 0.09 mg/g) and sodium chlorite
was centrifuged (5000g for 20 min), and the supernatant was filter- (1.18 ± 0.05 mg/g). Therefore, sodium hypochlorite pretreatment was
sterilized (0.22 µm filter), and used for antimicrobial activity assay. selected for further studies. Sodium hypochlorite is considered as a
The pathogenic bacteria used in the study were Staphylococcus aureus, potent delignifying agent as it creates adequate alkaline conditions, and
Escherichia coli, Micrococcus luteus, Enterococcus faecalis, Bacillus cereus, results in maximum lignin removal. Under alkaline pretreatment,
Pseudomonas aeruginosa, Streptococcus pneumoniae, Pseudomonas alcali­ different chemical bonds in lignin may show different type of reactions.
genes, Bacillus subtilis and Klebsiella sp. Either of the bacterial pathogen For instance, the linkages between phenolic α-aryl ether bonds and
was grown in nutrient broth (NB) under shaking (180 rpm) for 18 h at α-alkyl ether bonds are very weak, and get fractured in alkaline milieu
37◦ C. The cell biomass was collected by centrifugation (5000g for resulting in considerable depolymerisation of lignin. However, the non-
10 min), washed and resuspended in sterile saline (108-109 cells/ml) phenolic α-aryl ether linkages are very stable, and require much harsh
(Millette et al., 2007). conditions for their breakage. Furthermore, after pretreatment of SB, a
For antibacterial activity, a mixture consisting of 100 µl of NB, 100 µl yield of 91.55%, w/w was obtained, and weight loss of 8.45%, w/w was
of filter-sterile supernatant and 50 µl of either of the pathogen suspen­ accounted mainly due to removal of lignin, and some other extractives.
sion was incubated at 37◦ C for 16 h under shaking (180 rpm). In the However, delignification also incurred a weight loss of cellulose (1.72%)
control experiment C1, 100 µl of filter sterile supernatant was replaced and hemicelluloses (3.55%) content.
with 100 µl of NB, while control experiment C2 consisted of 100 µl of Furthermore, after delignification, the holocellulose content of SB
NB, 100 µl of MRS and 50 µl of either of the pathogen suspension. The was enriched by 12.93% (w/w), and resulted in increased cellulose
microbial growth was monitored with a spectrophotometer (A600) (41.49%, w/w) and hemicellulose content (29.31%, w/w). Also, the
against a blank which consisted of 100 µl NB, 100 µl of filter sterile glucose, xylose and arabinose content were increased to 46.11%,

4
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

25.36% and 7.96%, respectively. This increase in sugar content was The different response in various experimental runs indicated the
mainly due to the removal of lignin and various other plant cell wall distinct impact of each of the process parameters on the response.
components. The lignin removal increases the porosity, and concentra­ Analysis of variance (ANOVA) showed that the model was significant
tion of carbohydrate polymers, and makes the biomass more amenable with F value of 2.84 and Prob>F value of 0.0269 (Table 2).
to bioconversion. Cellulose content of sugarcane bagasse biomass was In this case, A (temperature), D2 (sodium hypochlorite concentra­
reported to get increased from 42% to 55% after alkaline pretreatment tion), AB (interactive effect of temperature and biomass loading) and AC
(Nalawade et al., 2020). Similarly, cellulose content of sugarcane (interactive effect of temperature and time), were the significant model
bagasse biomass was increased considerably i.e. 52–79%, after auto­ terms. The lack of Fit F-value (0.87) implied its non-significance which
hydrolysis (hydothermal pretreatment) and alkaline delignification (de in turn substantiated the robustness of the model. The coefficient of
Oliveira Rodrigues et al., 2021). variation (CV) of 2.45 indicated the reliability and precision of the
The alkali (NaOH 1%, w/v) pretreatment of sugarcane bagasse experiment. Generally, a CV value of < 10% is desired to ensure credi­
resulted in increase of holocellulosic and cellulosic contents from 68.3% bility of the experimental data.
to 70.1% (w/w), and from 42.5% to 47.4% (w/w), respectively (Nath Coefficient of determination R2 (0.9261) explained the goodness of
et al., 2021). However, the original hemicellulosic (25.8%, w/w) con­ the model (value closer to 1.0 denotes better correlation between
tent was decreased slightly (22.1%, w/w), but acid-insoluble lignin observed and predicted responses), and indicated that 92.61% of the
(22.1% (w/w) content was decreased appreciably (12.3%, w/w) (Nath variability in the response could be explained by the model. Further­
et al., 2021). The raw SB biomass was reported to contain > 55% (w/w) more, the adjusted R2 (0.9705) and predicted R2 (0.9447) indicated
carbohydrate content (glucose 33.27%, xylose 21.44% and arabinose good agreement between experimental and predicted responses.
2.09%) of which 23% (w/w) constituted the hemicelluloses (Kundu Adequate precision i.e. signal to noise ratio of 26.204 (> 4.0) indicated
et al., 2021). Considerably high fraction of hemicelluloses (especially an adequate signal. All the statistical parameters were appropriate and
the xylan) in SB reflects its potential as an excellent substrate for XOS statistically significant. Thus, the model was used to navigate the design
production. Thus, loss of carbohydrate macromolecules during deligni­ space. A regression equation (I) obtained after multiple regression
fication is quite variable and depends on the type of delignifying analysis, for the response (Y, xylan, %, w/w) as a function of various
agents/methods used, and the type of biomass substrate being studied. variables is given below:
Response Y (Xylan, %, w/w) = +51.43 + 0.41 A− 3.80D2
+ 2.19AB− 8.48AC ………….(I).
3.3. Process optimization for delignification of sugarcane bagasse
The polynomial equation describes variations in the response Y
(xylan, % w/w) as a function of various variables i.e. A (delignification
For DoE-based optimization of process variables (biomass loading,
temperature), B (biomass loading), C (delignification time) and D (so­
temperature and time, and sodium hypochlorite concentration) for
dium hypochlorite concentration), and also provides detailed informa­
delignification of sugarcane bagasse, thirty designed experiments were
tion about the interactions within (i.e. A2, B2, C2 and D2), and between
performed, and the responses were fed to the response column (Table 1).
the independent variables (AB, AC, BC and BD). The interactive effect of
temperature (A) and biomass loading (B), and that of temperature (A)
Table 1
and time (C) on the response (xylan yield %, w/w) is presented in Fig. 1
Design of Experiment based optimization for delignification of sugarcane
(a) and (b), respectively. Three-dimensional (3-D) response surface plots
bagasse.
showed that the maximum response was obtained at moderate level of
Run Experimental variables* Xylan (%,w/w) temperature and maximum level of biomass loading and time. Pertur­
A B C D Experimental Predicted bation plot (Fig. 1c) showed an increase or decrease in the response
response response when each of the variable was changed keeping other variables constant
1 40.00 10.00 90.00 1.00 29.75 30.83 with respect to the chosen reference point. Sodium hypochlorite (D) had
2 50.00 7.50 75.00 0.75 41.53 40.30
3 50.00 7.50 45.00 0.75 30.66 43.38
4 60.00 5.00 90.00 1.00 31.00 61.61 Table 2
5 40.00 10.00 60.00 1.00 18.20 57.20 Analysis of variance for delignification of sugarcane bagasse based on Design of
6 50.00 7.50 75.00 0.75 43.93 32.76 Experiment.
7 40.00 10.00 90.00 0.50 66.50 59.29
Source* Sum of DF Mean F Prob
8 60.00 10.00 90.00 1.00 31.95 43.61
Squares Square Value >F
9 60.00 5.00 60.00 0.50 46.30 27.99
10 50.00 2.50 75.00 0.75 54.80 45.32 Model 3250.72 14 232.19 2.84 0.0269 Significant
11 40.00 5.00 90.00 1.00 65.80 21.95 A 4.11 1 4.11 0.050 0.0056 Significant
12 50.00 7.50 105.00 0.75 47.66 48.05 B 34.68 1 34.68 0.42 0.5246 –
13 40.00 5.00 60.00 0.50 33.10 55.85 C 145.78 1 145.78 1.78 0.2016 –
14 60.00 5.00 90.00 0.50 39.20 39.27 D 333.69 1 333.69 4.08 0.0616 –
15 60.00 5.00 60.00 1.00 40.80 39.35 A2 338.67 1 338.67 4.14 0.0599 –
16 50.00 7.50 75.00 0.75 47.40 31.53 B2 57.96 1 57.96 0.71 0.4130 –
17 40.00 5.00 60.00 1.00 27.40 36.55 C2 277.60 1 277.60 3.40 0.0852 –
18 50.00 7.50 75.00 1.25 32.26 38.20 D2 395.57 1 395.57 4.84 0.0039 Significant
19 50.00 12.50 75.00 0.75 60.60 54.84 AB 76.78 1 76.78 0.94 0.0078 Significant
20 50.00 7.50 75.00 0.25 41.13 59.65 AC 1150.06 1 1150.06 14.07 0.0019 Significant
21 30.00 7.50 75.00 0.75 37.06 33.78 AD 61.82 1 61.82 0.76 0.3982 –
22 60.00 10.00 60.00 0.50 54.35 43.64 BC 109.46 1 109.46 1.34 0.2653 –
23 50.00 7.50 75.00 0.75 55.60 43.70 BD 345.50 1 345.50 4.23 0.0576 –
24 50.00 7.50 75.00 0.75 67.53 28.78 CD 2.21 1 2.21 0.027 0.8715 –
25 40.00 5.00 90.00 0.50 45.10 51.43 Residual 1225.99 15 81.73 –
26 50.00 7.50 75.00 0.75 52.60 51.43 Lack of 777.52 10 77.75 0.87 0.6053 Not
27 60.00 10.00 90.00 0.50 40.60 51.43 Fit significant
28 70.00 7.50 75.00 0.75 38.60 51.43 Pure 448.47 5 89.69 –
29 60.00 10.00 60.00 1.00 56.55 51.43 Error
30 40.00 10.00 60.00 0.50 48.05 51.43 Cor Total 4476.71 29 –
*
(*A- Delignification temperature, ºC, B- biomass loading, %, w/v, C- delignifi­ (A- Delignification temperature, B- biomass loading, C- delignification time,
cation time, min and D- sodium hypochlorite, %, v/v) and D- sodium hypochlorite)

5
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

Fig. 1. Response surface plots for delignification of sugarcane bagasse showing (a) interaction between temperature and biomass loading; (b) different temperature
and time; (c) perturbation plot showing the effectiveness of the variables on the delignification of sugarcane bagasse.

the maximum effect on the response, and was followed by biomass ANOVA-demonstrated model F-value of 1.30 implied the signifi­
loading (B), delignification time (C) and temperature (A). cance of the model (Table 4). Based on Prob>F value, the model terms i.
The experiments were performed at predicted optimal level of pro­ e. B (biomass loading), D (sodium hydroxide concentration), C2
cess variables i.e. temperature 51.86ºC, biomass loading 12.91% w/v, (extraction time), AB (interaction between temperature and biomass
time 75.12 min, and sodium hypochlorite concentration 0.53% v/v), for loading), BC (interaction between biomass loading and extraction time)
validating the model. The close proximity of experimental (75.05% w/ and BD (interaction between biomass loading and sodium hydroxide
w, xylan) and predicted (72.65%, w/w, xylan) responses validated the concentration), were significant. Insignificant lack of fit (P-value
process model. 0.1608), low standard deviation (3.04), and high R2 (0.9261), pointed
Presence of lignin and highly crystalline cellulose hinders the frac­ towards robustness and a reasonably good predictability of the model.
tionation of biomass, and necessitates a suitable pretreatment regime for Adequate precision of 26.204 indicated a sufficient signal. The regres­
efficient delignification of biomass especially under mild process con­ sion equation (II) obtained for the model is given below:
ditions to avoid sugar degradation (Yiin et al., 2017). An efficacious Response, Y (Xylan, %, w/w) = +51.43 + 1.20B− 3.73D− 3.18 C2
pretreatment promises effective and economic bioconversion at low + 2.19AB+ 2.62 BCE+ 4.65BD……………………(II).
operating cost. Several factors influence the polysaccharide extraction The variables in coded form were: temperature (A), biomass loading
from LB in terms of yield and functional properties of extracted polymer, (B), extraction time (C) and sodium hydroxide concentration (D).
therefore, for obtaining high quality and yield, the process variables The 3-D response surface plots representing the interactive effect of
must be optimized (Wei et al., 2018). Optimization of delignification of process parameters were studied. The interactions namely AB (temper­
sugarcane bagasse (Valim et al., 2018) was achieved using alkaline ature and biomass loading), BC (biomass loading and time), and BD
hydrogen peroxide with high efficiency (78.9%). Effectiveness of (biomass loading and sodium hydroxide concentration) significantly
delignification was indicated by the favourable morphological changes affected the response. The interactive influences of temperature (A) and
in the biomass. biomass loading (B) (Fig. 2a), biomass loading (B) and extraction time
(C) (Fig. 2b), and that of biomass loading (B) and sodium hydroxide
3.4. Process optimization for xylan extraction from delignified sugarcane concentration (D) (Fig. 2c), showed a direct relationship with the
bagasse response. Thus, increasing the level of these process parameters led to
concomitant enhanced response (xylan yield). The perturbation plot
Among various approaches examined for xylan extraction from the analysis illustrated that among all the variables, the variable B (biomass
delignified SB, sodium hydroxide coupled with thermal treatment yiel­ loading) deviated maximally from the reference point, and was followed
ded the maximum xylan (41.49%). Therefore, various process variables by other variables i.e. A, C and D (Fig. 2d).
such as temperature (A), biomass loading (B), extraction time (C) and Experiments performed at predicted optimal level of variables
sodium hydroxide concentration (D), were optimized for xylan extrac­ (temperature 89.03ºC, biomass loading 11.80% w/v, time 20.14 min
tion using this approach. The responses obtained for DoE-generated and sodium hydroxide concentration 14.19% v/v) showed a close pro­
thirty experimental runs are presented in the Table 3. pinquity between the experimental (96.11%, w/w) and the predicted

6
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

Table 3 process variables i.e. H2O2 5.63%, NaOH 12.91%, pretreatment time
Experimental and predicted response for xylan extraction (%, w/w) from 17.51 h (Hesam et al., 2020). A combination of alkalis (NaOH and
delignified sugarcane bagasse based on Design of Experiment. NH4OH) was reported to exert synergistic action on xylan extraction
Run Experimental variablesnew-a Xylan % (w/w) from sugarcane bagasse, and further the statistical optimization of alkali
A B C D Experimental Predicted
concentration and temperature increased the xylan recovery (Kundu
response response et al., 2021).
1 82.50 7.50 15.00 8.00 44.60 41.11
2 65.00 5.00 10.00 14.00 55.90 55.33 3.5. Physicochemical characterization of xylan extracted from sugarcane
3 100.00 5.00 20.00 10.00 51.00 51.92 bagasse
4 82.50 7.50 25.00 12.00 52.33 53.41
5 47.50 7.50 15.00 12.00 40.00 38.32
3.5.1. Thermogravimetric analysis of xylan
6 65.00 5.00 10.00 10.00 43.90 44.89
7 65.00 10.00 10.00 10.00 40.60 40.77 Thermogravimetric analysis of xylan was done to examine its ther­
8 82.50 7.50 15.00 12.00 43.86 41.63 mal stability and degradation pattern, to access its application potential
9 82.50 7.50 15.00 12.00 62.80 65.82 keeping in view that most of the industrial processes generally involve
10 100.00 5.00 20.00 14.00 31.50 31.79 high temperature milieu. TGA curve (Fig. S1a) showed that xylan
11 65.00 10.00 20.00 14.00 52.85 53.63
degradation occurred in three stages. The initial degradation at less than
12 65.00 5.00 20.00 10.00 36.80 38.87
13 100.00 10.00 10.00 14.00 66.70 67.80 150ºC could be attributed to the loss of water. The water content in the
14 65.00 10.00 20.00 10.00 53.60 46.12 sample is responsible for the low dip in the initiation of the thermal
15 100.00 10.00 20.00 10.00 49.75 48.26 degradation (Yiin et al., 2017). At around 80ºC, the degradation of xylan
16 117.50 7.50 15.00 12.00 75.95 71.85
led to a weight loss of 10%. After initial degradation phase, weight loss
17 82.50 7.50 15.00 12.00 61.2 62.77
18 82.50 12.50 15.00 12.00 84.16 86.58 occurred slowly which could be due to the existence of thermally stable
19 82.50 7.50 15.00 16.00 64.00 63.49 sugar moieties. In the second stage (150–450ºC), the decomposition and
20 100.00 5.00 10.00 14.00 31.60 31.02 elimination of carboxylic/hydroxyl groups resulted in a perceptible
21 100.00 10.00 20.00 14.00 85.75 84.26 weight loss (Fig. S1a) that was maximum (40%) around 200–300ºC
22 82.50 7.50 15.00 12.00 56.60 54.44
(Zhao et al., 2017c). The third stage of mass loss was observed around
23 65.00 10.00 10.00 14.00 56.90 56.87
24 82.50 7.50 15.00 12.00 58.33 57.81 500–600ºC (Fig. S1a) in which pyrolysis was completed and only char
25 82.50 7.50 15.00 12.00 39.66 59.74 residue was left (31–35%). At this point, all the elemental oxygen and
26 100.00 10.00 10.00 10.00 33.65 33.74 hydrogen had been liberated to gaseous compounds leaving only carbon
27 82.50 7.50 5.00 12.00 63.60 62.80
elements (Zhao et al., 2017c).
28 65.00 5.00 20.00 14.00 50.60 53.74
29 100.00 5.00 10.00 10.00 30.40 34.40
Since xylan is less thermally stable than cellulose and lignin, it de­
30 82.50 2.50 15.00 12.00 64.60 65.20 composes around 225–325ºC (Davila et al., 2016). Xylan degradation
a was reported to occur in two-step i.e. in first step, degradation of the
(A- Extraction temperature, ºC; B- biomass loading, %, w/v; C- extraction
glycosidic bonds along with the side-chain substituting groups occurs,
time, min; D-sodium hydroxide, %, w/v)
and in the second step breakage of the depolymerised fragments takes
place. The degradation and weight loss in this phase corresponds to
Table 4 random polymer backbone cleavages and depolymerization of the
Analysis of variance for xylan extraction (%, w/w) from delignified sugarcane remaining inorganic material from the xylan (Davila et al., 2016). Xylan
bagasse by employing Design of Experiment. from finger millet seed coat showed an initial degradation at 130 ◦ C
Source* Sum of DF Mean FValue Prob (Palaniappan et al., 2017), and significant decomposition at 200–400 ◦ C
Squares Square >F (weight loss 65–70%). Similarly, hemicelluloses from the apical, middle
Model 3392.86 14 242.35 1.30 0.0073 Significant
and basal segments of Neolamarckia cadamba exhibited an initial mass
A 713.95 1 713.95 3.84 0.0688 – loss due to moisture evaporation at 170 ◦ C, however, the maximum
B 903.07 1 903.07 4.86 0.0005 Significant weight loss (50%) was realized between 170 and 320 ◦ C (Zhao et al.,
C 155.55 1 155.55 0.84 0.3746 – 2017c). Thus, TGA analysis showed that the xylan extracted from the SB
D 36.65 1 36.65 0.20 0.0032 Significant
biomass was thermally stable and can be used for the production of
A2 111.53 1 111.53 0.60 0.4505 –
B2 247.44 1 247.44 1.33 0.2665 – xylooligosaccharides or other products.
C2 33.05 1 33.05 0.18 0.0091 Significant
D2 33.20 1 33.20 0.18 0.6785 – 3.5.2. Fourier transform infrared spectroscopy of xylan
AB 371.53 1 371.53 2.00 0.0007 Significant The FTIR analysis was carried out to ascertain the purity of the xylan
AC 5.52 1 5.52 0.030 0.8654
polymer for the presence of residual lignin components. FTIR spectral

AD 68.06 1 68.06 0.37 0.5540 –
BC 347.82 1 347.82 1.87 0.0093 Significant profile of xylan displayed typical signal patterns of xylan-rich hemicel­
BD 81.00 1 81.00 0.44 0.0090 Significant lulose (Li et al., 2016). The prominent absorption band at 3444.04 cm− 1
CD 218.30 1 218.30 1.18 0.2444 – was due to the stretching of the hydroxyl (OH) groups present in the
Residual 2786.30 15 185.75 –
xylan fractions as well as water involved in the hydrogen bonding
Lack of 2323.45 10 232.35 2.51 0.1608 Non
Fit significant (Fig. S1b) (Seesuriyachan et al., 2017). The fingerprint region
Pure 462.84 5 92.57 – (1800–600 cm− 1) allows the identification of major chemical groups in
Error the polysaccharides where as the position and intensity of the bands is
Cor Total 6179.16 29 – specific for each polysaccharide. A band observed at 1632 cm− 1 corre­
*
(A- extraction temperature, B- biomass loading, C- extraction time, D- so­ sponded to the mode of bending in water (Khaleghipour et al., 2021;
dium hydroxide) Zhang et al., 2016). The band at around 1400 cm− 1 and 1000 cm− 1
(Fig. S1b) were of typical xylan, and could be attributed to C-O, C-C
responses (92.75%, w/w), thus, validating the process model. The xylan stretching or C-OH bending in the hemicelluloses (Sukri and Sakinah,
yield was increased from initial 41.49% (w/w) to 96.11% (w/w) after 2018; Li et al., 2016).
optimization. Similar to current study, high yield of xylan (95.84%) was The band obtained at 883 cm− 1 was ascribed to the glycosidic link­
extracted from sugarcane bagasse after RSM based optimization of ages present between xylopyranose units in the xylan backbone (Kha­
leghipour et al., 2021). The band at 543.00 cm− 1 (Fig. 3b) resulted from

7
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

Fig. 2. Response surface plots for xylan extraction from delignified sugarcane bagasse showing (a) interaction between temperature and biomass loading; (b)
biomass loading and extraction time; (c) biomass loading and sodium hydroxide concentration; (d) perturbation plot showing the effectiveness of the variables on
xylan extraction from delignified sugarcane bagasse.

C–C–H or C–O–C stretching or bending (Seesuriyachan et al., 2017). The internal structure of the biomass (Fig. S2c), and it appeared heavily
absence of bands corresponding to lignin (1520 cm− 1) and pectin damaged i.e. large scale deformations and shredded look, complete
(1570 cm− 1) residues substantiated the purity of the extracted xylan surface exposure with drastic porosity, swelling and radically increased
(Davila et al., 2016). Furthermore, the removal of acetyl, uronic acid, or volume (Kaweeai et al., 2016; Zhao et al., 2017b). Furthermore, the
ester groups by alkaline xylan extraction was confirmed by the absence swelling of the biomass due to increased hydration of fibrous material in
of bands at 1736 cm− 1 (Sukri and Sakinah, 2018; Wang et al., 2016). the middle lamella causes softening of cell walls which in turn eases the
Therefore, the FTIR analysis suggested that the extracted xylan was breakage of structural tissue. The adjacent cells of the biomass get
significantly pure, and may potentially be used for XOS production or separated resulting in the formation of pores of varying sizes. This may
other industrial applications. be due to the removal of lignin, minerals, extractives and dissolution of
hemicelluloses from the biomass. In a similar study, SEM analysis of
3.5.3. Scanning electron microscopy of sugarcane bagasse after ultra high pressure pretreated corncob biomass showed sieve-like
delignification and xylan extraction structure with pores of different diameters (Seesuriyachan et al.,
Scanning electron microscopy (SEM) was used to examine the 2017). Similarly, liquid ammonia pretreatment of giant reed biomass led
structural aberrations occurred in the biomass due to delignification and to increased porosity as shown by SEM analysis due to cellulose swelling
xylan extraction. The structure of untreated material appeared strongly and lignin relocation, which resulted in easy extraction of xylan from
bundled, rigid, highly ordered, compact, smooth and intact (Fig. S2a). biomass (Zhao et al., 2017b). The SEM analysis of finger millet seed coat
However, after delignification and xylan extraction, the biomass un­ showed irregular and rough surface morphology, and appeared largely
derwent significant morphological changes, and appeared as disrupted, aggregated due to removal of lignin (Palaniappan et al., 2017). Thus,
shredded with increased surface area and porosity. The rough appear­ delignification and/or xylan extraction results in enormous alterations
ance of delignified biomass surface was due to disrupted cell wall con­ in the biomass structure and reduces its recalcitrance.
tents, and/or re-deposition of pseudo-lignin on the surface (Zhao et al.,
2017a). The increased volume and surface area, and formation of
enlarged pores were attributed to the forced removal of lignin from the 3.6. Xylosidase-free endoxylanase from fungal isolate A. flavus MG-7
biomass (Seesuriyachan et al., 2017). The effective removal of outer
lignin layer (peeling) decreased the cell wall thickness (Fig. S2b). Thus, The A. flavus MG-7 produced maximum xylanase (4.35 IU/ml) after
the biomass recalcitrance got decreased which made the hemicelluloses 120 h of submerged fermentation. The concentrated xylanase obtained
more accessible by breaking the lignin–carbohydrate complex. after ammonium sulphate precipitation (activity 33.91 IU/ml), followed
Xylan extraction made significant changes in the external as well as by lyophilisation (activity 49.59 IU/mg), was used for hydrolysis of
xylan. TLC analysis of xylan hydrolysate demonstrated the presence of

8
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

potential, higher specificity towards substrate, long shelf life, and


overall robust nature. Therefore, targeting process-apt xylanases from
the enormous microbial diversity has been a continuous practice (Kundu
et al., 2021; Khandeparker et al., 2017; Bajaj et al., 2011).
Commercial xylan from various biomass sources such as sugarcane
bagasse, rapeseed straw, corn cob and others have generally been used
for the XOS production. However, fewer studies have been undertaken
on XOS production using indigenously produced fungal xylanase (de
Freitas et al., 2021; Kundu et al., 2021; Khandeparker et al., 2017; Bajaj
et al., 2011). A. flavus MG-7 produced β-xylosidase-free endo-β-1,
4-xylanase which is suitable for the industrial production of xylooligo­
saccharides. Analysis of enzymatic hydrolysates via TLC and HPLC
revealed the presence of XOS with different degree of polymerization
(2− 4) i.e. xylobiose, xylotriose and xylotriose, but no xylose at all.
Furthermore, process for XOS production was optimized based on
‘one-variable-at-a-time’ (OVAT) approach to deduce the optimal level of
variables i.e. incubation period, xylan concentration and enzyme
loading. An enzyme dosage of 100 IU was found to be optimal as it
yielded maximum reducing sugar (RS) content (425 mg/g), and further
increase of dosage (120 IU) decreased the RS (378.42 mg/g) (Fig. 4a).
Since the endo-β-1,4-xylanase lacked β-xylosidase activity which is
required for complete hydrolysis of xylooligomers into xylose. There­
fore, accumulation of XOS might have caused end product inhibition
resulting in reduced RS yield at increased enzyme dose (Alvarez et al.,
2017).
Appropriate reaction time for enzymatic XOS production was 24 h as
Fig. 3. Thin layer chromatography analysis of xylan hydrolyzed products after it resulted in highest RS content (429.05 mg/g). However, upon longer
16 h where spot X: standard xylose; SX: standard beechwood xylan; HS1: xylan incubation (48 h), a decreased reducing sugar (RS) (Fig. 4b) yield was
hydrolyzed products with 20IU of enzyme; HS2: xylan hydrolyzed products
obtained which may be attributed to inaccessibility of hydrolytic sites on
with 40IU of enzyme; HS3: xylan hydrolyzed products with 60IU of enzyme;
xylan substrate and/or restrained enzymatic hydrolysis due to end
HS4: xylan hydrolyzed products with 80IU of enzyme and HS5: xylan hydro­
lyzed products with 100IU of enzyme. product inhibition by the accumulated XOS (Sukri and Sakinah, 2018).
Similar to current study, an optimal incubation period of 24 h was re­
ported for XOS production from sugarcane bagasse (de Oliveira et al.,
various xylooligomers of varying degree of polymerization but complete
2018), however, 72 h of incubation was documented to be optimal for
absence of xylose monomers thus, establishing the endoxylanolytic na­
achieving a high concentration of XOS from brewer’s spent grain
ture of xylanase, and complete absence of xylosidase activity (Fig. 3).
(Amorim et al., 2019). The optimum substrate (xylan) concentration for
TLC plate showed that standard xylan did not show any movement
maximum XOS production was 2% (w/v) at enzyme loading of 100 IU,
because of its high molecular weight while standard xylose moved
and incubation time of 24 h as indicated by the maximum RS content
maximally due to its low molecular weight, and xylanase-hydrolysed
(947.49 mg/g). A further increase in xylan concentration, however, led
xylan sample showed an intermediate movement between xylan and
to a decrease in RS yield (Fig. 4c) which might be due to partial inhi­
xylose. Thus, it was established that the xylan has been cleaved endo­
bition of enzyme and/or increased viscosity of the reaction medium that
lytically and resulted in the production of a mixture of oligosaccharides
might have hindered the access of substrate by the enzyme.
but no xylose at all. Total absence of xylose signified the β-xylosidase-
HPLC analysis of enzymatic hydrolysate of xylan obtained under
free nature of endo-β-1,4-xylanase from A. flavus MG-7 (Khandeparker
optimal conditions showed a high XOS concentration of 91.12% (w/w)
et al., 2017).
that was comprised of xylotetrose (69.57%, w/w), xylobiose (13.17%,
Similar to current study, β-xylosidase-free endo-β-1, 4 xylanases were
w/v) and xylotriose (8.37%, w/w). Thus, XOS hydrolysate was rich in
produced from fungal species Neocallimastix patriciarum and Orpino­
xylooligomers exhibiting degree of polymerization of 2–4 oligomers.
myces sp., and used for XOS production from alkali-extracted eucalyptus
This vindicated the effectiveness of the enzymatic regimen as the smaller
xylan (Puchart et al., 2018). A xylanase from Bacillus tequilensis BT21
XOS (X2 to X4) are always preferred for food and pharmaceutical ap­
had endoxylanolytic mode of action, and generated the xylobiose
plications as prebiotic compounds (Reddy and Krishnan, 2016). Enzy­
products upon enzymatic hydrolysis of beechwood xylan (Khandeparker
matic hydrolysis of xylan (sugarcane bagasse) to XOS showed a
et al., 2017). Thus, for production of XOS in high concentration,
conversion yield of 42.3% (Khaleghipour et al., 2021) with xylobiose as
β-xylosidase-free endoxylanases are highly preferred as the presence
the predominant XOS followed by xylotriose. Contrary to current study,
β-xylosidases activity may hydrolyse XOS to produce xylose, and un­
xylan from areca nut husk upon enzymatic hydrolysis (Singh et al.,
dermine the overall process (Surek and Buyukkileci, 2017).
2018) yielded a high amount of xylobiose (71%), moderate content of
xylotriose (26.2%), and a relatively low amount of xylotetrose (2%).
3.7. Production of xylooligosaccharides using A. flavus MG-7 xylanase Thus, the endoxylanase produced from the indigenously isolated fungus
A. flavus MG-7 could be used for the efficient production of xylooligo­
Xylan extracted from SB was used for the production of xylooligo­ saccharides from the sugarcane bagasse biomass for potential applica­
saccharides (XOS) by using A. flavus MG-7 xylanase. XOS production by tions in food and pharmaceutical industries.
enzymatic hydrolysis is generally preferred as it does not produce un­
wanted products like furfural, monosaccharides and other by-products 3.8. Nuclear magnetic resonance analysis of xylooligosaccharides
(Akpinar et al., 2009). For XOS production, xylanase must have
certain typical/ characteristic properties such as lack of β-xylosidase The ultrastructural characteristics of XOS produced from SB-
activity, and ability to generate XOS with various degrees of polymeri­ extracted xylan were studied by NMR. The peak signals in spectral re­
zation (xylobiose X2, xylotriose X3, xylotetrose X4), efficient catalytic gion of 3.0–4.5 ppm (Fig. 5) were ascribed to the protons of xylose,

9
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

Fig. 4. Reducing sugars produced at different (a) enzyme loading; (b) incubation time; and (c) substrate concentration.

Fig. 5. Nuclear magnetic resonance spectra of xylooligosaccharides from sugarcane bagasse (SB) xylan.

arabinose, and glucuronic acid residues (Li et al., 2016). The charac­ non-reducing ends (Bowman et al., 2015). The characteristic signature
teristic signals at 4.20–4.50 ppm (for anomeric protons of reducing end signals of XOS were similar to those reported previously, and some
of XOS) indicated the β-glycosidic linkages in the glycosyl residues, overlaps were due to the mixed nature of XOS (Reddy and Krishnan,
while those at 3.8–3.0 ppm (Fig. 5) signified the non-substituted internal 2016).
xylose units (Zhang et al., 2016). The signal peaks at 3.8–3.2 ppm The absence of signal at 5.00–5.40 ppm indicated that β-glycosidic
showed the presence of β-xylopyranose residues at reducing and linked side-chains (glucuronic acid, α-L-arbinofuranosyl) have been

10
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

cleaved (Zhao et al., 2017c), and XOS were almost linear with few hydrolysis was observed at pH 1 (19.94%) followed by that at pH 2 and 3
attached branches. XOS derived from finger millet seed coat xylan (16.28 and 16.09, respectively), and at pH 4 (14.24%) and 5 (11.83%).
showed major signal at 3.0–5.0 ppm which corresponded to the β-linked Similarly, FOS, a commercial prebiotic showed hydrolysis level of
xylopyranoside residues, and the chemical shift signal at 5.10 ppm 11.16–17.67% in the gastric juice at different pH. An inverse relation­
indicated that 4-O-methyl α-D-glucuronic acid was attached to the XOS ship between degree of XOS hydrolysis and pH of gastric juice (lower the
backbone (Palaniappan et al., 2017). Similarly, XOS from Miscanthus pH more is hydrolysis) may be due to higher vulnerability of glycosidic
xylan showed signals in the range of 3.0–5.5 ppm in the 1H NMR which bonds at low acidic pH (Azmi et al., 2012).
was ascribed to xylose, arabinose, and glucuronic acid residues, thus, Additionally, the incubation time of probiotics in gastric juice also
displaying a typical pattern for xylooligosaccharides (Li et al., 2016). contributes towards the degree of hydrolysis. Mild hydrolysis of XOS-SB
at high acidic pH (pH 1.0) over 6 h of incubation indicated their sub­
3.9. Prebiotic potential of xylooligosaccharides generated from sugarcane stantial resistance to gastric juice. Considering that food takes around
bagasse 4 h to pass through the stomach, it may be inferred that XOS-SB could
reach the intestine in reasonably unaltered form, and be utilized by
The functional potential of XOS as prebiotics was ascertained based probiotics present over there. In a similar study, oligosaccharides syn­
on their growth promoting ability for a probiotic bacterium Lactobacillus thesized from lactulose exhibited considerable resistance to gastroin­
plantarum M-13 (LPM-13), isolated from an exotic source, and charac­ testinal conditions, and showed a mild digestion (<15%)
terized previously (Gupta and Bajaj, 2018, 2017). The results illustrated (Ferreira-Lazarte et al., 2017). Similarly, XOS produced from sugarcane
that the XOS as a sole carbon source excellently supported the growth of straw and coffee husk (Avila et al., 2020), and from sugarcane bagasse
probiotic LPM-13 (18.77 log cfu/ml) that was comparable or even better (Ghosh et al., 2021) were reported to have considerable stability to
than that in standard MRS (with glucose as the carbon source). digestive enzymes.
Maximum growth of LPM-13 was observed at 72 h however, adequate Various factors such as composition of sugar oligosaccahrides,
growth occurred even after 48 h (16.62 log cfu/ml). Furthermore, the configuration of anomer, and linkages involved, may influence the de­
probiotic LPM-13 retained high viability (76%, 14.27 log cfu/ml) even gradability of carbohydrates. Generally, β-linkages are more resistant
after prolonged incubation in MRS-XOS (144 h). than α-ones, however, upon prolonged incubation at low pH (< 4.0)
The growth of probiotic LPM-13 was comparable in MRS-XOS and even the former could undergo hydrolysis (Wang et al., 2015). Thus,
standard MRS up to 48 h, however, at 72 h growth was higher in MRS- resistance of prebiotics to gastrointestinal milieu is absolutely essential
XOS than that in standard MRS. Furthermore, MRS-XOS supported for getting them in large intestine in relatively unaltered form for suc­
higher viability of probiotic LPM-13 (76%) as compared to standard cessfully delivering the health enhancing effects.
MRS (62%) after prolonged incubation (144 h). XOS due to their olig­
omeric structure are metabolized much slowly, and are capable of 3.11. Prebiotic activity score of xylooligosaccharides
supporting the growth/viability of probiotics for a much longer period.
In contrast, glucose being monomeric in nature is metabolized much Prebiotic activity score (PAS) refers to comparative growth analysis
faster, and thus, do not support the sustained growth/viability of pro­ of a bacterial probiotic bacterium (Bifidobacterium spp. and Lactobacillus
biotics over prolonged time periods. Furthermore, growth stimulatory spp.) and a potential bacterial pathogen (E. coli, S. aureus or Clostridium
effect of XOS prebiotics on probiotic LPM-13 (18.77 log cfu/ml) was spp. or others) using an oligosaccharide prebiotic or glucose as the sole
comparable or even better than that of commercial prebiotics such as carbon source. Growth study of probiotic LPM-13 on MRS-XOS or
FOS (18.78 log cfu/ml), lactulose (18.76 log cfu/ml) and inulin standard MRS (containing glucose) showed PAS of 1.5 and 1.46,
(18.65 log cfu/ml). Thus, XOS-SB enriched with X2, X3, and X4 may be respectively, using E. coli and S. aureus as the reference pathogens. The
an effective prebiotic for potential commercial applications. pectin-oligosaccharides produced from citrus peel pectin were reported
Moreover, a concentration of XOS-SB or commercial prebiotics at 1% to have a PAC of 0.41 and 0.92 for L. paracasei LPC-37 and B. bifidum
(w/v) as the sole carbon source was observed to be adequate for sup­ ATCC 29521, respectively, with E. coli as the reference pathogen (Zhang
porting growth of probiotic LPM-13. A higher prebiotic concentration et al., 2018). Similarly, XOS from finger millet seed coat exhibited PAC
(as sole carbon source) may cause disproportion of carbon and the ni­ of 0.7–0.8 for several Lactobacillus spp. using E. coli as the reference
trogen source, and may be detrimental for growth as the nutrient pathogen (Palaniappan et al., 2017).
imbalance affects the enzymatic/metabolic/transport capabilities of A high and positive PAS score indicates relatively higher growth of
bacteria and leads to underutilization of nutrients, and hence impaired probiotics in medium containing prebiotic-oligosaccharides than that of
growth/survival of probiotics (Noori et al., 2017). In a similar study, bacterial pathogens, and substantiates the premise that probiotics can
XOS generated from xylan (extracted from banana pseudostem), using selectively and more efficiently utilize the oligosaccharide-prebiotics for
an in-house produced endoxylanase (Aspergillus versicolour) exhibited growth than the pathogenic bacteria. Thus, PAS analysis of XOS-SB
desirable prebiotic attributes, and adequately supported the growth of established its prospects for application as an ingredient for devel­
probiotic bacteria L. plantarum and L. fermentum. Further, XOS with high oping functional foods.
degree of polymerization supported growth of probiotic bacteria more
efficiently (de Freitas et al., 2021). Thus, XOS generated enzymatically 3.12. Impact of xylooligosaccharides on antibacterial activity of
from LB-derived xylan may have excellent application potential for food L. plantarum M-13
industries especially for production of probiotics-fortified functional
foods. Ever-rising incidences of antibiotic resistance among pathogens
necessitated the development of new/novel approaches for disease
3.10. Stability of xylooligosaccharides under artificial gastric juice control. Probiotics and/or prebiotics mediated antibacterial action may
be a potent approach for controlling/managing infectious diseases
Gastric-stability of oligosaccharide prebiotics is an imperative attri­ caused by intestinal/food borne bacterial pathogens (Gupta and Bajaj,
bute, and refers to their capability to withstand digestive processes, 2018; Palaniappan et al., 2017). Cell free supernatant of L. plantarum
gastric acidity, hydrolysis by enzymes and gastrointestinal absorption. M-13 grown in standard MRS has previously been shown to exhibit mild
Prebiotics must resist the morbid gastrointestinal milieu, and reach the antimicrobial activity against various bacterial pathogens (Gupta and
large intestine to modulate the probiotic population (Ferreira-Lazarte Bajaj, 2018). However, the cell free supernatant obtained after growing
et al., 2017). XOS-SB exhibited moderate to mild hydrolysis in gastric L. plantarum M-13 in MRS with XOS as the sole carbon source, exhibited
juice at different pH (1− 5) over an incubation time of 6 h. Maximum a high inhibitory potential against a wide range of bacterial pathogens.

11
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

Moderate to high inhibition (30–80%) was observed against probiotics) in biorefining industries.
Gram-positive (Bacillus cereus, Bacillus subtilis, Staphylococcus aureus,
Streptococcus pneumoniae, Enterococcus faecalis, Micrococcus luteus) and CRediT authorship contribution statement
Gram-negative pathogenic bacteria (Escherichia coli, Klebsiella sp.,
Pseudomonas aeruginosa, Pseudomonas alcaligenes). Maximum inhibition Mahak Gupta: Methodology, Experimentation, Software, Data
was observed against Klebsiella sp. (74.51%), and was followed by that curation. Ridhika Bangotra: Writing – original draft, Writing – review
against P. aeruginosa (72.96%), E. coli (71.94%), B. cereus (65.42%), S. & editing. Surbhi Sharma: Writing – review & editing. Surbhi Vaid:
aureus (64.92%), M. luteus (63.91%), P. alcaligenes (58.65%), Writing – review & editing. Nisha Kapoor: Writing – review & editing.
S. pneumonia (53.37%), E. faecalis (51.27%), and against B. subtilis Harish Chander Dutt: Writing– review & editing. Bijender Kumar
(49.42%). Bajaj: Conceptualization, Project Administration, Funding acquisition,
The potentiated antimicrobial activity of cell free supernatant of Resources, validation Supervision.
L. plantarum M-13 upon cultivation in MRS-XOS may be due to pro­
duction of short chain fatty acids as a result of metabolism of XOS by the
Declaration of Competing Interest
probiotic bacterium. Production of organic acids resulted in pH reduc­
tion of cultural broth (pH 2.8). Thus, it may be inferred that the path­
The authors declare that they have no known competing financial
ogenic bacteria were inhibited due to low acidic pH as well as due to
interests or personal relationships that could have appeared to influence
some other antimicrobials generated by L. plantarum M-13 (Gupta and
the work reported in this paper.
Bajaj, 2018). In a similar study, the cell free supernatant of L. plantarum
grown in medium containing XOS from finger millet seed coat displayed
Acknowledgements
strong antibacterial activity against several pathogens such as Shigella
flexneri, E. coli, Salmonella typhi, and S. aureus (Palaniappan et al., 2017).
Professor Bijender Kumar Bajaj (BKB) thankfully acknowledges the
Thus, metabolism of prebiotic XOS by probiotic bacteria generates
Department of Science and Technology ‘DST’-Govt. of India, for granting
metabolic products that exert antimicrobial activity against bacterial
financial aid in the form of Research Project (Ref. SR/SO/BB-66/2007),
pathogens.
and other funding agencies such as the Council of Scientific and Indus­
trial Research (CSIR-(Ref. No.: 37(1378)/09/EMR-II), Indian Council of
3.13. Antioxidant potential of xylooligosaccharides
Medical Research (ICMR Project Ref. no: 5/4/1/EXM/12-NCD-II), Uni­
versity Grants Commission (UGC Ref. F. Nos. F-32–550/2006/SR; 42-
Oxidative stress refers to excessive accumulation of free radicals
226/2013/SR) for financial support in the form of research projects.
which disturb the anti-oxidant defence system of the body. Free radicals
BKB gratefully acknowledges the Indo-US Science and Technology
may be implicated in conditions such as ageing and degenerative dis­
Forum for Research Fellowship at Ohio State University, USA,
eases viz. cancer, immune dysfunctioning, arthritis, cardiovascular
Commonwealth Scholarship Commission, for providing Commonwealth
complications, cataracts, liver damage, inflammation, renal failure,
Fellowship for research stay at Institute of Biological, Environmental
brain damage and others. Therefore, an effective antioxidant defence
and Rural Sciences, Aberystwyth University, Aberystwyth, UK, and
mechanism is vital for scavenging the free radicals (Zhang et al., 2018).
Institute of Advanced Study, Durham University, Durham, UK, for
XOS-SB showed good antioxidant potential as indicated by their ability
providing COFUND-International Senior Research Fellowship for
to scavenge DPPH radicals, and further the antioxidant activity
research at Department of Biosciences Durham University, UK. Mrs.
increased proportionally with the XOS-SB concentration. Maximum
Surbhi Vaid thankfully acknowledges Department of Biotechnology,
DPPH scavenging activity of 83.33% was observed at 5.0 mg/ml of
Govt. of India and British Council, UK, for providing Placement under
XOS-SB. In a similar study, xylooliogosaccharides produced from xylan
Newton Bhabha Ph.D. Placement Programme for Research Stay at
extracted from sugarcane straw and coffee husk exhibited antioxidant
Durham University, Durham, UK. Mrs. Mahak Gupta, Mrs. Surbhi Vaid
activity of 71% and 78%, respectively, at a concentration of 2 g/l (Avila
and Ms Surbhi Sharma acknowledge University of Jammu, Jammu, for
et al., 2020). Thus, antioxidant potential of XOS further adds to their
providing University Research Fellowship for research. Authors express
bioactivity potential, and hence augments their prospective for appli­
gratitude to the School of Biotechnology, University of Jammu for lab­
cations in food and pharmaceutical industries.
oratory facilities. SAIF, STIC, Cochin University of Science and Tech­
nology (CUSAT), Kerala, India, is thanked for physicochemical
4. Conclusions
characterization of samples.

Sugarcane bagasse (SB) extracted xylan could be a promising feed­


Appendix A. Supporting information
stock for the production of XOS, therefore leading to value-addition and
efficient valorization of sugarcane bagasse. DoE based optimization of
Supplementary data associated with this article can be found in the
important variables effectively increased the yield of xylan from SB. The
online version at doi:10.1016/j.indcrop.2022.114591.
FTIR and TGA analysis confirmed the purity and thermal stability of
xylan whereas SEM analysis of the biomass indicated that delignification
References
and xylan extraction adequately altered the complex structure of SB
biomass. Enzymatic hydrolysis of SB xylan by a β-xylosidase-free Akpinar, O., Erdogan, K., Bostanci, S., 2009. Production of xylooligosaccharides by
endoxylanase indicated the high efficiency for XOS production. Struc­ controlled acid hydrolysis of lignocellulosic materials. Carbohydr. Res. 344 (5),
tural elucidation of the xylooligosaccharides through NMR analysis 660–666.
Alvarez, C., Gonzalez, A., Negro, M.J., Ballesteros, I., Oliva, J.M., Saez, F., 2017.
revealed that xylooligosaccharides were linear with no substitutions. Optimized use of hemicellulose within a biorefinery for processing high value-added
Furthermore, XOS prebiotics exhibited excellent functional attributes xylooligosaccharides. Ind. Crops Prod. 99, 41–48. https://doi.org/10.1016/j.
and bioactivity spectrum viz. growth promoting potential of probiotic indcrop.2017.01.034.
Amorim, C., Silverio, S.C., Rodrigues, L.R., 2019. One-step process for producing
L. plantarum M-13, high-positive prebiotic activity score, significant
prebiotic arabino-xylooligosaccharides from brewer’s spent grain employing
antioxidant, and antimicrobial activity against bacterial pathogens. The Trichoderma species. Food Chem. 270, 86–94. https://doi.org/10.1016/j.
prebiotic potential of XOS promises their application prospective in food foodchem.2018.07.080.
industries for production of prebiotic-fortified functional foods. Avila, P.F., Martins, M., de Almeida Costa, F.A., Goldbeck, R., 2020.
Xylooligosaccharides production by commercial enzyme mixture from agricultural
Furthermore, the study emphasizes valorization of agricultural residues wastes and their prebiotic and antioxidant potential. Bioact. Carbohydr. Diet. Fibre
(sugarcane bagasse) for production of value-added products (XOS 24, 100234. https://doi.org/10.1016/j.bcdf.2020.100234.

12
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

Azmi, A.F.M.N., Mustafa, S., Hashim, D.M., Manap, Y.A., 2012. Prebiotic activity of Nalawade, K., Baral, P., Patil, S., Pundir, A., Kurmi, A.K., Konde, K., Agrawal, D., 2020.
polysaccharides extracted from Gigantochloa levis (Buluh beting) shoots. Molecules 17, Evaluation of alternative strategies for generating fermentable sugars from high-
1635–1651. https://doi.org/10.3390/molecules17021635. solids alkali pretreated sugarcane bagasse and successive valorization to L (+) lactic
Bajaj, B.K., Sharma, M., Sharma, S., 2011. Alkalistable endo-β-1, 4-xylanase production acid. Renew. Energy 157, 708–717. https://doi.org/10.1016/j.renene.2020.05.089.
from a newly isolated alkalitolerant Penicillium sp. SS1 using agro-residues. 3 Biotech Nath, P., Maibam, P.D., Singh, S., Rajulapati, V., Goyal, A., 2021. Sequential
1, 83–90. https://doi.org/10.1007/s13205-011-0009-5. pretreatment of sugarcane bagasse by alkali and organosolv for improved
Bowman, M.J., Dien, B.S., Vermillion, K.E., Mertens, J.A., 2015. Isolation and delignification and cellulose saccharification by chimera and cellobiohydrolase for
characterization of unhydrolyzed oligosaccharides from switchgrass (Panicum bioethanol production. 3 Biotech 11 (2), 1–16. https://doi.org/10.1007/s13205-
virgatum, L.) xylan after exhaustive enzymatic treatment with commercial enzyme 020-02600-y.
preparations. Carbohydrate research 407, 42–50. https://doi.org/10.1016/j. Noori, N., Hamedi, H., Kargozari, M., Shotorbani, P.M., 2017. Investigation of potential
carres.2015.01.018. prebiotic activity of rye sprout extract. Food Biosci. 19, 121–127. https://doi.org/
Chapla, D., Pandit, P., Shah, A., 2012. Production of xylooligosaccharides from corncob 10.1016/J.FBIO.2017.07.001.
xylan by fungal xylanase and their utilization by probiotics. Bioresour. Technol. 115, de Oliveira, S.M., Moreno-Perez, S., Terrasan, C.R.F., Romero-Fernandez, M., Vieira, M.
215–221. https://doi.org/10.1016/j.biortech.2011.10.083. F., Guisan, J.M., Rocha-Martin, J., 2018. Covalent immobilization-stabilization of
Davila, I., Gordobil, O., Labidi, J., Gullon, P., 2016. Assessment of suitability of vine β-1, 4-endoxylanases from Trichoderma reesei: Production of xylooligosaccharides.
shoots for hemicellulosic oligosaccharides production through aqueous processing. Process Biochem. 64, 170–176. https://doi.org/10.1016/j.procbio.2017.09.018.
Bioresour. Technol. 211, 636–644. https://doi.org/10.1016/j.biortech.2016.03.153. de Oliveira Rodrigues, P., da Silva Barreto, E., Brandão, R.L., Gurgel, L.V.A., Pasquini, D.,
Dubois, M., Gilles, K.A., Hamilton, J.K., Rebers, P.T., Smith, F., 1956. Colorimetric Baffi, M.A., 2021. On-site produced enzyme cocktails for saccharification and
method for determination of sugars and related substances. Anal. Chem. 28 (3), ethanol production from sugarcane bagasse fractionated by hydrothermal and
350–356. https://doi.org/10.1021/ac60111a017. alkaline pretreatments. Waste Biomass Valoriz. 1–12. https://doi.org/10.1007/
Ferreira-Lazarte, A., Montilla, A., Mulet-Cabero, A.I., Rigby, N., Olano, A., Mackie, A., s12649-021-01499-7.
Villamiel, M., 2017. Study on the digestion of milk with prebiotic carbohydrates in a Oniszczuk, A., Oniszczuk, T., Gancarz, M., Szymańska, J., 2021. Role of gut microbiota,
simulated gastrointestinal model. J. Funct. Foods 33, 149–154. https://doi.org/ probiotics and prebiotics in the cardiovascular diseases. Molecules 26 (4), 1172.
10.1016/j.jff.2017.03.031. https://doi.org/10.3390/molecules26041172.
Forsan, C.F., de Freitas, C., Masarin, F., Brienzo, M., 2021. Xylooligosaccharide Palaniappan, A., Balasubramaniam, V.G., Antony, U., 2017. Prebiotic potential of
production from sugarcane bagasse and leaf using Aspergillus versicolor endoxylanase xylooligosaccharides derived from finger millet seed coat. Food Biotechnol. 31,
and diluted acid. Biomass. Convers. Biorefinery 1–16. https://doi.org/10.1007/ 264–280. https://doi.org/10.1080/08905436.2017.1369433.
s13399-021-01403-2. Puchart, V., Franova, L., Krogh, K.B.M., Hoff, T., Biely, P., 2018. Action of different types
de Freitas, C., Terrone, C.C., Masarin, F., Carmona, E.C., Brienzo, M., 2021. In vitro study of endoxylanases on eucalyptus xylan in situ. Appl. Microbiol. Biotechnol. 102,
of the effect of xylooligosaccharides obtained from banana pseudostem xylan by 1725–1736. https://doi.org/10.1007/s00253-017-8722-6.
enzymatic hydrolysis on probiotic bacteria. Biocatal. Agric. Biotechnol. 33, 101973 Reddy, S.S., Krishnan, C., 2016. Production of high-pure xylooligosaccharides from
https://doi.org/10.1016/j.bcab.2021.101973. sugarcane bagasse using crude β-xylosidase-free xylanase of Bacillus subtilis KCX006
Ghosh, A., Chandra, A., Dhar, A., Shukla, P., Baishya, D., 2021. Multi-efficient and their bifidogenic function. LWT-Food Sci. Technol. 65, 237–245. https://doi.
thermostable endoxylanase from Bacillus velezensis AG20 and its production of org/10.1016/j.lwt.2015.08.013.
xylooligosaccharides as efficient prebiotics with anticancer activity. Process Seesuriyachan, P., Kawee-ai, A., Chaiyaso, T., 2017. Green and chemical-free process of
Biochem. 109, 59–71. https://doi.org/10.1016/j.procbio.2021.06.011. enzymatic xylooligosaccharide production from corncob: enhancement of the yields
Guarino, M.P.L., Altomare, A., Emerenziani, S., Di Rosa, C., Ribolsi, M., Balestrieri, P., using a strategy of lignocellulosic destructuration by ultra-high pressure
Cicala, M., 2020. Mechanisms of action of prebiotics and their effects on gastro- pretreatment. Bioresour. Technol. 241, 537–544. https://doi.org/10.1016/j.
intestinal disorders in adults. Nutrients 12 (4), 1037. https://doi.org/10.3390/ biortech.2017.05.193.
nu12041037. Sharma, S., Nargotra, P., Sharma, V., Bangotra, R., Kaur, M., Kapoor, N., Bajaj, B.K.,
Gupta, M., Bajaj, B.K., 2017. Development of fermented oat flour beverage as a potential 2021a. Nanobiocatalysts for efficacious bioconversion of ionic liquid pretreated
probiotic vehicle. Food Biosci. 20, 104–109. https://doi.org/10.1016/j. sugarcane tops biomass to biofuel. Bioresour. Technol. 333, 125191 https://doi.org/
fbio.2017.08.007. 10.1016/j.biortech.2021.125191.
Gupta, M., Bajaj, B.K., 2018. Functional characterization of potential probiotic lactic acid Sharma, V., Nargotra, P., Bajaj, B.K., 2019. Ultrasound and surfactant assisted ionic
bacteria isolated from kalarei and development of probiotic fermented oat flour. liquid pretreatment of sugarcane bagasse for enhancing saccharification using
Probiotics Antimicrob. Proteins 10 (4), 654–661. https://doi.org/10.1007/s12602- enzymes from an ionic liquid tolerant Aspergillus assiutensis VS34. Bioresour.
017-9306-6. Technol. 285, 121319 https://doi.org/10.1016/j.biortech.2019.121319.
Hesam, F., Tarzi, B.G., Honarvar, M., Jahadi, M., 2020. Valorization of sugarcane bagasse Sharma, V., Nargotra, P., Sharma, S., Bajaj, B.K., 2021b. Efficacy and functional
to high value-added xylooligosaccharides and evaluation of their prebiotic function mechanisms of a novel combinatorial pretreatment approach based on deep eutectic
in a synbiotic pomegranate juice. Biomass Convers. Biorefinery 1–13. https://doi. solvent and ultrasonic waves for bioconversion of sugarcane bagasse. Renew. Energy
org/10.1007/s13399-020-01095-0. 163, 1910–1922. https://doi.org/10.1016/j.renene.2020.10.101.
Holscher, H.D., 2017. Dietary fiber and prebiotics and the gastrointestinal microbiota. Singh, R.D., Banerjee, J., Sasmal, S., Muir, J., Arora, A., 2018. High xylan recovery using
Gut Microbes 8, 172–184. https://doi.org/10.1080/19490976.2017.1290756. two stage alkali pre-treatment process from high lignin biomass and its valorisation
Kaweeai, A., Srisuwun, A., Tantiwa, N., Nontaman, W., Boonchuay, P., Kuntiya, A., to xylooligosaccharides of low degree of polymerisation. Bioresource technology
Chaiyaso, T., Seesuriyachan, P., 2016. Eco-friendly processing in enzymatic 256, 110–117. https://doi.org/10.1016/j.biortech.2018.02.009.
xylooligosaccharides production from corncob: Influence of pretreatment with Sluiter, A., Hames, B., Ruiz, R., Scarlata, C., Sluiter, J., Templeton, D., 2008.
sonocatalytic–synergistic Fenton reaction and its antioxidant potentials. Ultrason. Determination of sugars, byproducts, and degradation products in liquid fraction 3
Sonochem. 31, 184–192. https://doi.org/10.1016/j.ultsonch.2015.12.018. process samples. NREL/TP-510-42623. Nat. Renew. Energy Lab. Gold 4391924.
Khaire, K.C., Moholkar, V.S., Goyal, A., 2021. Bioconversion of sugarcane tops to Sukri, S.S.M., Sakinah, A.M., 2018. Production of high commercial value
bioethanol and other value added products: an overview. Mater. Sci. Energy xylooligosaccharides from Meranti wood sawdust using immobilised xylanase. Appl.
Technol. 4, 54–68. https://doi.org/10.1016/j.mset.2020.12.004. Biochem. Biotechnol. 184, 278–290. https://doi.org/10.1007/s12010-017-2542-0.
Khaleghipour, L., Linares-Pastén, J.A., Rashedi, H., Siadat, S.O.R., Jasilionis, A., Al- Surek, E., Buyukkileci, A.O., 2017. Production of xylooligosaccharides by autohydrolysis
Hamimi, S., Karlsson, E.N., 2021. Extraction of sugarcane bagasse arabinoxylan, of hazelnut (Corylus avellana L.) shell. Carbohydr. Polym. 174 https://doi.org/
integrated with enzymatic production of xylo-oligosaccharides and separation of 10.1016/j.carbpol.2017.06.109.
cellulose. Biotechnol. biofuels 14 (1), 1–19. https://doi.org/10.1186/s13068-021- Valim, I.C., Rego, A.S.C., Vieira, A.A.S., Vilani, C., Martins, A.R.F.A., Santos, B., 2018.
01993-z. Response surface methodologies to investigate the pretreatment of sugarcane
Khandeparker, R., Parab, P., Amberkar, U., 2017. Recombinant xylanase from Bacillus bagasse via alkaline hydrogen peroxide. Chem. Eng. Trans. 65, 421–426. https://doi.
tequilensis BT21: Biochemical characterization and its application in the production org/10.3303/CET1865071.
of xylobiose from agricultural residues. Food Technol. Biotechnol. 55, 164–172. Wang, X., Huang, M., Yang, F., Sun, H., Zhou, X., Guo, Y., Zhang, M., 2015. Rapeseed
https://doi.org/10.17113/ftb.55.02.17.4896. polysaccharides as prebiotics on growth and acidifying activity of probiotics in vitro.
Kundu, P., Kansal, S.K., Elumalai, S., 2021. Synergistic action of alkalis improve the Carbohydr. Polym. 125, 232–240. https://doi.org/10.1016/j.carbpol.2015.02.040.
quality hemicellulose extraction from sugarcane bagasse for the production of Wei, L., Yan, T., Wu, Y., Chen, H., Zhang, B., 2018. Optimization of alkaline extraction of
xylooligosaccharides. Waste Biomass Valoriz. 12 (6), 3147–3159. https://doi.org/ hemicellulose from sweet sorghum bagasse and its direct application for the
10.1007/s12649-020-01235-7. production of acidic xylooligosaccharides by Bacillus subtilis strain. Plos One 13 (4),
Li, Y., Cui, J., Zhang, G., Liu, Z., Guan, H., Hwang, H., Wang, P., 2016. Optimization MR44. https://doi.org/10.1371/journal.pone.0195616 e0195616.
study on the hydrogen peroxide pretreatment and production of bioethanol from Yang, J., Xu, Y., 2018. Functional Carbohydrate Polymers: Prebiotics. In Polymers for
seaweed Ulva prolifera biomass. Bioresour. Technol. 214, 144–149. https://doi.org/ Food Applications. Springer Cham, pp. 651–691. https://doi.org/10.1007/978-3-
10.1016/j.biortech.2016.04.090. 319-94625-2_24.
Meghana, M., Shastri, Y., 2020. Sustainable valorization of sugar industry waste: status, Yang, L., Deng, Y., Wang, X., Zhang, W., Shi, X., Chen, X., Zhang, F., 2021. Global direct
opportunities, and challenges. Bioresour. Technol. 303, 122929 https://doi.org/ nitrous oxide emissions from the bioenergy crop sugarcane (Saccharum spp. inter-
10.1016/j.biortech.2020.122929. specific hybrids). Sci. Total Environ. 752, 141795 https://doi.org/10.1016/j.
Millette, M., Luquet, F.M., Lacroix, M., 2007. In vitro growth control of selected scitotenv.2020.141795.
pathogens by Lactobacillus acidophilus-and Lactobacillus casei-fermented milk. Yiin, C.L., Quitain, A.T., Yusup, S., Uemura, Y., Sasaki, M., Kida, T., 2017. Choline
Letters in applied microbiology 44 (3), 314–319. https://doi.org/10.1111/j.1472- chloride (ChCl) and monosodium glutamate (MSG)-based green solvents from
765X.2006.02060.x. optimized cactus malic acid for biomass delignification. Bioresour. Technol. 244,
941–948. https://doi.org/10.1016/j.biortech.2017.08.043.

13
M. Gupta et al. Industrial Crops & Products 178 (2022) 114591

Zhang, J., Wang, Y.H., Wei, Q.Y., Du, X.J., Qu, Y.S., 2018a. Investigating desorption Zhao, C., Qiao, X., Cao, Y., Shao, Q., 2017a. Application of hydrogen peroxide presoaking
during ethanol elution to improve the quality and antioxidant activity of xylo- prior to ammonia fiber expansion pretreatment of energy crops. Fuel 205, 184–191.
oligosaccharides from corn stalk. Bioresour. Technol. 249, 342–347. https://doi.org/ https://doi.org/10.1016/j.fuel.2017.05.073.
10.1016/j.biortech.2017.09.203. Zhao, C., Cao, Y., Ma, Z., Shao, Q., 2017b. ). Optimization of liquid ammonia
Zhang, S., Hu, H., Wang, L., Liu, F., Pan, S., 2018. Preparation and prebiotic potential of pretreatment conditions for maximizing sugar release from giant reed (Arundo donax
pectin oligosaccharides obtained from citrus peel pectin. Food Chem. 244, 232–237. L.). Biomass Bioenergy 98, 61–69. https://doi.org/10.1016/j.
https://doi.org/10.1016/j.foodchem.2017.10.071. biombioe.2017.01.001.
Zhang, Y., Yu, G., Li, B., Mu, X., Peng, H., Wang, H., 2016. Hemicellulose isolation, Zhao, X., Tong, T., Li, H., Lu, H., Ren, J., Zhang, A., Wu, A.M., 2017c. Characterization of
characterization, and the production of xylo-oligosaccharides from the wastewater hemicelluloses from Neolamarckia cadamba (Rubiaceae) during xylogenesis.
of a viscose fiber mill. Carbohydr. Polym. 141, 238–243. https://doi.org/10.1016/j. Carbohydr. Polym. 156, 333–339. https://doi.org/10.1016/j.carbpol.2016.09.041.
carbpol.2016.01.022.

14

You might also like