You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/301757199

Bioconversion of Lignocellulosic biomass to xylitol: An overview

Article  in  Bioresource Technology · April 2016


DOI: 10.1016/j.biortech.2016.04.092

CITATIONS READS

84 1,855

4 authors, including:

Venkateswar Rao Linga Jyosthna Khanna Goli


Osmania University Osmania University
136 PUBLICATIONS   3,684 CITATIONS    6 PUBLICATIONS   96 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Yeast and lactic acid bacteria for probiotic applications View project

Bioethanol Production from Prosopis Juliflora Using Thermotolerant Saccharomyces Cereviseae View project

All content following this page was uploaded by Jahnavi Gentela on 01 February 2018.

The user has requested enhancement of the downloaded file.


Bioresource Technology xxx (2016) xxx–xxx

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Review

Bioconversion of lignocellulosic biomass to xylitol: An overview


Linga Venkateswar Rao ⇑, Jyosthna Khanna Goli, Jahnavi Gentela, Sravanthi Koti
Dept. of Microbiology, Osmania University, Hyderabad, Telangana State 500 007, India

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Effective pre treatment strategies


were discussed. Lignocellulosic Biomass Hydrolysis
 Recent detoxification methodologies
have been elaborated. Acid Based Enzyme Based
 Emphasis on the overall processes Processing Inhibitors

involved in the bioconversion of


biomass to xylitol. Hexoses Hexoses
Detoxification Pentoses
 Focus on the upcoming fermentation
Pretreatment
strategies.
Xylose
Reduction in crystallinity of HC Fermentation
Concentration,
Lignin removal
Purification &
Xylitol Crystallization
Hydrolysis

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

Article history: Lignocellulosic wastes include agricultural and forest residues which are most promising alternative
Received 31 December 2015 energy sources and serve as potential low cost raw materials that can be exploited to produce xylitol.
Received in revised form 16 April 2016 The strong physical and chemical construction of lignocelluloses is a major constraint for the recovery
Accepted 19 April 2016
of xylose. The large scale production of xylitol is attained by nickel catalyzed chemical process that is
Available online xxxx
based on xylose hydrogenation, that requires purified xylose as raw substrate and the process requires
high temperature and pressure that remains to be cost intensive and energy consuming. Therefore, there
Keywords:
is a necessity to develop an integrated process for biotechnological conversion of lignocelluloses to xylitol
Xylitol
Pretreatment
and make the process economical. The present review confers about the pretreatment strategies that
Lignocellulose facilitate cellulose and hemicellulose acquiescent for hydrolysis. There is also an emphasis on various
Detoxification detoxification and fermentation methodologies including genetic engineering strategies for the efficient
Fermentation conversion of xylose to xylitol.
Ó 2016 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2. Pretreatment of lignocellulosic materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.1. Physical pretreatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.2. Chemical pretreatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.3. Physico-chemical pretreatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.4. Biological pretreatment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3. Hydrolysis of biomass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

⇑ Corresponding author.
E-mail address: vrlinga@gmail.com (L. Venkateswar Rao).

http://dx.doi.org/10.1016/j.biortech.2016.04.092
0960-8524/Ó 2016 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
2 L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx

4. Inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5. Detoxification of acid hydrolysate. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5.1. Chemical detoxification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5.2. Nanofiltration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5.3. Vacuum membrane distillation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5.4. Electrochemical detoxification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5.5. Biological detoxification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
6. Enzymatic hydrolysis by hemicellulases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
7. Biotechnological production of xylitol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
7.1. Fed batch fermentation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
7.2. Cell recycling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
8. Fermentation parameters influencing xylitol production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
8.1. Carbon source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
8.2. Nitrogen source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
8.3. Aeration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
8.4. Effect of temperature and pH on xylitol production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
8.5. Co-substrate and oxygen transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
9. Continuous fermentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
10. Molecular approaches for xylitol production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
11. Strain Improvement strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
12. Enzymatic approach for xylitol production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
13. Purification and recovery. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
14. Cost economics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
15. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

1. Introduction total xylitol and rest of the production is by Europe, United States
and Australia (Rao et al., 2012). Asia pacific is the largest consumer
Xylitol (five carbon sugar alcohol), an artificial sweetener, could of xylitol. In Asia Pacific, China has the largest market for xylitol
be produced from second most abundant polysaccharide, xylan and next India is also projected to have fastest growing market
rich hemicellulose which on hydrolysis produces xylose. It is a (http://www.marketresearch.com/product/sample-8164119.pdf)
sugar substitute used in dietary, food and pharmaceutical indus- (Fig. 1). Chewing gum industry has the largest market for xylitol
tries due to its properties like low energy content, anticariogenic- consumption compared to other industries and was found to be
ity, tooth rehardening, preventive against otitis, ear and upper 80% in 2010 and estimated to consume about 163 thousand metric
respiratory infections etc. Though most of xylitol is produced tons by 2020 with 67% of global xylitol consumption (Fig. 2).
chemically based on catalytic reduction of pure xylose under high (http://www.marketresearch.com/product/sample-8164119.pdf).
pressure and temperature using expensive catalyst (usually Chewing gum brands like Gross and Wrigley use only xylitol for
Ni-catalyst), biotechnological production of xylose to xylitol offers the manufacture of sugar-free chewing gums. DuPont, Cargill, CSPC
a better alternative in terms of energy intensiveness of the produc- Shengxue Glucose, Ingredion, Mitsubishi Shoji Food tech were
tion process of xylitol and overall process cost. found to be the leading vendors of xylitol in the market, other
As the consumers are inclining towards sugar free and low calo- important vendors include ZuChem, Hangzhou Shouxin Biological
rie food products due to health and weight consciousness, the Technology, Novagreen, Tata Chemicals, Roquette Frres and Thom-
demand for xylitol is increasing. The global consumption of xylitol son Biotech. http://www.prnewswire.com.
was approximately 160 thousand metric tons in 2013 which is Lignocellulosic wastes are abundantly available, renewable and
equivalent to market value of US$ 670 million and is expected to are inexpensive energy sources generally found in nature (Misra
reach 242 thousand metric tons by 2020 equivalent to US $ 1bn et al., 2013). Lignocelluloses comprise of cellulose, hemicellulose
with a CAGRS of over 6% by value and volume (http://www. and lignin. Hemicellulose content in lignocellulosic biomass is
prnewswire.com). Globally Asia alone produces 50% of the world’s usually in the range of 20–35%. Some of the lignocellulosic raw

Fig. 1. Asia-pacific xylitol market overview (2009–2020) in metric tons.

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx 3

Fig. 2. Global xylitol chewing gum (2009-2020) market by geographic region –in metric tons.

materials used for the production of xylitol mainly include input. Alkaline pretreatment digests lignin and make holocellulose
corncobs, wheat straw, corn stover, wheat bran, miscanthus etc. accessible for acid or enzymatic degradation. Generally hydroxides
A prerequisite step in the production of xylitol is the pretreat- of sodium, potassium, calcium and ammonium are employed in
ment of the substrate which is essential for removal of lignin and this process. Xu et al. (2007) studied the influence of equal concen-
to reduce the crystallinity of the substrate. Hydrolysis of lignocel- trations of sodium and potassium hydroxides in the recovery of
lulosic biomass using acids is most widely used, during which hemicellulose from perennial ryegrass leaves, and found sodium
optimization of various process parameters is required with a focus hydroxide was effective for the removal of hemicelluloses whereas
on xylose recovery. During the process of acid hydrolysis, various use of potassium hydroxide resulted in higher purity of hemicellu-
inhibitors are released which show negative impact on fermenta- lose. Manas and Krishnan (2015) studied the pretreatment of rice
tion. Detoxification process reduces the concentration of inhibitors. straw using 15% aqueous ammonia at 120 °C, which removed sig-
Later the sugars especially xylose will be converted to xylitol nificant amount of lignin i.e., 65.84% and enhanced subsequent
during fermentation and the xylitol thus produced is further enzymatic digestibility by four fold.
concentrated, purified and crystallized. In acid pretreatment method, acids like sulfuric, hydrochloric,
In the earlier reviews there was focus on the conversion of lig- hydrofluoric, nitric and phosphoric acid etc. are used as catalysts.
nocellulosic wastes, the effect of process parameters and evalua- Dilute sulfuric acid hydrolysis is a well-known process for depoly-
tion on biotechnological strategies involved in xylitol production. merization of hemicelluloses in lignocellulosic biomass into mono-
This review mainly focuses on the recent technologies employed meric sugars (Garcia Martin et al., 2013). In the presence of
in the exploitation of lignocellulosic wastes with respect to pre- concentrated acids, hemicellulose and cellulose get hydrolyzed
treatment, detoxification strategies for removal of inhibitors and completely and high yield of monomeric sugars are obtained but
the use of various genetic engineering aspects that necessitates the major drawback is their corrosive nature and the need to recy-
xylitol production. cle acids to lower the cost of pretreatment. In recent times, organic
acids were also recommended as alternatives to inorganic acids to
avoid reactor corrosions and to reduce energy requirement for the
2. Pretreatment of lignocellulosic materials
recovery of acid. Acids like maleic, succinic, oxalic, fumaric and
acetic acid have been extensively studied. Qin et al. (2012)
Pretreatment is the prerequisite step which plays a key role in
observed that, when corn stover was pretreated with sulfuric, oxa-
subsequent saccharification of biomass and its conversion to xyli-
lic, citric, tartaric and acetic acid, it was found that 91.7–96.8% glu-
tol. The methods of pretreatment can be divided into physical,
can as well as higher xylan and less furfural was obtained using
chemical, physico-chemical and biological methods as there are
acetic acid. Pretreating lignocellulosic biomass using oxidative
many reports available on pretreatment, in the current review
reagents involves the reduction of crystallinity in hemicellulose
we tried to put forth in brief about the recent pretreatment
and cellulose and also disrupts the association between carbohy-
technologies.
drates and lignin. Nevertheless the cost of oxidative pretreatment
is more than conventional alkali and acid pretreatments hence it
2.1. Physical pretreatment is generally used as a support for other pretreatments to remove
residual lignin from lignocellulosic substrates (Qi et al., 2009).
Physical pretreatment methods disrupt the structural integrity The oxidative methods comprise of alkaline peroxide pretreat-
of lignocellulosic substrates and thereby increase their accessibil- ment, ozonolysis, and wet oxidation. In wet oxidation, the biomass
ity to acids or enzymes. Various types of physical pretreatment is treated with water and air or oxygen at high temperatures for
methods which can improve the efficiency of lignocellulosic sac- comparatively short residence time. Recent studies have stated
charification include reducing size through chipping, grinding, or that hydrogen peroxide is the most frequently used oxidizing agent
milling and irradiation by gamma rays, electron beam or micro- that disintegrate about 50% of lignin and most of the hemicellulose
wave (Taherzadeh and Karimi, 2008). with 2% H2O2 at 30 °C. As ozone has the electron deficiency in the
terminal oxygen, it is used as the strongest oxidizing agent. It
2.2. Chemical pretreatment attacks lignin, an electron-rich substrate selectively than carbohy-
drates. In a study conducted by Garcia Cubero et al. (2009), O3
Chemical pretreatment methods involve the use of alkali, dilute appreciably reduced lignin content with meagre loss in hemicellu-
acid, ammonia, organic solvents, oxidative agents and other chem- lose during oxidative delignification process.
icals. These methods are effortless and give good conversion yields Organic solvent pretreatment is an additional probable method
in less span of time. However, the alkali used in the treatment has in the extraction of hemicelluloses. As compared to other chemical
to be recovered after the process which needs additional energy pretreatments, this process has many advantages such as easy

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
4 L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx

recovery of solvent by distillation, low environmental impact, and in the saw dust fibers and after pretreatment there was a twenty
recovery of high quality lignin as by-product (Pan et al., 2006). fold increase in reducing sugars. Fungal laccases are the most
Dimethyl sulfoxide (DMSO) is the most common neutral solvent widely used lignolytic enzymes in biotechnology but pretreatment
that has been used to recover hemicelluloses from holocellulose with a single enzyme cannot completely remove lignin in either
without disturbing the acetyl ester compounds and the glycosidic wood or herbaceous species.
bonds. Though most of the solvents used in the organosolv pre-
treatment can be recycled from the reactor to decrease the cost,
the high price and potential hazards of handling large volumes of 3. Hydrolysis of biomass
organic solvents limit the utilization of organosolv process. Process
using ionic liquids (IL) is another alternative for pretreatment of Hydrolysis of lignocellulosic biomass can be exploited using
lignocellulosic materials. Ionic liquids disturb the non-covalent acids/enzymes. In chemical hydrolysis, sulfuric acid is the most
interactions in the lignocellulosic matrix without any significant investigated and predominantly used acid since it is less volatile,
degradation. Hemicellulose liberated from IL solutions show do not corrode the equipment and less expensive. Although other
enhanced enzymatic digestibility. Although the inhibitor formation acids such as hydrochloric acid, phosphoric acid, nitric acid have
is minimal, a small amounts of ILs remained in the residual also been reported. Dilute acid hydrolysis is preceded by enzymatic
pretreated materials are toxic to fermentative microorganisms hydrolysis and is more favored over enzymatic hydrolysis due to
(Yang and Wyman, 2008) significantly. On the other hand, ILs are its lower cost, simplicity, effectiveness, economic feasibility and
mostly expensive and thus further research is essential to faster kinetics (Garcia Martin et al., 2013).
standardize the solubility of ILs for lignocellulosic substrates. Varying the concentration of acid, temperature, residence time,
and solid to liquid ratio play a crucial role during acid hydrolysis in
2.3. Physico-chemical pretreatment sugar recovery. In a study conducted by Misra et al. (2013), hydrol-
ysis of corn cob at optimized condition of 121 °C for 30 min. using
This category includes the majority of pretreatment technolo- 1.0% H2SO4, with solid: liquid ratio of 1:8 resulted in the (g/L):
gies such as ultrasonication, microwave irradiation, steam xylose 21.98, glucose 3.56, arabinose 2.17, acetic acid 1.31, furfural
explosion, twin screw extruder and so on. These forms of pretreat- 0.21. Recently, Zahed et al. (2015) conducted dilute acid hydrolysis
ment exploit the use of conditions and compounds that affect the of rice straw by varying the concentrations of sulfuric acid (2%,
physical and chemical properties of biomass. Twin-screw extruder 3.5%, and 5%) at fixed temperature of 100 °C. At an optimum con-
or extrusion reactor is a thermo-mechanic and chemical fractiona- centration of 3.5% sulfuric acid, 2.3 and 15.05 g/L of glucose and
tion system that allows the integration of extrusion, cooking, xylose was obtained.
liquid–solid extraction, and filtration in a single step. The twin- Meranti wood sawdust, is the most promising and economic
screw reactor extracts up to 90% of the original hemicellulose lignocellulosic biomass used for manufacturing xylitol. The opti-
depending upon the content of xylan. The yield after extraction mum temperature, acid concentration and time was found to be
depends significantly on temperature whereas reduction in screw 124 °C, 3.26%, and 80 min. respectively to attain maximum saccha-
rotation speed and solid flow rate influenced the average residence rification. Under these conditions xylose 18.8 g/L, glucose 4.64 g/L,
time of the liquid phase which increased the extraction yield of arabinose 2.55 g/L, acetic acid 4.14 g/L, furfural 0.55 g/L, HMF
hemicellulose (N0 diaye and Rigal, 2000). 0.08 g/L, and lignin degradation products of 1.55 g/L with a xylose
Hydrothermal pretreatment includes steam explosion and yield and selectivity attained of 90.6% and 4.05 g/g, respectively
liquid hot water hydrolysis. In the process of steam explosion, bio- (Rafiqul et al., 2014).
mass is generally treated with high pressure and temperature for a Saleh et al. (2014) conducted acid hydrolysis of olive stones
certain amount of time after which the sample is rapidly decom- with 0.025 M sulfuric acid at a temperature of 195 °C for 5 min.
pressed resulting in the breakdown of the lignin carbohydrate in a 2 L Parr reactor, series 4522 (Moline, IL, USA). The hydrolysis
complex. The principal fraction of the biomass digested in the liq- resulted in the recovery of 89.7% of the total D-xylose (20.5 g
uid phase during pretreatment is hemicellulose whereas the lignin extracted from 100 g raw material), which was further concen-
is transformed as result of the high temperature (Kumar et al., trated and detoxified under vacuum at 40 °C. The hydrolysate con-
2012). In hot-compressed water or liquid hot water pretreatment, tained D-xylose 20.0 g/L, L-arabinose 0.88 g/L, D-galactose 0.55 g/L,
temperature above 200 °C at various pressures below the critical D-glucose 0.83 g/L, and acetic acid 2.86 g/L.
point is used. This method can recover comparatively high hemi- The use of phosphoric acid has been advantageous since the
cellulose with low levels of inhibitory compounds (Abdullah neutralization of acid hydrolysate is performed using NaOH, and
et al., 2014). Moretti et al. (2014) reported that microwave- the salt, sodium phosphate thus formed is used as nutrient by
assisted extraction can be a novel eco-friendly way for the recovery microorganisms during xylitol fermentation. Using phosphoric
of hemicelluloses. The difficulty in microwave-assisted extraction acid Martínez et al., 2012) performed hydrolysis in a stirred batch
is to get a good yield of sugars without extensive degradation of reactor of 1 L capacity, where sunflower stalk was used as sub-
the hemicelluloses and contamination with dissolved lignin. This strate and hydrolysis was performed at 95 °C. At phosphoric acid
method is efficient than the traditional chemical pretreatment. concentration of 2.67 mol/L, the maximum hemicellulose of
82.7% was obtained. The polyphenols like hydroxytyrosol
2.4. Biological pretreatment (12.98 mg/L), vanillin (1.25 mg/L), veratric acid (1.23 mg/L), syrin-
gic acid (0.88 mg/L), vanillic acid (0.52 mg/L), protocatechuic acid
Biological pretreatment using either organism directly or its (0.42 mg/L) and caffeic acid (0.17 mg/L) were detected in the
metabolite for pretreatment of biomass is a promising eco- hydrolysate.
friendly technology as it has various advantages and is economi- The use of potential catalysts such as amorphous carbon mate-
cally viable strategy for enhancing the rate of enzymatic sacchari- rial bearing SO3H and OH groups, micro porous crystalline material
fication. Microorganisms like brown, white and soft rot fungi were (zeolites), and sulfated zirconia were found to be promising for the
used for degradation of lignin in biological pretreatment. Biological hydrolysis of lignocelluloses. In a recent study by Zhong et al.
pretreatment of Eucalyptus grandis, saw dust degradation patterns (2015), a nanoscale catalyst, solid acid SO2
4 / Fe2O3 with both Bron-
and saccharification kinetics with white rot fungi was reported by sted and Lewis acidity was used for hydrolysis of wheat straw,
Castoldi et al. (2014). The treatment produced structural changes which led to the hemicellulose recovery of 63.5%. This catalyst

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx 5

was found to be more promising as green catalyst and could be 5. Detoxification of acid hydrolysate
recycled six times with high activity.
The use of plasma acid is more beneficial over sulfuric acid in Several detoxification strategies were used to remove toxic
cost as it can be obtained and separated more easily. Large num- compounds from the hydrolysates as the minute amounts may
bers of protons are produced by the dielectric barrier discharge interfere and prevent the microbial growth. The efficiency of
of water vapor and these protons make plasma acid that could detoxification process depends on the raw material, type of hydrol-
hydrolyze hemicellulose. At optimum conditions of 100 °C, ysis process, and microorganism employed (Rao et al., 2006).
50 min. and pH 2.81, 38.67% xylose; 9.28% glucose and 3.09% galac-
tose was obtained. It was observed and finally concluded that the 5.1. Chemical detoxification
use of plasma acid over sulfuric acid is environmental-friendly
and likely to bring about efficient hydrolysis of hemicellulose Adsorption with activated charcoal is a well-established and
(Wang et al., 2013). However, the major bottleneck with acid low cost detoxification method. The detoxification of sago trunk
hydrolysis is the degradation of sugars into inhibitory components hydrolysate with 2.5% activated charcoal at the reaction time of
that affects yeast and there by consequently impinge the process of 60 min facilitated the reduction of 53% and 78% of furfural and
xylitol fermentation. phenolic compounds. It is evident that the maximum removal of
phenolic compounds is attained by activated charcoal with a xyli-
4. Inhibitors tol yield of 0.78 g/g when compared with the untreated hydroly-
sate (0.307 g/g). This study proposes that detoxification method
The hydrolysis of lignocellulosic biomass (LB) using acids using activated charcoal has a significant impact on xylitol produc-
results in the generation of various inhibitors in addition to sugars. tion (Siti et al., 2011).
The inhibitors released affect the growth of microorganism with an Arruda et al. (2011) elaborated the process involved in using
impact on fermentation process. The tolerable degree of inhibitors ion-exchange resins. Sugarcane hemicellulosic hydrolysate was
is reliant on the microorganism used and its level of adaptation, the treated with the resins at the ratio of 1:2 v/v. The hydrolysate
fermentation process employed, and the presence of other inhibi- was vacuum evaporated and then subjected to ion-exchange resins
tors (Kumar et al., 2009). like A-860S (macroporous strong base anion exchanger), A-500PS
Phenolics are the major inhibitors obtained as a result of lignin (macroporous type I strong base anion exchanger) and C-150
breakdown which lead to the loss in integrity of biological mem- (macro porous strong acid cation exchanger), which led to the
branes. Usually 0.1 g/L concentration of phenolics compounds removal of 93% phenolics and 64.9% acetic acid.
affects the xylose consumption, cell growth and xylitol production. Zhu et al. (2011) performed trialkylamine extraction to remove
Phenolics include vanillic acid, syringic acid, vanillone, and the inhibitors from corn stover prehydrolysate. The advantages in
syringaldehyde. The most commonly observed phenolic deriva- using trialkylamine is that it is highly efficient in removing inhibi-
tives in the lignocellulosic acid hydrolysate include 4-hydroxy ben- tors, requires low temperature, low energy, low cost. The trialky-
zoic acid, ferulic acid, and guaiacol (Mussatto and Roberto, 2004). lamine contained 50% n-octanol-20% kerosene. It was able to
Furfural and HMF: The higher temperature and pressure during remove 73.3% of acetic acid, 45.7% of 5-HMF and 100% of furfural.
acid hydrolysis result in the degradation of xylose that leads to the
formation of furfural while 5-HMF is generated during hexose 5.2. Nanofiltration
degradation. Furfural and HMF not only affect the enzymatic and
biological activities of yeast but damage DNA and further inhibit Nanofiltration (NF) is a potential and cost-competitive mem-
protein and RNA synthesis (Mussatto and Roberto, 2004). brane separation technology. The molecular weight cut-off ranges
Aliphatic acids such as acetic, formic and levulinic acid are also from 150 to 1000 g/mol, thus facilitating the retention of com-
formed by de-acetylation of hemicellulose or HMF breakdown. The pounds with molecular weight up to 150–250 g/mol as well as
lower concentration i.e., 1.0 g/L in the fermentation medium ema- charged molecules. NF finds its applications in detoxification,
nates the bioconversion of xylose to xylitol. The presence of acetic fermentation broth separation, fractionation of sugars and sugar
acid at a concentration of 2.6 g/L compared to the control has concentration (Weng et al., 2009).
reduced the consumption of xylose, xylitol production, and cell Recently Nguyen et al. (2015) worked on NF and reverse osmo-
growth by 13%, 18%, and 30%, respectively (Kumar et al., 2009). sis (RO). The low molecular weight cut-off (150–400 g/mol) of the
The concentration of inhibitors obtained by hydrolysis of different membranes of NF and RO were monitored on a flat-sheet plant for
lignocellulosic wastes and their respective detoxifying agents used their capability in separating C5 and C6 sugars from acetic acid,
were given in Table 1. furfural, 5-HMF and vanillin in a model solution. RO led to the

Table 1
Concentration of inhibitors from different lignocellulosic wastes and their respective detoxifiying agents.

Lignocellulosic Inhibitors (g/L) Detoxifying agents employed for inhibitor References


biomass removal
Mixed hard wood Acetic acid, 11.25; HMF, 0.19; furfural, 1.75; Total Resin XAD, XAD-ED, ED, ED-XAD Kundu et al. (2015)
phenolic compound, 4.52
Yellow poplar Acetic acid, 11.33; HMF, 0.18; Furfural, 1.4; Total XAD, XAD-ED, ED, ED-XAD Kundu et al. (2015)
phenolic compound, 4.53
Sugarcane bagasse Acetic acid, 2.31; HMF, 0.02; Furfural, 0.19; Total Ion exchange resin (A-860S; A500PS and C- Arruda et al. (2011)
phenolic compound, 5.48 150-Purolite
Brewery’s spent grain Acetic acid, 1.31; HMF, 0.05; Furfural, 0.64; Total pH adjustment to 5.5, overliming, adsorption Carvalheiro et al. (2005)
phenolic compound, 1.32; Formic acid, 0.81; Levulinic with activated charcoal, Anion exchanger pH
acid, 0.16 5.5
Anion exchanger pH 0.77
Cation exchanger pH 5.5
Cation exchanger pH 0

ED-Electrodialysis.

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
6 L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx

sugar rejection of >97% but with the low transmission in inhibitors. Zhang et al. (2013) performed detoxification of oil palm empty
The membranes of NF chiefly NF270, NF- and NF245 (Dow) and DK fruit bunch hydrolysate with a newly isolated Enterobacter sp. FDS8
(GE Osmonics) were suitable with >94% glucose rejection and >80% and reported the depletion of furfural and HMF at a rate of up to
inhibitors transmission. 0.54 g/L/h and 0.12 g/L/h with total sugar loss of <5%. It is even pos-
Brás et al. (2014) used nanofiltration under diafiltration mode sible to recover the cells and used again for at least 5 times without
for the removal of inhibitory compounds from diluted acid hydro- losing their ability of detoxification. Biological detoxification is
lysate of olive pomace substrate to improve its xylose fermentabil- performed in single step without any additional manipulation,
ity. They also used the membranes NF270 and NF90 to explore the and without volume loss. It also offers numerous other benefits
best membrane suitable for detoxification. NF270 was chosen to be like complete reduction of furfural, acetic acid, and guaiacol over
used in the diananofiltration experiment as it showed the lowest chemical detoxification.
rejection for toxic compounds and highest permeate flux and can The enzyme mediated detoxification involves the enzymes lac-
deplete 99.7% of acetic acid and 100% of formic acid and furfural. cases and peroxidises. Lee et al. (2012) was the first to report a
novel laccase from the yeast, Yarrowia lipolytica, which demon-
5.3. Vacuum membrane distillation strates a higher efficiency in the reduction of non-phenolic and
phenolic components than any previously reported enzyme from
Chen et al. (2013) studied the efficiency of vacuum membrane bacteria or fungi.
distillation (VMD), a membrane separation technology which is
an increasingly popular and cost-effective in recovering inhibitors. 6. Enzymatic hydrolysis by hemicellulases
It works on the principle of liquid–vapor phase equilibrium that
controls the process selectivity, facilitating the removal of volatile Bioconversion of hemicellulose into monomeric sugars by the
components. VMD was used to eliminate two inhibitors acetic acid action of xylanases is promising, economical but time consuming.
and furfural from lignocellulosic hydrolysates. More than 98% of The complex structure of xylan needs different enzymes for its
furfural could be removed by VMD under optimal conditions but, complete hydrolysis. Endo-1, 4-b-xylanases (1, 4-b-D-
the removal of acetic acid was considerably lower when compared xylanxylanohydrolase, E.C.3.2.1.8) depolymerise xylan by the
with furfural. random hydrolysis of xylan backbone and 1,4-b-D-xylosidases (1,4,
b-D-xylan xylohydrolase E.C.3.2.1.37) split off small oligosaccha-
5.4. Electrochemical detoxification rides. List of different pretreated lignocellulosic materials
hydrolyzed by xylanases is depicted in Table 2. Though these organ-
Electrochemical detoxification is used for the removal of pheno- isms produce high yields of enzyme, there is still a need to improve
lic compounds, p-coumaric acid, ferulic acid, syringaldehyde, and the production of xylanases to make the enzymatic conversion of
vanillin. The phenolic compounds were converted electrochemi- hemicellulose to xylitol more economic.
cally to a radical form at oxidation potential and are then removed Microbial hemicellulases for biomass hydrolysis are cost inten-
after radical polymerization as polymeric compounds. Nearly 71% sive. Mostly xylanases need to be either modified genetically or
of total phenolics were removed from rice straw hydrolysate with- immobilized to meet the requirements like activity, substrate
out any sugar-loss (Lee et al., 2015). specificity, enantioselectivity, sterospecificity, stability and toler-
ance to toxic reagents and extreme conditions. Overexpression of
5.5. Biological detoxification these enzymes has been attained in many studies. Some of the
recombinant xylanases with heterologous expression are summa-
There is increasing interest on biological mode of detoxification rized in Table 3. The over expressed enzyme in majority of the
because of the formation of low side reaction products, requires reports obtained increased thermal stability up to 80–90 °C. When
less energy without the use of chemicals (Parawira and Tekere, the recombinant C. thermocellum GH43 hemicellulase genes
2011). For the proficient microbial detoxification to occur, there expressed in E. coli the enzyme production has been enhanced
should be a focus on the process parameters like optimal nutrient from 1.6 to 2.9 fold in repetitive batch mode of fermentation
addition, pH 4–6, temperature 25–50 °C, incubation time (Das et al., 2013).
12–144 h, size of inoculum 1–10% v/v or 0.5–10 g/L (dry weight)) Hydrogels comprising polyanion xylan and polycation chitosan
and the strain employed in carrying out detoxification. The use of were used as a carrier for immobilization of the endo-xylanase
an efficient Coniochaeta ligniaria NRRL30616 metabolizes furfural, from Trichoderma viridae. Because of the induced electrostatic
HMF, and other inhibitory compounds present in the dilute acid interaction between the enzyme and gel, the immobilized endo-
hydrolysate (Nichols et al., 2008). xylanase was active at 98 °C while the free enzyme was active only

Table 2
List of saccharification of different lignocellulosic substrates by xylanase Zhang and Sang (2015).

Enzymes sources Pretreatment conditions Saccharification conditions Substrate Reducing sugar Xylose (mg/g dry
(mg/g dry substrate)
substrate)
Penicillium chrysogenum Aqueous ammonia solution (15%), Substrate loading (2.0%, w/v), Corn cob powder 553.94 236.63
QML-2 60 °C 50 °C, pH 4.8, 24 h
Aspergillus foetidus MTCC Ammonia solution (15%), 121 °C, Substrate loading (2.5%, w/v), Wheat straw 196 –
4898 60 min 45 °C, pH 5.3, 20–24 h Rice straw 188 –
Corn cob 184 –
Penicillium sp. ECU0913 Steam-pretreated at 200 °C, 2.0 Substrate loading (10%, w/v) Corn stover powder – 139.33
MPa, 4 min
Trichoderma reesei Sodium hydroxide solution (1.0%), Room temperature, 60 min. Corn Stover powder 467 156.91
Substrate loading (1.0%, w/v), pH 4.8,
50 °C, 120 h
Aspergillus candidus Ammonium solution (15%) 121 °C, 60 min. Substrate loading Corn cob powder 438.47 190.66
(0.5%, w/v), pH 5.0, 43 °C, 7.12 h

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx 7

Table 3 maximal xylitol yield of 0.64 g/g and productivity of 0.43 g/l/h
Patents on expression of xylanases in different hosts (Juturu and Wu, 2012). were obtained at 13.68 KLa.
Source of gene Enzyme Expression host
Bacillus sp. NG-27 Endo Chloroplasts of tobacco 7.2. Cell recycling
xylanase A plant
T. reesei Endo H. insolens Cunha et al. (2007) investigated the conversion of xylose to xyl-
xylanase
N. frontalis Endo E. coli, P. Methanolica
itol by Candida guillermondii cells entrapped within poly vinyl alco-
N. patriciarum xylanase hol hydrogel by repeated batch method with cell recycling. The
Mixed population of rumen Endo- E. coli BL21(DE3) yeast cells were immobilized on to the support using freeze thaw
microorganism xylanase method, and fermentation was carried out in a reactor at 30 °C,
Trichoderma reesei Endo H. polymorpha
400 rpm at air flow rate of 1.04 vvm for 6 cycles. It was found that
xylanase
Alicyclobacillus acidocaldarius b xylosidase E. coli, P. pastoris the biocatalytic activity of cells retained till the 6th cycle but the
Selenomonas ruminantium GA192 b xylosidase E. coli conversion performance was slightly decreased in the 5th cycle.
Aspergillus niger b xylosidase H. polymorpha The xylitol production, volumetric productivity and yield increased
with continuous recycling of cells and reached to a maximum of
39.7 g/L, 0.53 g/L/h and 0.77 g/g after third cell cycling.
Kwon et al. (2006) reported a maximal volumetric productivity
up to 45 °C. b-Xylosidase was extracted from Aspergillus niger and of 3.5 g/L/h from an initial xylose concentration of 200 g/L by an
30% of the enzyme was immobilized on a polyamide membrane osmophilic C. tropicalis KCTC 10457. In a fed-batch fermentation
by adsorption. The immobilized enzyme showed only 6.8% of the of xylose and glucose, xylitol production of 234 g/L was obtained
maximal activity of the free enzyme but has a better thermal sta- in 48 h, with a production rate of 4.88 g/L/h utilizing 260 g/L
bility than the free enzyme. The immobilized enzyme was still xylose. This strain produced moderate xylitol of 2.07 g/L/h even
active after 20 cycles of usage for xylan hydrolysis releasing xylose at a very high xylose concentration of 350 g/L due to its high osmo-
at 1.7 g/L enzyme (Delcheva et al., 2008). tolerance. To increase the productivity of xylitol, cell recycling was
carried out for 10 cycles during which an average xylitol concen-
7. Biotechnological production of xylitol tration of 180 g/L was produced with a productivity and yield of
8.5 g/l/h and 0.84 g/g respectively for each recycle. When recycling
The microbial xylitol production has been extensively studied was carried out using two fold concentrated cell mass, average xyl-
as an alternative to the chemical process as it does not need exten- itol concentration, productivity and yield for each recycle was
sive steps for purification of xylose and operated at low cost. found to be 182 g/L, 12 g/L/h and 0.85 g/g respectively.
Microorganisms like bacteria, yeasts and fungi possess the ability
of fermenting commercial xylose or xylose present in the hydroly- 8. Fermentation parameters influencing xylitol production
sate derived from lignocellulosic residues to xylitol.
Xylose metabolism occurs in two steps. Xylose is reduced to Microbial production of xylitol is influenced by various factors.
xylitol in the presence of an enzyme xylose reductase (XR), which Hence optimization of these factors is necessary to obtain efficient
is either secreted from the cell or oxidized to xylulose by xylitol xylitol production and it depends upon the specificity of the
dehydrogenase (XDH). The above two reactions are considered to microorganism used.
be rate-limiting steps in xylitol production. Further by the enzyme
xylulokinase (XKS), xylulose is phosphorylated to xylulose 8.1. Carbon source
5-phosphate (EC 2.7.1.17) and catabolized by entering to pentose
phosphate, glycolytic, or phosphoketolase pathways. Xylitol pro- Lignocellulosic materials are cheap, renewable and abundantly
duction is extensively carried out using free cells or immobilized available promising carbon sources. Upon hydrolysis of hemicellu-
cells in batch, continuous or fed-batch process using synthetic losic fraction of lignocellulosic biomass xylose is released, which is
media or lignocellulosic substrates (Arruda et al., 2011). further fermented to xylitol. Corncobs, rice straw, rice husk, sugar-
cane bagasse, sorghum straw, sawdust and oat hulls are generally
7.1. Fed batch fermentation used as lignocellulosic substrates.
Misra et al. (2013) used corncobs as potential substrate for xyl-
Francisco et al. (2015) carried out simultaneous production of itol production by C. tropicalis. A maximum of 20.92 g/L xylose was
ethanol and xylitol under different operation bioreactor modes obtained by acid hydrolysis. The hydrolysate was further concen-
using co-cultures of Candida tropicalis IEC5-ITV and Saccharomyces trated by rotavapor that increased the amount of xylose to
cerevisiae ITV01-RD in sugarcane bagasse hydrolysates. Sequential 40.16 g/L that resulted in the xylitol production of 22.63 g/L and
cultivation using simultaneous co-culture of S. erevisiae and yield of 0.57 g/g. But, by microwave assisted concentration of the
C. tropicalis, the productivity and yield for xylitol were found to hydrolysate, the amount of xylose increased to 52.71 g/L resulted
be xylitol, QXylOH = 0.10 g/L/h and YXylOH/S = 0.31 g/g, respectively. in the xylitol production of 15.19 g/L and yield of 0.31 g/g.
Whereas by using fed-batch culture, the productivity and yield of
xylitol were found to be QxytOH = 0.27 g/L/h and YxytOH/s = 0.57 g/g 8.2. Nitrogen source
respectively. Continuous multistep production showed a 50%
increase than in fed-batch culture, where the productivity and Several organic nitrogen sources such as yeast, peptone, urea,
yield of xylitol were found to be QxytOH = 0.26 g/L/h and YxytOH/s casamino acid and inorganic sources like ammonium sulfate,
= 0.62 g/g, respectively. Himabindu and Gummadi (2015) carried ammonium nitrate, and ammonium chloride are generally used
out fed batch fermentation to increase the yield, productivity and in the production of xylitol. Organic nitrogen sources are more
to determine the effect of volumetric oxygen transfer coefficient beneficial than inorganic nitrogen sources and show a greater
(KLa) on xylitol production by Debaromyces nepalensis. By varying impact on the cell growth and xylitol production. Zhang et al.
oxygen transfer coefficients in the range of 12–39.6 h1and provid- (2012) investigated the effect of nitrogen sources on xylitol pro-
ing xylose and nitrogen source to the medium intermittently, duction from horticultural waste hemicellulosic hydrolysate using

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
8 L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx

Fig. 3. Effect of different nitrogen sources on xylitol yield.

the strain Candida athensensis SB18. The YNB-YE synthetic medium a higher temperature optimum (40–45 °C) because of acceleration
was supplemented with different nitrogen sources like ammonium of their enzymes at high temperature. It was observed that, conver-
sulfate, ammonium nitrate, urea and casaminoacid. Among which, sion of xylose to xylitol by Candida sp. B-22 was found to be con-
the addition of urea enhanced the conversion of xylose with a stant in the temperature range of 35–40 °C. Whereas at 45 °C and
higher xylitol yield of 0.75 g g1. These results were similar and higher, there was a great decline in the xylitol yield, may be due
consistent when compared with Rao et al. (2006) where urea to the inactivation of NADPH and NADH dependent XR
was found to be beneficial for maximum xylitol yield of 0.87 g/g (Ur-Rehman et al., 2012).
when compared with other different nitrogen sources like casein Sirisansaneeyakul et al. (2013) evaluated the xylitol production
hydrolysate, yeast extract, urea, sodium nitrate, ammonium chlo- by Candida magnolia TISTR 5663 under oxygen limited conditions.
ride and ammonium sulfate (Fig. 3). Xylitol yield of 0.72 g/g was obtained utilizing the biomass of
50 g/L at pH 7.0.
8.3. Aeration
8.5. Co-substrate and oxygen transfer
The availability of oxygen, which is often expressed in terms of
volumetric oxygen transfer coefficient (KLa) and oxygen transfer Hernández-Pérez et al. (2016) demonstrated the effect of
rate (OTR), is an important process parameter that plays a major various sugars as co-substrate for xylitol production by Candida
role in regulating cell growth and formation of xylitol by affecting guilliermondii FTI20037, by supplementing sugar cane bagasse
the activity of xylose reductase and xylitol dehydrogenase (Branco hydrolysate containing 57 g/L xylose with sugars such as maltose,
et al., 2008). To study the influence of KLa on XR and XDH, batch sucrose and glycerol. It was observed that the addition of 10 g/L
fermentation under different initial KLa values was carried out. sucrose and 0.7 g/L glycerol as co-substrate led to the significant
Generally to maintain micro-aerobic conditions, low oxygen trans- increase in the xylose uptake by 8.8% and 6.8% respectively. But
fer volumetric coefficient (KLa) values are used. However Branco addition of sucrose alone increased the xylitol concentration to
et al. (2008) reported that high KLa values resulted in high xylitol 36 g/L. This study makes the process economical by using sucrose
production. By evaluating the effect of aeration on XR and XDH derived from lignocellulosic substrates.
activities, it was found that at lowest KLa value of 12 h1, maxi- To avoid the requirement of co-substrate for co-factor regener-
mum XR specific activity of 1.45 ± 0.21 U/mg protein was obtained, ation and cell growth, Ko et al. (2011) followed a new method of
whereas the maximum XD specific activity of 0.19 ± 0.03 U/mg decreasing the XDH activity by engineering the attenuation of
protein was obtained at a high KLa value of 25 h1. Further when XDH of C. tropicalis strain, by reducing its gene copy number. Stud-
KLa was increased from 12 to 50 h1, there was a greater increase ies on xylitol production by mutants, strain XDH-5 (with only one
in xylitol production with a volumetric productivity of copy of the XDH gene), and ARSdR-16 (with a mutated XDH gene)
1.50 ± 0.08 g/L/h and efficiency of 71%. showed 70% and 40% of wild type (WT) XDH activity respectively.
To evaluate the influence of different KLa values on xylitol pro- Xylose to xylitol conversion yield by wild type and mutant strains
duction by C. tropicalis W103, Cheng et al. (2010) carried out batch XDH-5 and ARSdR-16 were 62%, 64%, and 75%, respectively, with
fermentation and observed that at lower KLa of 10.8 h1, the bio- productivity of 0.52, 0.54, and 0.62 g/L/h respectively. ARSdR-16
mass was less with low xylitol productivity but, as KLa increased mutant showed high xylitol yield and productivity without using
from 10.8 h1 to 16.5 h1, biomass is increased with increased xyl- any co-substrate.
itol yield. Further increase of KLa to 18.3 h1 led to the increase of
biomass but decrease in xylitol yield, may be due to increased XRD 9. Continuous fermentation
activity which converts xylitol to xylulose. The maximum xylitol
yield of 0.73 g g1 and productivity of 1.43 g L1 h1 were obtained Continuous fermentation maintains high productivities for
at KLa 16.5 h1 and 18.3 h1 respectively. longer periods of time by eliminating the idle time for cleaning
and sterilization. It shows greater steadiness in product synthesis,
8.4. Effect of temperature and pH on xylitol production hence several bioreactors have been used for continuous produc-
tion of xylitol.
Most of the yeasts grow at 30 °C to 37 °C for biotechnological However, immobilized cells are generally used during continu-
xylitol production. However, some thermotolerant yeasts possess ous fermentation as they possess high cell density, improved

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx 9

stability with higher productivity and can be re-used when com- Qi et al. (2016) conducted a study in which xdh gene of Glu-
pared to free cells. conobacter oxydans CGMCC 1.49 was cloned and expressed in
A three-phase fluidized-bed bioreactor using immobilized Can- E. coli BL21 successfully, and recombinant XDH was characterized.
dida sp. ZU04 cells was built by Ding (2011) for efficient continu- The results showed that, in the mixed culture of resting cells, the
ous production of xylitol with high yield from corn cob mixed strains (G. oxydans and recombinant strain BL21-xdh)
hemicellulosic hydrolysate. It was found that at high aeration rate improved the yield of xylitol from 0.297 to 0.960 g/g using 30 g/L
of 1.00 vvm, within first 24 h, the glucose was rapidly consumed D-xylulose as substrate.
and biomass in alginate beads was quickly increased. But when Xylose from mother liquor was converted to xylitol by using a
the aeration rate was decreased to 0.30 vvm in the second phase, novel strategy of combining Bacillus subtilis and Candida maltosa.
xylitol yield increased remarkably. The immobilized cells were In this process C. maltosa was used to detoxify the inhibitors like
reused for further batches leading to the xylitol yield of 0.73 g/g furfural and 5-HMF present in the mother liquor by utilizing the
and productivity of 0.84 g/L/h. glucose that causes catabolite repression in B. subtilis. Further
Sarrouh et al. (2010) determined hydrodynamic characteristics xylose isomerase gene disrupted B. subtilis was constructed for uti-
and fermentation parameters of a bench-scale three-phase flu- lizing the arabinose, thereby increasing the concentration of xylose
idized bed reactor (FBR) for xylitol production using calcium algi- in the media. Finally C. maltosa is reused for the reduction of xylose
nate immobilized C. guilliermondii. When the affect of total to xylitol with a yield and productivity of 0.85 g/g and 4.25 g/L/h
particle density, terminal velocity, particles drag force, minimum respectively (Cheng et al., 2011).
fluidization velocity and bed porosity (n) on cell immobilization
was evaluated, it was found that the reactor operates similar to a
fixed bed reactor at porosity <0.5 and as fluidized bed reactor at 11. Strain Improvement strategies
>0.5 porosity. It is observed that maximum bed fluidization in
the reactor can be obtained when the flow rate is equal to the ter- Strain improvement methods such as mutation, adaptation and
minal velocity of immobilized cells. When the effects of the aera- recombination are generally employed to enhance the perfor-
tion rate and fermentation time on the process productivity and mance of microorganisms used in fermentation process. To evalu-
yield were analyzed, a high aeration rate of 600 ml/min favors ate xylitol production capabilities of K. marxianus, Kim et al. (2015)
better oxygen transfer into the immobilized cells and lead to the carried out random mutagenesis and selected seven K. marxianus
maximum xylitol productivity, yield and xylitol concentration of strains (ATCC 124224, ATCC 46537, ATCC 200963, ATCC 36907,
0.41 g/L/h, 0.582 g /g and 28.9 g/L respectively at 70 h of ATCC 26548, CBS 1555, and CBS 607). Among seven strains,
fermentation. K. marxianus ATCC 36907 produced maximum amount of xylitol
Comparison of regeneration and bioconversion efficiencies of 24 g/L with a yield and productivity of 0.48 g/g and 0.17 g/L/h
free and immobilized cells of C. guilliermondii FTI 20037 on xylitol respectively from 50 g/L xylose at 30 °C for 144 h, indicating that
production from oat hull hemicellulosic hydrolysate was evaluated XR from ATCC 36907 may possess higher activity than
and found that the aeration rate or oxygen mass transfer coeffi- other strains. Further mutagenesis of K. marxianus 36907 lead to
cient KLa plays a crucial role on xylitol production. When fermen- K. marxianus 36907-FMEL1 strain which exhibited more than two
tation of hydrolysate containing 74.5 g/L xylose was carried out fold improvement in xylitol production which corresponds to
using free cells of Candida at an aeration rate of 1.25 vvm or KLa 0.67 g/g xylitol yield and productivity of 0.36 g/L/h. Xylitol concen-
of 15.8 h1, a maximum xylitol yield of 0.87 g/g and production tration and yield from K. marxianus 36907-FMEL1 were 120% and
of 55 g/L was obtained which is almost similar to the fermentation 39% more than the parental strain, K. marxianus ATCC 36907,
carried out by immobilized cells with aeration rate of 1.25–1.5 vvm respectively. The reason being, XR activity of the mutant is twofold
(Soleimani and Tabil, 2014). higher than its parent and genes sequencing of its KmXYL1 gene
show that cysteine was substituted to tyrosine at position 36 after
strain development which might cause enhanced XR activity in
10. Molecular approaches for xylitol production K. marxianus 36907-FMEL1.
Mahmud et al., 2013 disrupted the xdhA and ladA genes of
Xylose reductase (XR) is considered as a key enzyme in biotech- Aspergillus oryzae which encode XDH and LAD respectively to
nological xylitol production. In recent decades, various metabolic increase xylitol production. The xdhA- and ladA-disrupted mutants
engineering strategies have been explored to modify the XR and were constructed by homologous transformation into A. oryzae P5
XDH in microbes to increase the production of xylitol. (DpyrG) and pyrG was used as a selectable marker. The mutants
In a study conducted by Jo et al. (2015), NADPH-dependent XR were grown on different carbohydrate-containing media, colony
and NADH-preferring mutant XR along with co expression of ZWF1 diameters of mutants were measured, and gene disruption was
for NADPH regeneration and ACS1 for NADH regeneration was confirmed by PCR. The xylitol productivity of the mutants was
engineered into S. cerevisiae that resulted in the xylitol productivity studied using D-xylose and oat spelt xylan as carbon sources. The
of 4.27 g/l/h and a yield of 1.0 g/g. The XR gene from N. crassa was xdhA-disrupted mutant (xdhA2-1) produced 16.6 g/L xylitol with
codon optimized and expressed in XYL2 deficient C. tropicalis a yield of 0.43 g/g D-xylose and productivity of 0.248 g/L/h from
mutant. This engineered C. tropicalis was able to metabolize both D-xylose, while 10.2 g/L xylitol was produced at a yield of
glucose and xylose that resulted in higher xylitol yield of 0.96 g/g 0.204 g/g from oat spelt xylan.
than without the expression of XR gene from N. crassa. Adaptation is an inexpensive approach used as an alternative to
The disruption of Glucose specific phospho transferase system the detoxification methods which adapts the microbes to inhibi-
(ptsG) which is phosphoenol dependent resulted in the elimination tory compounds. Fermentative efficiency for xylitol production
of catabolite repression and allowed immediate uptake of glucose can be improved by adapting the cells to the fermentation med-
and xylose. Deleting the genes xylA (xylose isomerase) and xylB ium. Misra et al. (2013) reported xylitol production of 7.46 g/l with
(xylulokinase) resulted in the blockage of xylose catabolic pathway a yield of 0.37 g/g by adapted C. tropicalis from 21.98 g/L of xylose
in E. coli W3110 to avoid xylose consumption. The phospho enol present in untreated corn cob hydrolysate. Further adaptation of C.
pyruvate-dependent fructose phospho transferase system (ptsF) tropicalis in optimized and treated corncob hydrolysate containing
was disrupted that consequently blocked the xylitol phosphoryla- 20.92 g/l xylose, found to produce 12.23 g/L of xylitol with a yield
tion pathway with a xylitol yield of 0.95 g/g (Su et al., 2015). of 0.61 g/g which is 1.22-fold higher compared to parent strain. In

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
10 L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx

comparison with the above study, Rao et al. (2006) reported an hydrolysate was decolourised and by-products such as acetic acid,
increase in xylitol yield from of 0.43 g/g to 0.58 g/g and 0.45 g/g lignin derivatives were removed. The clarified fermentation broth
to 0.65 g/g by adaptation of C. tropicalis cells for 20 cycles in neu- was concentrated 40 fold, and then the xylitol content in concen-
tralized, activated charcoal and ion exchange resin treated corn trated medium was 726.5 g dm3, a density of 1.3443 g cm3 and
fiber and sugarcane bagasse hydrolysates respectively. viscosity of 15cP. Further to promote xylitol crystallization, the
concentrated medium was cooled in the presence of alcohol as
the solubility of xylitol is decreased in the presence of ethanol.
12. Enzymatic approach for xylitol production
Finally to favor the nucleation pure xylitol crystal were seeded in
to crystallizer at a temperature below the saturation so that small
Enzymatic xylitol production technology is an alternative and
amount of crystals 3.3 g with a crystalline yield of 43.5% is
promising approach for conversion of xylose to xylitol. During
obtained which is close to the 51.4% obtained for commercial xyl-
enzymatic process, xylose was stoichiometrically converted to xyl-
itol solution.
itol with almost equivalent consumption of NADPH using NADP+
to xylose ratio of over 1:30, and the coenzyme was regenerated
and retained successfully using a membrane reactor. About 90% 14. Cost economics
conversion of xylose to xylitol could be achieved at 35 °C and pH
7.5 after a 24 h reaction period (Rafiqul and Sakinah, 2012). Xylitol production cost varies and depends mainly on cost of
To interpret the xylitol production by enzymatic process, raw material used and its transport cost which depends on feed-
Branco et al. (2011) used RSM, a statistical design which resulted stock mass and location of the manufacturing plant. The estimated
in quadratic model that could predict 98.6% volumetric productiv- cost for production of xylose from rye straw hemicellulose was
ity of xylitol with respect to xylose and NADPH concentration. observed to be €59.2 per tonne. This can be further broken down
A maximum xylitol volumetric productivity of 1.5 g/L/h was in to 30% for raw materials, utilities 24.4%, enzyme cost 6.5%,
obtained using the predicted experimental conditions of pH 7.0, 10.2% maintenance, 22.9% depreciation and 6.0% labor cost. This
temp 25 °C, 1.2 mM NADPH, 0.34 M xylose and glucose, 0.2 U/mg has to be further added up with fermentation and downstream
xylose reductase and 0.2 U/mg of xylitol dehydrogenase. The xyli- processing cost (Franceschin et al., 2011).
tol productivity of 1.58 g/L/h obtained by the statistical model was
higher and significant compared to the traditional mode. 15. Conclusions
Zhoua et al. (2012) investigated the biotransformation of
D-arabitol into xylitol with focus on the conversion of xylulose into The production of xylitol is receiving greater interest because of
xylitol by co expression of XDH gene from G. oxydans and co-factor its increasing demand in food and pharmacy industries. Efforts are
regenerating enzyme gene glucose dehydrogenase (GDH) from being made by several researchers to improve xylitol yield to make
B. subtilis in E. coli (system 1) and co-expression of alcohol dehy- the process economical and competitive. The focus should be more
drogenase gene ADH from G. oxydans and co-factor regenerating on novel strategies like genetic engineering for complete utiliza-
enzyme gene GDH from B. subtilis in E. coli (system 2). Both the tion of lignocellulosic substrates and also making organisms more
above systems successfully converted xylulose to xylitol without stress tolerant. Understanding the technology involved in the
addition of NADH. Approximately 92% conversion yield i.e., hydrolysis, detoxification and the conversion of wastes to xylitol
26.91 g/L xylitol was obtained from 30 g/L xylulose by system plays vital role in commercial production of xylitol. Economic fea-
1 and 24.9 g/L xylitol with 85.2% conversion yield by system 2. sibility of the process involved in the conversion of xylose to xylitol
The xylitol yields of both systems are 3 fold higher than G. oxydans has to be elucidated.
NH-10 cells (7.32 g/L).
Acknowledgements
13. Purification and recovery
Authors wish to acknowledge the financial support from UGC
Depending upon the type of lignocellulosic substrate used for DRS, SAP-I and UGC BSR Faculty fellowship.
fermentation, its acid hydrolysis process varies and generates dif-
ferent type of sugars and inhibitors making the purification and References
recovery process complicated. Crystallization is an energetically
advantageous process generally used to separate a single com- Abdullah, R., Ueda, K., Saka, S., 2014. Hydrothermal decomposition of various
crystalline celluloses as treated by semi-flow hot-compressed water. J. Wood
pound from a mixture of raw materials or reaction by-products Sci. 60 (4), 278–286.
(Misra et al., 2013). Arruda, P.V., Rodrigues, R.D.C.L.B., Silva, D.D.V., Felipe, M.D.G.D.A., 2011. Evaluation
Economical and environmental-friendly method for the purifi- of hexose and pentose in pre-cultivation of Candida guilliermondii on the key
enzymes for xylitol production in sugarcane hemicellulosic hydrolysate.
cation and crystallization of xylitol from corn cob hydrolysate Biodegradation 22 (8), 15–22.
was reported by Wei et al. (2010). Initially by using 1% activated Branco, R.F., Julio, C., Dos, Santos, Sarrouh, F. Boutros, Rivaldi, Juan D., Pessoa Jr.,
carbon at a temperature of 60 °C and 165 rpm, the xylitol fermen- Adalberto, da Silva, Silvio S., 2008. Profiles of xylose reductase, xylitol
dehydrogenase and xylitol production under different oxygen transfer
tation broth was decolourised and then desalted with a two ion volumetric coefficient values. J. Chem. Technol. Biotechnol. 84, 326–330.
exchange resins 732 and D301. Later the left over sugars in the Branco, R.F., Santos, J.C., Silva, S.S., 2011. A solid and robust model for xylitol
broth were separated with the resin UBK-555(Ca2+). Further the enzymatic production optimization. Bioproces. Biotech. 1, 4.
Brás, Teresa, Guerra, Vera, Torrado, Ivone, Lourenco, Pedro, Carvalheiro, Florbela,
broth was vacuum-concentrated up to xylitol super saturation of Luís, C., Duarte, Luísa, Neves, A., 2014. Detoxification of hemicellulosic
750 g/L. Finally to the clarified fermentation broth 1% xylitol crys- hydrolysates from extracted olive pomace by diananofiltration. Process.
tal seeds, were added and then the supersaturated solution was Biochem. 49, 173–180.
Canilha, Larissa, Carvalho, Walter, Giulietti, Marco, Das Grac, Maria, Felipe, Almeida,
cooled to 20 °C for 48 h to obtain crystalline xylitol of tetrahedral
Silva, Joao Batista Almeida E., 2008. Clarification of a wheat straw-derived
shape with 95% purity and crystallization yield of 60.2%. medium with ion-exchange resins for xylitol crystallization. J. Chem. Technol.
To purify and crystallize xylitol present in wheat straw hemicel- Biotechnol. 83, 715–721.
lulosic hydrolysate, Canilha et al. (2008), initially used two anion Carvalheiro, F., Duarte, L.C., Lopes, S., Parajó, J.C., Pereira, H., Girio, F.M., 2005.
Evaluation of the detoxification of brewery’s spent grain hydrolysate for xylitol
exchange resins A-860S and A-500PS to clarify the hemicellulosic production by Debaryomyces hansenii CCMI 941. Process Biochem. 40, 1215–
hydrolysate by sequential adsorption. By adsorption 95% of 1223.

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx 11

Castoldi, R., Bracht, A., Morais, G.R., Baesso, M.L., Correa, R.C.G.C., Peralta, R.A., of phenolic compounds in lignocellulosic hydrolysate for Clostridium
Moreieira, R.F.P.M., Polizeli, M.L.T.M., Souza, C.G.M., Peralta, R.M., 2014. fermentation. Bioresour. Technol. 187, 228–234.
Biological pretreatment of Eucalyptus grandis sawdust with white-rot fungi: Mahmud, Asif, Hattori, Koji, Hongwen, Chen, Kitamoto, Noriyuki, Suzuki, Tohru,
study of degradation patterns and saccharification kinetics. Chem. Eng. J. 258, Nakamura, Kohei, Takamizawa, Kazuhiro, 2013. Xylitol production by NADþ-
240–246. dependent xylitol dehydrogenase (xdhA)- and L-arabitol-4-dehydrogenase
Chandan, Kundu, Ly Thi Phi Trinh, Lee, Hong-Joo, Lee, Jae-Won, 2015. Bioethanol (ladA)-disrupted mutants of Aspergillus oryzae. J. Biosci. Bioeng. 115 (4), 353–
production from oxalic acid-pretreated biomass and hemicellulose-rich 359.
hydrolysates via a combined detoxification process. Fuel 161, 129–136. Martínez, M.L., Sancheza, S., Bravo, V., 2012. Production of xylitol and ethanol by
Chen, Jingwen, Zhang, Yaqin, Wang, Yafei, Ji, Xiaosheng, Zhang, Lin, Mi, Xigeng, Hansenula polymorpha from hydrolysates of sunflower stalks with
Huang, He, 2013. Removal of inhibitors from lignocellulosic hydrolyzates by phosphoricacid. Ind. Crop Prod. 40, 160–166.
vacuum membrane distillation. Bioresour. Technol. 144, 680–683. Misra, S., Raghuwanshi, S., Saxena, R.K., 2013. Evaluation of corncob hemicellulosic
Cheng, K.K., Ling, H.Z., Zhang, J.A., Ping, W.X., Huang, W., Ge, J.P., Xu, J.M., 2010. hydrolysate for xylitol production by adapted strain of Candida tropicalis.
Strain isolation and study on process parameters for xylose to xylitol Carbohydr. Polym. 92, 1596–1601.
bioconversion. Biotechnol. Biotech. Equip. 24 (1), 1606–1611. Mussatto, S.I., Roberto, I.C., 2004. Alternatives for detoxification of diluted-acid
Cheng, Hairong, Wang, Ben, Lv, Jiyang, Jiang, Mingguo, Lin, Shuangjun, Deng, Zixin, lignocellulosic hydrolyzates for use in fermentative processes: a review.
2011. Xylitol production from xylose mother liquor: a novel strategy that Bioresour. Technol. 93, 1–10.
0
combines the use of recombinant Bacillus subtilis and Candida maltosa. Microb. N diaye, S., Rigal, L., 2000. Factors influencing the alkaline extraction of poplar
Cell. Fact. 10, 5. hemicelluloses in a twin-screw reactor: correlation with specific mechanical
Cunha, M.A.A., Rodrigues, R.C.B., Santos, J.C., Converti, A., Da Silva, S.S., 2007. energy and residence time distribution of the liquid phase. Bioresour. Technol.
Repeated-batch xylitol bioproduction using yeast cells entrapped in polyvinyl 75, 13–18.
alcohol–hydrogel. Curr. Microbiol. 54, 91–96. Nguyen, N., Fargues, C., Guiga, W., Lameloise, M.L., 2015. Assessing nano filtration
Das, S.P., Ghosh, A., Gupta, A., Goyal, A., Das, D., 2013. Lignocellulosic fermentation and reverse osmosis for the detoxification of lignocellulosic hydrolysates. J.
of wild grass employing recombinant hydrolytic enzymes and fermentative Membr. Sci. 487, 40–50.
microbes with effective bioethanol recovery. BioMed. Res. Int., 14 Nichols, N.N., Sharma, L.N., Mowery, R.A., 2008. Fungal metabolism of fermentation
Delcheva, G., Dobrev, G., Pishtiyski, I., 2008. Performance of Aspergillus niger B 03 b- inhibitors present in corn stover dilute acid hydrolysate. Enzyme Microb.
xylosidase immobilized on polyamide membrane support. J. Mol. Catal. B Technol. 42, 624–630.
Enzyme 54, 109–115. Pan, X.J., Gilkes, N., Kadla, J., Pye, K., Saka, S., Gregg, D., Ehara, K., Xie, D., Lam, D.,
Francisco, C.R.J., Domínguez-González, José Manuel, Ortíz-Muñiz, Benigno, Saddler, J., 2006. Bioconversion of hybrid poplar to ethanol and co-products
Torrestiana-Sanchez, Beatriz, Ramírez de León, José Alberto, Aguilar-Uscanga, using an organosolv fractionation process: optimization of process yields.
María Guadalupe, 2015. Continuous multistep versus fed-batch production of Biotechnol. Bioeng. 94 (5), 851–861.
ethanol and xylitol in a simulated medium of sugarcane bagasse hydrolysates. Parawira, W., Tekere, M., 2011. Biotechnological strategies to overcome inhibitors in
Eng. Life. Sci. 15 (1), 96–107. lignocellulose hydrolysates for ethanol production: review. Crit. Rev.
de Souza Moretti, Marcia Maria, Bocchini-Martins, Daniela Alonso, da Costa Carreira Biotechnol. 31, 20–31.
Nunes, Christiane, Villena, Maria Arévalo, Micali Perrone, Olavo, da Silva, Qi, B., Chen, X., Shen, F., Su, Y., Wan, Y., 2009. Optimization of enzymatic hydrolysis
Roberto, Boscolo, Maurício, Gomes, Eleni, 2014. Pretreatment of sugarcane of wheat straw pretreated by alkaline peroxide using response surface
bagasse with microwaves irradiation and its effects on the structure and on methodology. Ind. Eng. Chem. Res. 48, 7346–7353.
enzymatic hydrolysis. Appl. Energy 122, 189–195. Qi, Xiang Hui, Zhu, Jing Fei, Yun, Jun Hua, Lin, Jing, Qi, Yi Lin, Guo, Qi, Hong, Xu,
Ding, Xinghong, 2011. Fermentation of xylitol using immobilized Candida sp. ZU04 2016. Enhanced xylitol production: expression of xylitol dehydrogenase from
cells in three-phase fluidized-bed bioreactor. In: Remote Sensing, Environment Gluconobacter oxydans and mixed culture of resting cell. J. Biosci. Bioeng. 1–6.
and Transportation Engineering (RSETE), 2011 International Conference on, http://dx.doi.org/10.1016/j.jbiosc.2016.02.009.
Nanjing, pp. 7591–7593. http://dx.doi.org/10.1109/RSETE.2011.5966129. Qin, L., Liu, Z.H., Li, B.Z., Dale, B.E., Yuan, Y.J., 2012. Mass balance and transformation
Franceschin, G., Sudiro, M., Ingram, T., Smirnova, I., Brunner, G., Bertucco, A., 2011. of corn stover by pretreatment with different dilute organic acids. Bioresour.
Conversion of rye straw in to fuel and xylitol: a technical and economical Technol. 112, 319–326.
assessment based on experimental data. Chem. Eng. Res. Des. 89, 631–640. Rafiqul, I.S.M., Sakinah, A.M.M., 2012. Perspective bioproduction of xylitol by
Garcia Cubero, M.A., Gonzalez Benito, G., Indacoechea, I., Coca, M., Bolado, S., 2009. enzyme technology and future prospects Int. Food Res. J. 19 (2), 405–408.
Effect of ozonolysis pretreatment on enzymatic digestibility of wheat and rye Rafiqul, S.M., Sakinah, A.M.M., Karim, M.R., 2014. Production of Xylose from meranti
straw. Bioresour. Technol. 100 (4), 1608–1613. wood sawdust by dilute acid hydrolysis. Appl. Biochem. Biotechnol. 174, 542–
Garcia Martin, J.F., Sanchez, S., Cuevas, M., 2013. Evaluation of the effect of the 555.
dilute acid hydrolysis on sugars release from olive prunings. Renew. Energy 51, Rao, R.S., Jyothi, P.C., Prakasham, R.S., Sarma, P.N., Rao, V.L., 2006. Xylitol production
382–387. from corn fibre and sugarcane bagasse hydrolysates by Candida tropicalis.
Hernández-Pérez, A.F., Costa, I.A.L., Silva, D.D.V., Dussán, Villela, T.R., Canettieri, E.V., Bioresour. Technol. 97, 1974–1978.
Carvalho, J.A., Soares Neto, T.G., M.G.A., Felipe, 2016. Biochemical conversion of Rao, R.S., Gallagher, Joe, Fish, Steve, Prakasham, Reddy Shetty, 2012. Overview on
sugarcane straw hemicellulosic hydrolysate supplemented with co-substrates Commercial Production of Xylitol, Economic Analysis and Market Trends. D-
for xylitol production. Bioresour. Technol. 200, 1085–1088. Xylitol Fermentative Production, Application and Commercialization. Springer,
Himabindu, K., Gummadi, S.N., 2015. Effect of kLa and Fed-batch strategies for Berlin Heidelberg, pp. 291–306. http://dx.doi.org/10.1007/978-3-642-31887-0-
enhanced production of xylitol by Debaryomyces nepalensis NCYC 3413. Br. 13.
Biotechnol. J. 5 (1), 24–36. Saleh, Marwa, Cuevas, M., Garcia, J.F, Sanchez, S., 2014. Valorization of olive stones
http://www.marketresearch.com/product/sample-8164119.pdf. for xylitol and ethanol production from dilute acid pretreatment via enzymatic
http://www.prnewswire.com/. hydrolysis and fermentation by Pachysolentannophilus. Biochem. Eng. J. 90, 286–
Jo, J.H., Oh, S.Y., Lee, H.S., Park, Y.C., Seo, J.H., 2015. Dual utilization of NADPH and 293.
NADH cofactors enhances xylitol production in engineered Saccharomyces Sarrouh, Boutros F., Silvério da Silva, Silvio, 2010. Evaluation of the performance of a
cerevisiae. Biotechnol. J. http://dx.doi.org/10.1002/biot.201500068. three-phase fluidized bed reactor with immobilized yeast cells for the
Juturu, Veeresh, Wu, Jin Chuan, 2012. Microbial xylanases: engineering, production biotechnological production of xylitol. Int. J. Chem. React. Eng. 6 (1), 1542–
and industrial applications. Biotechnol. Adv. 30, 1219–1227. 6580.
Kim, Jin.-Seong, Park, Jae.-Bum, Jang, Seung.-Won, Ha, Suk.-Jin, 2015. Enhanced Sirisansaneeyakul, S., Wannawilai, S., Chisti, Y., 2013. Repeated fed-batch
xylitol Production by mutant Kluyveromyces marxianus 36907-FMEL1 due to production of xylitol by Candida magnolia TISTR 5663. J. Chem. Technol.
improved xylose reductase activity. Appl. Biochem. Biotechnol. 176, 1975–1984. Biotechnol. 88, 1121–1129.
Ko, Byoung Sam, Kim, Dong.-Min, Yoon, Byoung Hoon, Bai, Suk, Kim, Hyeon Yong Siti, M., Kamal, Mustapa, Nurul Mohamad, L., Abdul, G., Abdullah, Liew, Abdullah,
Lee Jung Hoe, Kim, Il.-Chul., 2011. Enhancement of xylitol production by Norhafizal, 2011. Detoxification of sago trunk hydrolysate using activated
attenuation of intracellular xylitol dehydrogenase activity in Candida tropicalis. charcoal for xylitol production. Proc. Food Sci. 1, 908–913.
Biotechnol. Lett. 33, 1209–1213. Soleimani, Majid, Tabil, Lope, 2014. Evaluation of biocomposite-based supports for
Kumar, P., Barrett, D.M., Delwiche, M.J., Stroeve, P., 2009. Methods for pretreatment immobilized-cell xylitol production compared with a free-cell system. Biochem.
of lignocellulosic biomass for efficient hydrolysis and biofuel production. Ind. Eng. J. 82, 166–173.
Eng. Chem. Res. 48, 3713–3729. Su, B., Wu, M., Zhang, Z., Lin, J., Yang, L., 2015. Efficient production of xylitol from
Kumar, L., Arantes, V., Chandra, R., Saddler, J., 2012. The lignin present in steam hemicellulosic hydrolysate using engineered Escherichia coli. Metab. Eng. 31,
pretreated softwood binds enzymes and limits cellulose accessibility. Bioresour. 112–122.
Technol. 103 (1), 201–208. Swain, Manas R., Krishnan, Chandraj, 2015. Improved conversion of rice straw to
Kwon, S.G., Park, S.W., Oh, D.K., 2006. Increase of xylitol productivity by cell-recycle ethanol and xylitol by combination of moderate temperature and ammonia
fermentation of Candida tropicalis using submerged membrane bioreactor. J. pretreatment and sequential fermentation using Candida tropicalis. Ind. Crop
Biosci. Bioeng. 101 (1), 13–18. Prod. 77, 1039–1046.
Lee, K.M., Kalyani, D., Tiwari, M.K., Kim, T.S., Dhiman, S.S., Lee, J.K., 2012. Enhanced Taherzadeh, M.J., Karimi, K., 2008. Pretreatment of lignocellulosic wastes to
enzymatic hydrolysis of rice straw by removal of phenolic compounds using a improve ethanol and biogas production: a review. Int. J. Mol. Sci. 9, 1621–1651.
novel laccase from yeast Yarrowia lipolytica. Bioresour. Technol. 123, 636–645. Ur-Rehman, Salim, Mushtaq, Zarina, Zahoor, Tahir, Jamil, Amir, Murtaza, Mian
Lee, Kyung Min, Min, Kyoungseon, Choi, Okkyoung, Kim, Ki.-Yeon, Woo, Han Min, Anjum, 2012. Xylitol: a review on bioproduction, application, health benefits,
Kim, Yunje, Han, Sung Ok, Um, Youngsoon, 2015. Electrochemical detoxification and related safety issues. Crit. Rev. Food Sci. 55 (11), 1514–1528.

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
12 L. Venkateswar Rao et al. / Bioresource Technology xxx (2016) xxx–xxx

Wang, Ying, Yuan, Bo, Yingchao, Ji, Hong, Li, 2013. Hydrolysis of hemicellulose to Zhang, Jinming, Geng, Anli, Yao, Chuanyi, Yinghua, Lu, Li, Qingbiao, 2012. Xylitol
produce fermentable mono saccharides by plasma acid. Carbohydr. Polym. 97, production from D-xylose and horticultural waste hemicellulosic hydrolysate
518–522. by a new isolate of Candida athensensis SB18. Bioresour. Technol. 105,
Wei, Jinchao, Yuan, Qipeng, Wang, Tianxin, Wang, Le, 2010. Purification and 134–141.
crystallization of xylitol from fermentation broth of corncob hydrolysates. Zhang, Dongxu, Onga, Yee Ling, Li, Zhi, Jin Chuan, Wu, 2013. Biological
Front. Chem. Eng. China 4 (1), 57–64. detoxification of furfural and 5-hydroxyl methyl furfural in hydrolysate of oil
Weng, Y.H., Wei, H.J., Tsai, T.Y., Chen, W.H., Wei, T.Y., Hwang, W.S., Wang, C.P., palm empty fruit bunch by Enterobacter sp. FDS8. Biochem. Eng. J. 72, 77–82.
Huang, C.P., 2009. Separation of acetic acid from xylose by nanofiltration. Sep. Zhong, Chao, Chunming, Wang, Fan, Huang, Fengxue, Wang, Honghua, Jia, Hua,
Purif. Technol. 67, 95–102. Zhou, Ping, Wei, 2015. Selective hydrolysis of hemicellulose from wheat straw
Xu, F., Sun, J.X., Geng, Z.C., Liu, C.F., Sun, R.C., Fowler, P., 2007. Comparative study of by a nanoscale solid acid catalyst. Carbohyd. Polym. 131, 384–391.
water-soluble and alkali-soluble hemicelluloses from perennial ryegrass leaves Zhoua, Peng, Li, Sha, Xua, Hong, Fenga, Xiaohai, Ouyang, Pingkai, 2012. Construction
(Lolium peree). Carbohydr. Polym. 67, 56–65. and co-expression of plasmid encoding xylitol dehydrogenase and a cofactor
Yang, B., Wyman, C.E., 2008. Pretreatment: the key to unlocking low-cost cellulosic regeneration enzyme for the production of xylitol from d-arabitol. Enzyme
ethanol. Biofuel. Bioprod. Biorefin. 2, 26–40. Microb. Technol. 51, 119–124.
Zahed, Omid, Jouzani, Gholamreza Salehi, Abbasalizadeh, Saeed, Khodaiyan, Zhu, Junjun, Yong, Qiang, Yong, Xu, Shiyuan, Yu, 2011. Detoxification of corn stover
Faramarz, Tabatabaei, Meisam, 2015. Continuous co-production of ethanol prehydrolyzate by trialkylamine extraction to improve the ethanol production
and xylitol from rice straw hydrolysate in a membrane bioreactor. Folia. with Pichia stipitis CBS 5776. Bioresour. Technol. 102, 1663–1668.
Microbiol. 1–11. http://dx.doi.org/10.1007/s12223-015-0420-0.
Zhang, Hui, Sang, Qing, 2015. Production and extraction optimization of xylanase
and mannanase by Penicillium chrysogenum QML-2 and primary application in
saccharification of corn cob. Biochem. Eng. J. 97, 101–110.

Please cite this article in press as: Venkateswar Rao, L., et al. Bioconversion of lignocellulosic biomass to xylitol: An overview. Bioresour. Technol. (2016),
http://dx.doi.org/10.1016/j.biortech.2016.04.092
View publication stats

You might also like