You are on page 1of 18

This article was downloaded by: [University of Newcastle, Australia]

On: 05 January 2015, At: 10:53


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Food Reviews International


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/lfri20

Xylitol Biological Production: A Review


of Recent Studies
a a a
N. L. Mohamad , S. M. Mustapa Kamal & M. N. Mokhtar
a
Department of Process and Food Engineering, Faculty of
Engineering, Universiti Putra Malaysia, Selangor, Malaysia
Accepted author version posted online: 15 Sep 2014.Published
online: 26 Nov 2014.

Click for updates

To cite this article: N. L. Mohamad, S. M. Mustapa Kamal & M. N. Mokhtar (2015) Xylitol
Biological Production: A Review of Recent Studies, Food Reviews International, 31:1, 74-89, DOI:
10.1080/87559129.2014.961077

To link to this article: http://dx.doi.org/10.1080/87559129.2014.961077

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015
Food Reviews International, 31:74–89, 2015
Copyright © Universiti Putra Malaysia
ISSN: 8755-9129 print / 1525-6103 online
DOI: 10.1080/87559129.2014.961077

Xylitol Biological Production: A Review


of Recent Studies

N. L. MOHAMAD, S. M. MUSTAPA KAMAL,


AND M. N. MOKHTAR
Department of Process and Food Engineering, Faculty of Engineering, Universiti
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

Putra Malaysia, Selangor, Malaysia

Xylitol is an alternative sweetener that is recommended for use in food and pharmaceu-
tical products, as it has some health benefits. It is currently produced on a large scale
using a chemical reduction that requires high energy and is costly. Biological conversion
of xylitol using microorganisms is an alternative process that is environmentally friendly
and cost-effective. This process has been studied in an effort to provide one that is high
yielding and competitive with chemical processes. This article reviews recent studies in
the development of biological conversion processes for the production of xylitol, includ-
ing biomass conversion, fermenting microorganisms, and new technology for full-scale
process development.

Keywords Bioengineering, Biomass, Xylitol, Xylose, Yeast

Introduction
Xylitol, pentahydroxypentane (C5 H12 O5 ), is a sugar alcohol or polyol that has many com-
mercial applications. It is widely used in the food and pharmaceutical industries as an
alternative sweetener. This sugar alcohol has been used as a food additive and sweet-
ening agent since the 1960s, but in recent years its use in food formulation has greatly
increased as a result of several favorable properties. Its high chemical and biological sta-
bilities make it useful as a food preservative agent that can extend the shelf life of food
products, and it does not react with amino acids that are responsible for browning effects,
which could reduce the nutritional value of proteins.(1) Xylitol has a sweetening power
similar to sucrose, but a lower caloric value (2.4 cal/g). It is widely used alone or in com-
bination with other sweeteners to enhance flavor.(2) In chewing gums, confectioneries, and
chocolates, xylitol is used to promote a cooling effect because of its high endothermic heat
of solution (34.8 cal/g), and to provide texture and more flexible products with a flavor
similar to sucrose.(3) In addition, xylitol is used as an antioxidant, moisturizer, stabilizer,
cryoprotectant, and freezing point reducer.
The most important benefit of xylitol is its anticariogenic properties, which help
to reduce the formation of caries and plaque on teeth by controlling the growth of
Lactobacillus and Streptococcus mutans in saliva.(3)
Xylitol is recognized as an important chemical derived from carbohydrates, and poten-
tially as a co-product from a plant biomass–based biorefinery.(4) It is present naturally in
Address correspondence to S. M. Mustapa Kamal, Department of Process and Food Engineering,
Faculty of Engineering, Universiti Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia. E-mail:
smazlina@upm.edu.my

74
Xylitol Biological Production 75

Table 1
Factors of biological and chemical processes for xylitol production

Factor Biological Chemical


Carbon source Xylose from lignocellulose Pure xylose
Xylose from lignocellulose
Catalyst Yeast/ bacteria/fungi that Nickel and hydrogenation
required xylose reductase and
xylitol dehydrogenase enzyme.
Process steps 1. Acid or enzymatic hydrolysis 1. Acid hydrolysis of
of lignocellulose lignocellulose
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

2. Detoxification of hydrolysate 2. Purification of hydrolysate to


obtain pure xylose
3. Fermentation of hydrolysate to 3. Hydrogenation of xylose to
xylitol xylitol
4. Xylitol purification 4. Xylitol crystallization
Purification Complex downstream process Ion-exchange resins
because of different microbial
by-products
Cost Lower energy and mild High (need two steps of
temperature purification process, high
energy required, and laborious)

small amounts in fruits and vegetables, algae, and mushrooms, thus hindering its economi-
cal extraction for industrial applications. Therefore, large-scale xylitol production involves
the reduction of D-xylose to xylitol via a catalytic chemical reaction that requires high
pressure and temperature, which increases the cost. Therefore, many researchers have
undertaken extensive studies to develop alternative processes for the production of xylitol.
The biological conversion of xylitol has been targeted for this purpose, as it is less expen-
sive because of the availability of biomass and the use of microorganisms that require less
energy and cost.(5) Table 1 summarizes the difference of xylitol production by chemical
and biological conversion.
The market for xylitol has been increasing since it was first produced in the 1960s
by a Finnish company. In Asia, xylitol is used by gum manufacturers, and it is estimated
that 80–90% of chewing gum sold in the region contains xylitol as a sweetener. One par-
ticular China-based company is estimated to produce approximately 35,000 t of xylitol
per year.(6) The global xylitol market is estimated to be USD $340 million per year(7)
and is expect to increase up to USD $540 million per year within 3 years.(6) However,
the final cost of the product varies and is dependent on raw material costs and transport
costs, which in turn depends on feedstock mass and the location of the manufacturing
plant.(8,9)
Because of the widespread interest in biological xylitol production, research has
increased through the years, with more than 50% of documents being published since
2005 to the present.(10) Therefore, this article aims to provide an overview of the recent
studies available in the literature on the biological production of xylitol.
76 Mohamad et al.

Biomass-to-Xylitol Production
Lignocellulosic biomass is an abundant material that is readily available and cheap.
Generally, lignocellulose is made up of three major components: cellulose, hemicellulose,
and lignin. Hemicellulose is the second most common polysaccharide in nature.(7) It is rich
in xylan, which can be used to prepare a hydrolysate used for the conversion of xylitol by
biological means. A large number of types of lignocellulose biomass has been evaluated as
xylan-rich; therefore, the selection of biomass should meet three criteria: it must be abun-
dant and located within the transportation radius, it must contain high amounts of xylan
and xylose, and it should not contain too many impurities that will increase the risk of con-
tamination during bioconversion and purification.(11) Corn cobs, sugarcane bagasse, and
rice straw are the major materials that have been investigated as biomass for the produc-
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

tion of xylitol.(12−15) Other potential materials include oil palm empty fruit bunch,(16) sago
trunk,(17) Eucalyptus wood,(18) brewery’s spent grain,(19) barley bran, and corn leaves.(20)

Hydrolysis of Biomass
In recent years, extensive research into various fractionation technologies that involve bio-
logical and nonbiological processes used to extract or hydrolyze xylose from hemicellulose
have been investigated. Unfortunately, the complex structural nature of hemicellulose yields
multiple compounds upon sugar degradation, including aliphatic or phenolic acids, furalde-
hydes, and other weak acids. It is known that many of these compounds act as inhibitors
toward the microorganisms during the conversion of xylitol.(19)
Acid hydrolysis is a simple and rapid method that can be performed in highly con-
centrated acid (30–70%, w/w) or dilute acid (<2%, w/w). Generally, highly concentrated
acid hydrolysis is a nonselective process because of the presence of a high amount of acid,
the energy consumption required for acid recovery, the longer reaction time, and the corro-
sive and hazardous nature of the process.(21) Dilute acid hydrolysis is the preferred method
because of its low cost, high rates of reaction, and effectiveness in solubilizing hemicellu-
losic sugars from the lignocellulose structure. Table 2 summarizes the sugar composition
of selected hemicellulose hydrolysates using a dilute acid hydrolysis method.
In recent years, efforts have been undertaken to extract these monomeric sugars
using greener, more environmentally friendly and efficient methods. The acid hydrolysis
method is claimed to cause pollution, and the residue may harm the environment and
equipment. Furthermore, the acid hydrolysis process may produce fermentation inhibitors
(e.g., furfural, phenolic compounds, aliphatic acids) from the sugar degradation. Therefore,
extraction processes that use distilled water at high temperature or hydrothermal processes
have also been investigated. A hydrothermal process is normally used as a pretreatment
process before enzyme hydrolysis.(22) Autohydrolysis and steam explosion are two such
hydrothermal methods. Both processes involve water and steam at high temperature and
pressure. An autohydrolysis process uses compressed hot water with an operation temper-
ature from 150 to 230 ◦ C. The mechanism of degradation by autohydrolysis is similar to
dilute acid hydrolysis where the catalytic hydronium ions are generated in situ by water
autoionization.(19) This process has been successfully applied to the degradation of several
types of lignocellulose materials such as Eucalyptus wood,(23) rice husk,(24) and brewery’s
spent grain.(25) The steam explosion process uses high pressure for wetting the material that
explodes when the pressure in the reactor is rapidly released. Recently, a new process that
uses high temperature and steaming (HTS) has been developed as an alternative hydrolysis
process. The results indicated that the extracted sugar produced by this method was similar
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

Table 2
Composition of hemicellulose hydrolysate from various lignocellulose materials (g L−1 )

Material Hydrolysis Xyl Glc Ara Gal Man Acetic Furfural HMF Phenolics Reference
Eucalyptus Dilute sulfuric 18.00 3.60 0.60 — — 5.2 > 0.50 — — (96)
wood acid
Brewery spent Dilute sulfuric 14.90 5.20 6.20 — — 1.3 0.64 0.05 1.32 (25)
grain acid
Sorghum straw Dilute 4.00 2.90 1.10 — — 1.2 13.70 — — (96)
phosphoric
acid
Oak Dilute sulfuric 106.00 13.20 1.60 8.6 6.8 26.6 4.20 0.50 (97)
acid
Corn stover Dilute sulfuric 67.30 14.30 11.80 7.3 5.6 — — — — (90)
acid

77
Eucalyptus Dilute sulfuric 12.30 0.60 0.84 — — 3.4 0.26 0.07 2.23 (36)
wood acid
Wheat straw Dilute sulfuric 15.40 4.40 2.20 — — 1.2 0.50 0.09 — (87)
acid
Rice straw Dilute sulfuric 23.40 48.30 3.50 — — — — — — (91)
acid
Spruce residue Dilute sulfuric 12.90 9.80 — 4.5 25.3 6.3 0.60 1.20 — (92)
acid
Sorghum straw Dilute 16.20 3.80 — — — 1.9 2.00 — — (93)
hydrochloric
acid
Oil palm Dilute sulfuric 30.80 7.61 — — — 6.5 3.90 — — (16)
empty fruit acid
bunch

(Continued)
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

Table 2
(Continued)

Material Hydrolysis Xyl Glc Ara Gal Man Acetic Furfural HMF Phenolics Reference
Corn cobs Dilute 70.00 7.40 5.30 — — 8.9 — — — (94)
hydrochloric
acid
Corn cobs Dilute sulfuric 70.40 5.60 4.70 — — 8.4 — — — (94)
acid

78
Corn cobs Dilute sulfuric 28.70 5.40 3.70 0.7 0.4 2.0 0.80 0.20 — (12)
acid
Sugarcane Dilute sulfuric 24.30 1.00 — — — 2.1 0.32 0.18 — (83)
bagasse acid
Rice straw Dilute sulfuric 91.15 15.26 18.30 — — 1.5 0.02 0.25 — (35)
acid
Wheat straw Dilute sulfuric 39.60 8.40 6.40 — — 1.6 0.05 0.12 — (87)
acid
Note. Xyl = xylose; Glc = glucose; Ara = arabinose; Gal = galactose; Man = mannose; HMF = hydroxymethylfurfural.
Xylitol Biological Production 79

to that obtained from an acid hydrolysis process.(26) This alternative method, which can be
used for pretreatment or hydrolysis, has been discussed by Franceschin et al.(27) in an effort
to develop an economically sustainable process.
Enzymatic hydrolysis is an alternative hydrolysis process that offers milder operational
conditions (temperature and pH). However, it is often not an efficient process to degrade
untreated biomass. Therefore, necessary pretreatments are needed to reduce the crystallinity
to allow penetration of the enzyme to the hemicellulose structure. Enzymatic hydrolysis is
usually carried out by the xylanase enzyme. Xylanases are enzymes that hydrolyse xylan.
There are two enzymes involved, an endo-1,4-β-xylanase (EC 3.2.1.8) and a β-xylosidase
(EC 3.2.1.37). The endo-xylanase internally cleaves the main chain of xylan, producing a
mixture of xylooligosaccharides, whereas the β-xylosidase liberates xylose by removing
the terminal monosaccharide from the nonreducing end of the short oligosaccharides.(28)
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

Yoon et al.(29) has successfully demonstrated the use of the only commercial xylanase
enzyme for the extraction of pentoses from corn residue. Moreover, a mixture of cellulase
and xylanase has been reported to be more efficient at breaking the structure of biomass,
but the yield depends on the solid-to-biomass ratio, reaction time, and enzyme loading.(30)
Enzymatic hydrolysis of hemicellulose has been studied using maize straw,(31) agricultural
waste,(32) and wheat straw.(33)

Detoxification
Generally, toxic compounds or inhibitors are generated during an acid hydrolysis process.
Furan derivatives, aliphatic acids, and phenolic compounds are inhibitors that are produced
from the degradation of hexoses and pentoses, acetyl groups, and lignin. It is necessary
to remove these metabolic inhibitors to improve the fermentabilities of hemicellulose
hydrolysates by microorganisms. Detoxification treatments have been developed to
minimize the inhibitory effect of hydrolysates. However, the detoxification strategy does
not increase the total cost of purifying the hydrolysate liquor because, in fact, it is a process
that transforms the purification of the fermentation products to the purification of the
substrate.(11)
The efficiency of detoxification methods directly depends on the type of biomass,
method of hydrolysis, and microorganisms used.(34) Vacuum evaporation is a physical
detoxification method that reduces volatile compounds such as acetic acid, furfural,
hydroxymetylfurfural (HMF), and vanillin that are normally present in hydrolysates.
Mussatto and Roberto(35) reported that about 90% of furfural is removed when vacuum
evaporation is applied to wood, rice straw, and sugarcane bagasse hydrolysates. However,
this method increased the concentration of volatile compounds and reduces the overall
hydrolysate volume.
Overliming, organic solvent extraction, ion-exchange resin treatment, and activated
carbon adsorption are the chemical methods that are generally used to remove toxic com-
pounds after hydrolysate generation. Overliming or pH adjustment is effective in terms of
cost and chemical detoxification compared with the other methods. Calcium hydroxide and
sulfuric acid are commonly used for the efficient removal of phenolic compounds, ketones,
furfurals, and hydroxymethylfurfurals from hydrolysates.(36) However, if the hydrolysates
are subjected to vacuum concentration immediately after precipitation, they will form a
heavy calcium sulfate scale on the inner wall of the heater that will inhibit heat trans-
fer. Therefore, for large-scale production, the neutralized hydrolysates can be subjected
to evaporation concentration only, either through decalcification by ion exchange, or by
the addition of antiscaling agents. This is crucial to the stable running of the evaporation
80 Mohamad et al.

equipment. Activated charcoal is a low-cost process with the ability to absorb pigments,
free fatty acids, n-hexane, and other oxidation products. However, the effectiveness of this
treatment generally depends on variables such as pH, temperature, contact time, and the
solid-to-liquid ratio.(7) In hydrolysates treated by ion exchange, a certain amount of inor-
ganic salts, such as phosphate and potassium, must be retained so that they can be used
for yeast cell growth. After the end of the fermentation, they can be removed from the
fermentation broth.
If the employed yeast strain exhibits rather good tolerance to toxic substances, the
complete removal of toxic compounds from the fermentation medium is not necessary(37) ;
in fact, some toxic substances at low concentrations may even be beneficial to xylitol pro-
duction. These yeasts exhibit good activity in degrading toxic substances and many small
molecule impurities (acetic acid, furfural, phenol), which significantly improves xylitol
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

fermentative production performance and effectively reduces the burden of physical and
chemical purification or detoxification processes.(38) Determining the appropriate degree of
desalination in the ion exchange of hydrolysates can promote reasonable use of inorganic
salts, improve detoxification, and reduce costs.

Bioconversion of Xylitol Production


Two main areas are normally considered when investigating a fermentation process: strain
development and process development. Xylitol can be biologically produced by yeasts,
bacteria, and fungi. Among them, yeast has been reported to have the capability to produce
the highest yield of xylitol. Many potential xylitol-producing strains have been isolated,
and attention has been paid to the improvement of the strains, either through mutation or
genetic engineering techniques.
Winkelhausen and Kusmanova(39) has briefly discussed the metabolism of D-xylose
in yeast where the conversion occurs via two steps: reduction and oxidation. In the first
step, D-xylose is reduced to xylitol by a xylose reductase (XR) and then oxidized to xylu-
lose by xylitol dehydrogenase (XDH) before phosphorylation into xylulose-5-phosphate,
catalyzed by xylulokinase (XK). These two steps are considered to be rate limiting, and
XR and XDH are key enzymes in D-xylose fermentation and xylitol production, respec-
tively (Figure 1). Both enzymes require the pyridine nucleotide cofactors NAD+ /NADH
or NADP+ /NADPH depending on the type of yeast.(40) Moreover, the role and regenera-
tion of cofactors are highly dependent on oxygen level and oxygen transfer rates.(41) Under
anaerobic conditions, yeasts are unable to metabolize D-xylose because of redox imbalance
between NAD+ and NADH. At a low oxygen level, the electron transport system is unable
to oxidize intracellular NADH completely, increasing the NADH concentrations and per-
mitting xylitol excretion. At a high oxygen level, the oxidation of NADH to NAD+ occurs,
and the high NAD+ /NADH ratio favors xylitol oxidation to xylulose(13) where NAD+ is
regenerated by the respiratory chain. Excessive NADH cannot be oxidized by yeast, as it
does not possess an enzyme with transhydrogenase activity.(41)
Among yeasts, Candida sp. are known to be a good potential source of xylitol from
xylose as a carbon source.(42) Candida tropicalis produced xylitol in 80–90% yield from
xylose as the sole carbon source, with a productivity of more than 3 g L−1 h−1 .(43) Tada
et al.(44) reported that Candida magnoliae produced 18 g L−1 xylitol from 25 g L−1 xylose
extracted from corn cob hydrolysate, to give a yield of 0.6 g xylitol per g xylose under
oxygen limited conditions. A maximum xylitol yield of 0.56 g g−1 was obtained from the
optimum fermentation conditions for Candida peltata, whereas Candida boidinii produced
52% xylitol compared with theoretical yields from 150 g L−1 xylose when cultivated for
Xylitol Biological Production 81

Xylose Glucose

Cell membrane

Xylose Glucose
xylose NADPH
reductase HMP
NADP+ pathway
xylitol
dehydrogenase
XYLITOL Xylitol Xylulose Xlu-5P Glu-6P Biomass

Transport NAD+ NADH

Respiratory
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

chain

H2O O2

Figure 1. Metabolic pathways for xylitol production by Candida mogii (adapted from
Tochampa(86) ).
© Springer Science + Business Media. Reproduced with kind permission from Springer Science and
Business Media. Permission to reuse must be obtained from the rightsholder.

14 days.(45) Previous studies conducted with Candida guilliermondii found that by control-
ling the oxygen supply, 0.7–0.8 g g−1 of xylitol with a productivity above 0.8 g L−1 h−1 was
obtained from a medium containing 20 g L−1 xylose.(13,46,47) Sampaio et al.(48) screened
270 yeasts and found that Debaryomyces hansenii UFV-170 exhibited the highest xylitol
production, with a 0.54 g g−1 yield and a productivity of 0.24 g L−1 h−1 . The effects of
fermentation conditions(48) and model identification(49) have been further investigated for
this species.
Many studies have been done to evaluate the effect of mixed substrates on the
growth and productivity of the yeast for xylitol production. Because the fermentation of
lignocellulose hydrolysates often produces various sugars (e.g., xylose, glucose, mannose,
galactose, and arabinose), it is important to investigate their effects on the induction of XR
and XDH.(39) The presence of mixed substrates may interfere with the XR/XDH pathway,
either by directly influencing both substrate uptake and XR/XDH gene expression or by
varying the NADP/NADH ratio. Tamburini et al.(50) evaluated the activity of XR and XDH
on different substrates for Candida tropicalis and found that glucose significantly inhibited
xylose reduction, galactose stimulated xylitol use, and maltose improved biomass growth
and xylitol accumulation. Earlier studies have investigated the effect of glucose and xylose
feeding at different ratios to increase the xylitol yield. Oh and Kim(51) found a 15% feeding
ratio corresponds to a yield of 93% from 270 g L−1 xylose. Another study reported the
presence of glucose inhibited xylose use, and glucose was initially consumed. The maxi-
mum xylitol yield (0.84 g g−1 ) and volumetric productivity (0.49 g L−1 h−1 ) were obtained
for substrates containing high arabinose and low glucose and mannose.(52)
Besides yeasts, bacteria and fungi are also capable of producing xylitol, albeit with
a lower yield of xylitol compared with yeasts. A few bacteria have been reported to pro-
duce xylitol in small amounts, including Mycobacterium smegmatis(53) and Glucunobacter
oxydans.(54) Rangaswamy and Agblevor(55) Rangaswamy and Agblevor (55) screened
82 Mohamad et al.

17 cultures of facultative bacteria and found that Corynebacterium sp. produced the highest
amount of xylitol, with a maximum yield of 0.57 g g−1 from a 75 g L−1 initial xylose con-
centration within 24 hours of cultivation. The research on xylitol bioproduction using fungi
has been conducted by Sampaio et al.(56) who screened 11 filamentous fungi belonging
to the genera of Aspergillus and Penicillium. Penicillium crustosum presented the highest
xylitol production, with 0.52 g L−1 xylitol from 11.5 g L−1 xylose, representing a xylose
reduction of 76%. In conclusion, bioconversion of xylitol by bacteria and fungi does not
favorably compare with that performed by yeasts and generally has not found widespread
interest because of the low amount of xylitol produced.

Xylitol Fermentation Process Factors


Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

Xylitol is an extracellular metabolite and many factors influence the xylitol production
rate, including the components of the nutrient medium, culture conditions, cell density, and
growth rate.
High initial xylose concentrations are needed to produce a high amount of xylitol.
The effects of a high xylose concentration on the ability of microorganisms to produce
xylitol has been studied by Ikeuchi et al.(42) A high xylose concentration (300 g L−1 ) in
the medium slowed cell growth and resulted in the need to increase the fermentation time
to obtain the maximum xylitol concentration. The optimum concentration was obtained at
200–300 g L−1 xylose for C. guilliermondii, whereas growth and xylitol production were
limited for xylose concentrations of 400 g L−1 .(57)
The optimum production of xylitol using yeast was obtained at 30 ◦ C. However, a
temperature range of between 30 and 37 ◦ C is acceptable, as generally yeasts are not sen-
sitive to temperature changes within this range.(58) The conversion of D-xylose to xylitol
by Candida sp. B-22 was relatively constant from 35 to 40 ◦ C, whereas the conversion was
significantly reduced when the cultivated temperature was higher than 45 ◦ C.(59)
Normally, the pH range for yeast cultivation is between 4 and 6. The maximum growth
for C. parapsilosis and C. guilliermondii are at pH 6.0,(57,60) whereas C. mogii and P. stipitis
grow the best at pH 5 and 5.5, respectively.(61) Candida tropicalis can be cultivated at any
pH between 4.5 and 5.5.(34) In contrast, C. tropicalis is highly resistant to acidic conditions
and provided a maximum xylitol yield at pH 2.5.(58)
Oxygen is a key element for the D-xylose metabolism in yeast. Accumulation of xylitol
always occurs under oxygen-limited conditions.(41,47,57,62) In aerobic conditions, xylose
fermentation is favored for cell growth, whereas yeast cannot use xylose under strict anaer-
obic conditions. Therefore, appropriate oxygen control is necessary for effective xylitol
production.(63) Walther et al.(52) studied the influence of oxygen limitation and the initial
xylose concentration using C. tropicalis as the catalyst and found the maximum xylitol yield
of 0.7 g g−1 was achieved under semiaerobic conditions with an initial substrate concentra-
tion of 150 g L−1 . Santos et al.(64) reported that when C. guilliermondii was immobilized
on porous glass in a fluidized bed reactor and the aeration rate was increased from 0.031 to
0.093 min−1 , the cell concentration increased by more than 8 g L−1 , but the xylitol concen-
tration decreased by about 2 g L−1 . According to these authors, the cell growth and xylitol
yield were influenced by the availability of oxygen, which was generated when bubbles
were contained in the medium burst.

Developing New Techniques for Xylitol-Producing Microbes


Recent studies on xylitol production have focused on developing techniques for safer,
greener, and environmentally friendly routes using microbes. Commonly investigated
Xylitol Biological Production 83

strains include the natural xylose-fermenting yeasts C. tropicalis and D. hansenii, a genet-
ically engineered strain or a recombinant strain. The recombinant strains are developed
to obtain a higher xylitol yield that is more tolerant to the hydrolysate in the presence
of inhibitors for xylose fermentation.(7) In addition, the use of Candida sp. in the food
industry is undesirable because of the pathogenic nature of many Candida sp. The con-
struction of recombinant yeasts to produce xylitol has been recognized by introduction
of xylose reduction capabilities into Saccharomyces cerevisiae,(65) Lactococcus lactis,(66)
Corynebacterium glutamicum,(67) and microalgae.(68) Most of these microorganisms are
generally recognized as safe (GRAS) and structurally stable owing to their lipid-rich
outer layer that is tolerant to lysis by lytic enzymes and to osmotic stress.(67) Cheng
et al.(69) studied xylitol production by combining recombinant Bacillus subtilis and C.
maltosa in the xylose mother liquor. C. maltosa was used to detoxify the hydrolysate,
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

whereas recombinant B. subtilis was used for purification of xylose from the mother
liquor.
Saccharomyces cerevisiae is a preferred species for biofuel production because it is
an efficient glucose fermentor, but it cannot use xylose. Therefore, extensive genetic engi-
neering has been performed to overcome this drawback by developing xylose-fermenting S.
cerevisiae.(70−72) Krahulec et al.(73) reported the analysis and predicted the performance of
S. cerevisiae from kinetic comparisons of wild-type and engineered forms of XR and XDH.
The results showed that the lowest xylitol yield was obtained for strains that harbored XR
engineered for use of NADH compared with strains harboring XDH engineered for use of
NAPD+ . Kim et al.(74) found that for S. cerevisiae, the overexpression of XYL2 coding for
XDH improved the ethanol yields but decreased xylitol yields from xylose. The expres-
sion of XYL1 coding for XR from Pichia stipitis in S. cerevisiae in the presence of the
cellodextrin transporter (cdt-1) and intracellular β-glucosidase (g h−1 ) produced xylitol via
simultaneous use of cellobiose and xylose, resulting in a 37–63% improvement in xylitol
productivity.(65)
Generally, the formation of XR by yeasts is significantly repressed when grown on a
medium that contains glucose as a carbon source. Jeon et al.(75) developed XR from the
ascomycetes Neurospora crassa (NcXR), which has high catalytic efficiency, but it was not
expressed in C. tropicalis because of differences in codon use between the two species.
High expression of the NcXR gene was confirmed by determining XR activity in cells
grown on glucose and resulted in a xylitol production rate of 1.44 g L−1 h−1 and a xylitol
yield of 96% at 44 hours of cultivation. A recombinant strain of Kluyveromyces marxianus
constructed by expressing the N. crassa xylose reductase gene YZJ015 and maintaining
the xylitol dehydrogenase gene produced xylitol at 1.01 g g−1 xylose with glycerol as a
co-substrate at temperatures as high as 45 ◦ C.(76)
Candida tropicalis has been shown to ferment a hemicellulose hydrolysate prepared
from dilute acid hydrolysis without a detoxification step. A xylitol yield of 0.71 g g−1
xylose was obtained when using with rice-straw hydrolysate.(77) Misra et al.(78) used an
adapted version of C. tropicalis in the presence of corncob hemicellulose hydrolysate to
obtain an increase of 1.22-fold of xylitol yield when compared with the parent strain.
Similar work was carried out on sugarcane bagasse,(63) corn cobs,(44,79) and rice hull.(80) An
engineered strain of S. cerevisiae, which carried xylanase, β-xylosidase, and xylose reduc-
tase genes controlled by different transcriptional regulations, was constructed to directly
convert xylan to xylitol. A xylitol yield of 0.71 g g−1 xylan was obtained by optimizing the
transcriptional regulation and fermentation processes.(81) In other research, the reduction
of xylose to xylitol from hemicellulose hydrolysate was successfully carried out with engi-
neered Escherichia coli by eliminating L-arabinitol formation to produce xylitol at almost
100% purity.(82)
84 Mohamad et al.

Downstream Processes for Xylitol Production


Fermentation broth that contains a certain concentration of xylitol must be separated from
the yeast cells before the purification and crystallization can be performed. Generally,
centrifugation and membrane filtration are used to separate cells from the fermented broth.
Martínez et al.(83) developed a downstream processes for xylitol produced from sugarcane
hemicellulose hydrolysate using purification by ion exchange and crystallization. This
research resulted in xylitol crystals with 92–94% purity. An earlier study used the same
hydrolysate by adding activated carbon to the fermented broth at 80 ◦ C for 60 minutes at
pH 6.0.(84) Faveri et al.,(85) who developed a method to recover xylitol from fermented
and purified broth consisting of evaporation up to supersaturation, supersaturation cool-
ing, separation of crystals by centrifugation, and final filtration, found that the best results
were obtained from the concentrated solutions (730 g L−1 ) at −5 ◦ C. These parameters
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

were further optimized by using response surface methodology, and the suggested operating
conditions were xylitol concentration 728 g L−1 at −6 ◦ C under purity degree of 0.97 and
xylitol crystallization yield of 0.54.(86) According to Hou-Rui,(11) different purification pro-
cesses, including the use of activated carbon, ion-exchange, or ultrafiltration processes,
have both advantages and disadvantages. Activated carbon is useful for clarification and
deodorization because of its ability to adsorb large-molecule pigments. Ion exchange effec-
tively removes inorganic salts and adsorbs small molecule pigments, whereas ultrafiltration
removes proteins, and improves the working performance of ion-exchange resins, thus
extending their service life.

Conclusions
Xylitol is an important alternative sweetener that has attracted commercial demand because
of its potential applications in the food and pharmaceutical industries. Current studies in
xylitol research have focused on the development of engineered strains of microorgan-
isms in an effort to improve stability and tolerance to toxicity compared with the wild-type
strains. Biological production of xylitol has greatly improved; however, there are still many
opportunities to improve the performance of metabolically engineered microorganisms,
develop efficient xylitol recovery processes from the fermentation broth, and enhance the
yield and productivity in a cost-effective manner. No reports are available in the litera-
ture concerning the direct production of xylitol through saccharification and simultaneous
fermentation or enzymatic synthesis from lignocellulosic biomass. Bioconversion of xylitol
from lignocellulosic biomass using enzyme technology offers significant advantages for the
biorefinery industry. The components of hydrolysis residue, which are complexes of cellu-
lose and lignin, can be dried using tail gas from boilers and used as fuels. In addition, the
components of salinated wastewater produced from the purification of fermentation liquor
that have economic value can be recycled, whereas the yeast recovered from the xylitol
fermentation process can be served as single-cell protein in feed industry. Mathematical
modeling and simulations should prove useful for predicting the performance of processes
and production on a large scale in the future.

Funding

The authors gratefully acknowledge the financial support from the Ministry of Education of Malaysia
under Exploratory Research Grant Scheme (ERGS).
Xylitol Biological Production 85

References
1. Parajo, J.C.; Dominguez, H.; Dominguez, J.M. Biotechnological production of xylitol. Part 1:
Interest of xylitol and fundamentals of its biosynthesis. Bioresour. Technol. 1998, 66, 25–40.
2. Zacharis, C. Xylitol. In Sweeteners and sugar alternatives in food technology; Donell, K.,
Kearsley, M., Eds.; Wiley-Blackwell Publishing: Oxford, UK, 2012; pp 347–382
3. Mussatto, S.I. Application of xylitol in food formulations and benefits for health. In d-Xylitol;
Silva, S.S., Chandel, A.K., Eds.; Springer: Berlin and Heidelberg, 2012; pp 309–323.
4. Bozell, J. Feedstocks for the future-biorefinery production of chemicals from renewable carbon.
Clean 2008, 36, 641–647.
5. Nigam, P.; Singh, D. Processes for fermentative production of xylitol—A sugar substitute.
Process Biochem. 1995, 30, 117–124.
6. Ravella, S.R.; Gallagher, J.; Fish, S.;.Prakasham, R.S. Overview on commercial production
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

of xylitol, economic analysis and market trends. In d-Xylitol; Silva S.S., Chandel, A.K., Eds.;
Springer, Berlin and Heidelberg, 2012; pp 291–306.
7. Prakasham, R.S.; Rao, R.S.; Hobbs, P.J. Current trends in biotechnological production of xylitol
and future prospects. Curr. Trends Biotechnol. Pharm. 2009, 3, 8–36.
8. Kazi, F.; Fortman, J.; Anex, R.; Hsu, D.; Aden, A.; Dutta, A.; Kothandaraman, G. Techno-
economic comparison of process technologies for biochemical ethanol production from corn
stover. Fuel 2010, 89, S20–S28.
9. Lange, J. Lignocellulose conversion: An introduction to chemistry, process and economics.
Biofuels Bioprod. Biorefin. 2007, 1, 39–48.
10. Scopus. http://www.scopus.com/term/analyzer.url (accessed May 27, 2014).
11. Zhang, H.R. Key drivers influencing the large scale production of xylitol. In d-Xylitol; Silva,
S.S., Chandel, A.K., Eds.; Springer, Berlin and Heidelberg, 2012; pp 267–289.
12. Cheng, K.K.; Zhang, J.A.; Ling, H.Z.; Ping, W.X.; Huang, W.; Ge, J.P.; Xu, J.M. Optimization
of pH and acetic acid concentration for bioconversion of hemicellulose from corncobs to xylitol
by Candida tropicalis. Biochem. Eng. J. 2009, 43, 203–207.
13. Martínez, E.A.; da Silva, S.S.; Silva, J.B.A.E.; Solenzal, A.I.N.; Felipe, M.G.A. The influence of
pH and dilution rate on continuous production of xylitol from sugarcane bagasse hemicellulosic
hydrolysate by C. guilliermondii. Process Biochem. 2003, 38, 1677–1683.
14. Mussatto, S.I.; Roberto, I.C. Xylitol production from high xylose concentration: Evaluation
of the fermentation in bioreactor under different stirring rates. J. Appl. Microbiol. 2003, 95,
331–337.
15. Rao, R.S.; Jyothi, C.P.; Prakasham, R.S.; Sarma, P.N.; Rao, L.V. Xylitol production from corn
fiber and sugarcane bagasse hydrolysates by Candida tropicalis. Bioresour. Technol. 2006, 97,
1974–1978.
16. Rahman, S.H.A.; Choudhury, J.P.; Ahmad, A.L. Production of xylose from oil palm empty fruit
bunch fiber using sulfuric acid. Biochem. Eng J. 2006, 30, 97–103.
17. Mohamad, N.L.; Mustapa Kamal, S.M.; Abdullah, A.G.L. Optimization of xylose production
from sago trunk cortex by acid hydrolysis. Afr. J. Food Sci. Technol. 2011, 2, 102–108.
18. Diz, J.; Cruz, J.M.; Dominguez, H.; Parajo, J.C. Xylitol production from Eucalyptus wood
hydrolysates in low-cost fermentation media. Food Technol. Biotechnol. 2002, 40, 191–197.
19. Gírio, F.M.; Carvalheiro, F.; Duarte, L.C. Deconstruction of the hemicellulose fraction from
lignocellulosic materials into smple sugars. In d-Xylitol; Silva S.S.; Chandel, A.K. Eds.; Springer,
Berlin and Heidelberg, 2012; pp 3–37.
20. Cruz, J.M.; Dominguez, J.M.; Dominguez, H.; Parajo, J.C. Preparation of fermentation media
from agricultural wastes and their bioconversion into xylitol. Food Biotechnol. 2000, 14, 79–97.
21. Taherzadeh, M. J.; Karimi, K. Acid-based hydrolysis processes for ethanol from lignocellulosic
materials: A review. Bioresources 2007, 2, 472–499.
22. Walch, E.; Zemann, A.; Schinner, F.; Bonn, G.; Bobleter, O. Enzymatic Saccharification of
hemicellulose obtained from hydrothermally pretreated sugar cane bagasse and beech bark.
Bioresour. Technol. 1992, 39, 173–177.
86 Mohamad et al.

23. Garrote, G.; Dominguez, H.; Parajo, J.C. Generation of xylose solutions from Eucalyptus
globulus wood by autohydrolysis-posthydrolysis processes: Posthydrolysis kinetics. Bioresour.
Technol. 2001, 79, 155–164.
24. Vegas, R.; Kabel, M.; Schols, H.; Alonso, J.; Parajo, J.C. Hydrothermal processing of rice husks:
Effects of severity on product distribution. J. Chem. Technol. Biotechnol. 2008, 83, 965–972.
25. Carvalheiro, F.; Esteves, M.; Parajo, J.C.; Pereira, H.; Girio, F.M. Production of oligosaccharides
by autohydrolysis of brewery’s spent grain. Bioresour. Technol. 2004, 91, 93–100.
26. Wang, L.; Yang, M.; Fan, X.; Zhu, X.; Xu, T.; Yuan, Q. An environmentally friendly and effi-
cient method for xylitol bioconversion with high-temperature-steaming corncob hydrolysate by
adapted Candida tropicalis. Process Biochem. 2011, 46, 1619–1626.
27. Franceschin, G.; Sudiro, M.; Ingram, T.; Smirnova, I.; Brunner, G.; Bertucco, A. Conversion of
rye straw into fuel and xylitol: A technical and economical assessment based on experimental
data. Chem. Eng. Res. Des. 2011, 89, 631–640.
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

28. Peng, F.; Peng, P.; Xu, F.; Sun, R.C. Fractional purification and bioconversion of hemicelluloses.
Biotechnol. Adv. 2012, 25, 879–903.
29. Yoon, K.Y.; Woodams, E.E.; Hang, Y.D. Enzymatic production of pentoses from the
hemicellulose fraction of corn residues. Food Sci. Technol. 2006, 39, 387–391.
30. Lin, L.; Yan, R.; Liu, Y.; Jiang, W. In-depth investigation of enzymatic hydrolysis of biomass
wastes based on three major components: Cellulose, hemicellulose and lignin. Bioresour.
Technol. 2010, 101, 8217–8223.
31. Chen, M.; Zhao, J.; Xia, L. Enzymatic hydrolysis of maize straw polysaccharides for the
production of reducing sugars. Carbohydr. Polym. 2008, 71, 411–415.
32. Tran, L.; Yogo, M.; Ojima, H.; Idota, O.; Kawai, K.; Suzuki, T.; Takamizawa, K. The production
of xylitol by enzymatic hydrolysis of agricultural wastes. Biotechnol. Bioprocess Eng. 2004, 9,
223–228.
33. Liaviga, A.; Bian, Y.; Seib, P. Release of D-xylose from wheat straw by acid and xylanase
hydrolysis and purification of xylitol. J. Agric. Food Chem. 2007, 55, 7758–7766.
34. Rao, R.S.; Jyothi, C.P.; Prakasham, R.S.; Rao, C.S.; Sarma, P.N.; Rao, L.V. Strain improve-
ment of Candida tropicalis for the production of xylitol: Biochemical and physiological
characterization of wild type and mutant strain CT-OMV5. J. Microbiol. 2006, 44, 113–120.
35. Mussatto, S.I.; Roberto, I.C. Alternatives for detoxification of diluted acid lignocellulosic
hydrolysates for use in fermentative processes: A review. Bioresour. Technol. 2004, 93, 1–10.
36. Villarreal, M.L.M.; Prata, A.M.R.; Felipe, M.G.A.; Silva, J.B.A.E. Detoxification procedures of
eucalyptus hemicellulose hydrolysate for xylitol production by Candida guilliermondii. Enzyme
Microb. Technol. 2006, 40,17–24.
37. Pereira, R.S.; Mussatto, S.I.; Roberto, I.C. Inhibitory action of toxic compounds present in
lignocellulosic hydrolysates on xylose to xylitol bioconversion by Candida guilliermondii. J.
Ind. Microb. Biotechnol. 2011, 38, 71–78.
38. Zhang, H.R.; Qin, X.X.; Silva, S.S. Novel isolates for biological detoxification of lignocellulose
hydrolysates. Appl. Biochem. Biotechnol. 2009, 152, 199–212.
39. Winkelhausen, E.; Kusmanova, S. Microbial conversion of D-xylose to xylitol. J. Ferment.
Bioeng. 1998, 86, 1–14.
40. Goli, J.K.; Panda, S.H. Molecular mechanism of D-xylitol production in yeasts: Focus on molec-
ular transportation, catbolic sensing and stress response. In D-Xylitol; Silva S.S.; Chandel, A.K.
Eds.; Springer, Berlin and Heidelberg, 2012; pp 85–107.
41. Furlan, S.; Bouilloud, P.; Castro, H.F. Influence of oxygen on ethanol and xylitol production by
xylose fermenting yeasts. Process Biochem. 1994, 29, 657–662.
42. Ikeuchi, T.; Azuma, M.; Kato, J.; Ooshima, H. Screening of microorganisms for xylitol pro-
duction and fermentation behavior in high concentrations of xylose. Biomass Bioeng. 1999, 16,
333–339.
43. Jeon, Y.J.; Shin, H.S.; Rogers, P.L. Xylitol production from amutant strain of Candida tropicalis.
Lett. Appl. Microbiol. 2011, 53, 106–113.
Xylitol Biological Production 87

44. Tada, K.; Kanno, T.; Horiuchi, J. Enhanced production of bioxylitol from corn cobs by Candida
magnoliae. Ind. Eng. Chem. Res. 2012, 51, 10008–10014.
45. Saha, B.C.; Bothast, R.J. Production of xylitol by Candida peltata. J. Ind. Microbiol. Biotechnol.
1999, 22, 633–636.
46. Faria, L.F.F.; Gimenes, M.A.P.; Nobrega, R.; Pereira, N. Influence of oxygen availability on cell
growth and xylitol production by Candida guilliermondii. Appl. Biochem. Biotechnol. 2002,
98-100, 449–458.
47. Gimenes, M.A.P.; Carlos, L.C.S.; Faria, L.F.F.; Pereira, J.N. Oxygen uptake rate in production of
xylitol by Candida guilliermondii with different aeration rates and initial xylose concentrations.
Appl. Biochem. Biotechnol. A Enzyme Eng. Biotechnol. 2002, 98–100, 1049–1059.
48. Sampaio, F.C.; Chaves-Alves, V.; Converti, A.; Passos, F.M.L.; Coelho, J.L. Influence of cultiva-
tion conditions on xylose-to-xylitol bioconversion by a new isolate of Debaryomyces hansenii.
Bioresour. Technol. 2008, 99, 502–508
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

49. Na, J.; Gi, F.; Roseiro, C.; Nahlik, J.; Palatova, M.; Girio, F. Model identification and physio-
logical control of xylitol production using Debaryomyces hansenii. Process Biochem. 2003, 38,
1695–1705.
50. Tamburini, E.; Bianchini, E.; Bruni, A.; Forlani, G. Cosubstrate effect on xylose reductase and
xylitol dehydrogenase activity levels, and its consequence on xylitol production by Candida
tropicalis. Enzyme Microb. Technol. 2010, 46, 352–359.
51. Oh, D.K.; Kim, S. Increase xylitol yield by feeding xylose and glucose in Candida tropicalis.
Appl. Microbiol. Biotechnol. 1998, 50, 419–425.
52. Walther, T.; Hensirisak, P.; Agblevor, F.A. The influence of aeration and hemicellulosic sugars
on xylitol production by Candida tropicalis. Bioresour. Technol. 2001, 76, 213–220.
53. Izumori, K.; Tuzaki, K. Production of xylitol from D-xylulose by Mycobacterium smegmatis. J.
Ferment. Technol. 1988, 66, 33–36.
54. Suzuki, S.; Sugiyama, M.; Mihara, Y.; Hashiguchi, Y.; Yokozeki, K. Novel enzymatic method
for the production of xylitol from D-arabitol by Gluconobacter oxydans. Biosci. Biotechnol.
Biochem. 2002, 66, 2614–2620.
55. Rangaswamy, S.; Agblevor, F.A. Screening of facultative anaerobic bacteria utilizing D-xylose
for xylitol production. Appl. Microbiol. Biotechnol. 2002, 60, 88–93.
56. Sampaio, F.C.; Da Silveira, W.B.; Chaves-Alves, V.M.; Lopes Passos, F.M.; Cavalcante Coelho,
T.L. Screening of filamentaous fungi for production of xylitol from D-xylose. Braz. J. Microbiol.
2003, 34, 325–328.
57. Nolleau, V.; Preziosi-Belloy, L.; Delgenes, J.P.; Navarro, J.M. Xylitol production from xylose by
two yeast strains: Sugar tolerence. Curr. Microbiol. 1993, 27, 191–197.
58. Silva, S.S.; Afschar, A.S. Microbial production of xylitol from D-xylose using Candida
tropicalis. Bioprocess Eng. 1994, 11, 129–134.
59. Slininger, P.J.; Bolen, P.L.; Kurtzman, C.P. Pachysolen tannophilus: Properties and process
consideration for ethanol production from D-xylose. Enzyme Microb. Technol. 1987, 9, 5–15.
60. Sampaio, F.C, de Moraes, C.A.; De Faveri, D.; Perego, P.; Converti, A.; Passos, F.M.L.; Faveri,
D.De. Influence of temperature and pH on xylitol production from xylose by Debaryomyces
hansenii UFV-170. Process Biochem. 2006, 41, 675–681.
61. Ghindea, R.; Csutak, O.; Stoica, I.; Tanase, A.M.; Vassu, T. Production of xylitol by yeasts.
Romanian Biotechnol. Lett. 2010, 15, 5217–5222.
62. Zou, Y.; Qi, K.; Chen, X.; Miao, X.; Zhong, J.J. Favorable effect of very low initial KLa value
on xylitol production from xylose by a self-isolated strain of Pichia guilliermondii. J. Biosci.
Bioeng. 2010, 109, 149–152.
63. Branco, R.F.; Santos, J.C.; Murakami, L.Y.; Mussatto, S.I.; Dragone, G.; Silva, S.S. Xylitol pro-
duction in a bubble column bioreactor: Influence of the aeration rate and immobilized system
concentration. Process Biochem. 2007, 42, 258–262.
64. Santos, J.C.; Converti, A.; De Carvalho, W.; Mussatto, S.I.; Da Silva, S.S. Influence of aera-
tion rate and carrier concentration on xylitol production from sugarcane bagasse hydrolyzate in
immobilized-cell fluidized bed reactor. Process Biochem. 2005, 40, 113–118.
88 Mohamad et al.

65. Oh, E.J.; Ha, S.J.; Rin Kim, S.; Lee, W.H.; Galazka, J.M.; Cate, J.H.D.; Jin, Y.S. Enhanced
xylitol production through simultaneous co-utilization of cellobiose and xylose by engineered
Saccharomyces cerevisiae. Metab. Eng. 2013, 15, 226–234.
66. Nyysola, A.; Pihlajaniemi, A.; Palva, A.; von Weyman, N.; Leisola, M. Production of xylitol
from D-xylose by recombinant Lactococcus lactis. J. Biotechnol. 2005, 118, 55–66.
67. Kim, S.S.G.; Yun, J.Y.; Seo, J.; Park, J.B. Production of xylitol from D-xylose and glucose with
recombinant Corynebacterium glutamicum. Enzyme Microb. Technol. 2010, 46, 366–371.
68. Pourmir, A.; Noor-Mohammadi, S.; Johannes, T.W. Production of xylitol by recombinant
microalgae. J. Biotechnol. 2013, 165, 178–183.
69. Cheng, H.; Wang, B.; Lv, J.; Jiang, M.; Lin, S.; Deng, Z. Xylitol production from xylose mother
liquor: A novel strategy that combines the use of recombinant Bacillus subtilis and Candida
maltosa. Microb. Cell Factories 2011, 10, 1–12.
70. Hahn-Hagerdal, B.; Karnumoa, K.; Jeppsson, H.; Gorwa-Grauslund, M.F. Metabolic engineering
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

for pentose utilization of Saccharomyces cerevisiae. Adv. Biochem. Eng. Biotechnol. 2007, 108,
147–177.
71. Krahulec, S.; Klimacek, M.; Nidetzky, B. Engineering of a matched pair of xylose reductase
and xylitol dehydrogenase for xylose fermentation by Saccharomyces cerevisiae. Biotechnol. J.
2009, 4, 684–694.
72. Watanabe,.S.; Saleh, A.; Pack, S.; Annaluru, N.; Kodaki, T.; Makino, K. Ethanol production
from xylose by recombinant Saccharomyces cerevisiae expressing protein engineered NADP+
dependent xylitol dehydrogenase. J. Biotechnol. 2007, 130, 316–319.
73. Krahulec, S.; Klimacek, M.; Nidetzky, B. Analysis and prediction of the physiological effects
of altered coenzyme specificity in xylose reductase and xylitol dehydrogenase during xylose
fermentation by Saccharomyces cerevisiae. J. Biotechnol. 2012, 158, 192–202.
74. Kim, S.R.; Ha, S.J.; Kong, I.I.; Jin, Y.S. High expression of XYL2 coding for xylitol dehydro-
genase is necessary for efficient xylose fermentation by engineered Saccharomyces cerevisiae.
Metab. Eng. 2012, 14, 336–343.
75. Jeon, W.Y.; Yoon, B.H.; Ko, B.S.; Shim, W.Y.; Kim, J.H. Xylitol production is increased by
expression of codon-optimized Neurospora crassa xylose reductase gene in Candida tropicalis.
Bioprocess Biosyst. Eng. 2012, 35, 191–198.
76. Zhang, J.; Zhang, B.; Wang, D.; Gao, X.; Hong, J. Xylitol production at high temperature by
engineered Kluyveromyces marxianus. Bioresour. Technol. 2014, 152, 192–201.
77. Huang, C.F.; Jiang, Y.F.; Guo, G.L.; Hwang, W.S. Development of a yeast strain for xylitol
production without hydrolysate detoxification as part of the integration of co-product generation
within the lignocellulosic ethanol process. Bioresour. Technol. 2011, 102, 3322–3329.
78. Misra, S.; Raghuwanshi, S.; Saxena, R.K. Evaluation of corncob hemicellulosic hydrolysate
for xylitol production by adapted strain of Candida tropicalis. Carbohydr. Polym. 2012, 92,
1596–1601.
79. Guo, X.; Zhang, R.; Li, Z.; Dai, D.; Li, C.; Zhou, X. A novel pathway construction in Candida
tropicalis for direct xylitol conversion from corncob xylan. Bioresour. Technol. 2013, 128,
547–552.
80. Hickert, L.R.; de Souza-Cruz, P.B.; Rosa, C.A.; Ayub, M.A.Z. Simultaneous saccharification
and co-fermentation of un-detoxified rice hull hydrolysate by Saccharomyces cerevisiae ICV
D254 and Spathaspora arborariae NRRL Y-48658 for the production of ethanol and xylitol.
Bioresour. Technol. 2013, 143, 112–116.
81. Li, Z.; Qu, H.; Li, C.; Zhou, X. Direct and efficient xylitol production from xylan by
Saccharomyces cerevisiae through transcriptional level and fermentation processing optimiza-
tions. Bioresour. Technol. 2013, 149, 413–419.
82. Nair, N.U.; Zhao, H. Selective reduction of xylose to xylitol from a mixture of hemicellulosic
sugars. Metab. Eng. 2010, 12, 462–468.
83. Martínez, E.A.; de Almeida, e Silva, J.B.; Giulietti, M.; Solenzal, A.I.N. Downstream process for
xylitol produced from fermented hydrolysate. Enzyme Microb. Technol. 2007, 40, 1193–1198.
Xylitol Biological Production 89

84. Gurgel, P.; Manchilha, I.M.; Pecanha, R.; Siqueira, J. F. Xylitol recovery from fermented
sugarcane bagasse hydrolysate. Bioresour. Technol. 1995, 52, 219–223.
85. Faveri, D.De.; Torre, P.; Perego, P.; Converti, A. Optimization of xylitol recovery by crystal-
lization from synthetic solutions using response surface methodology. J. Food Eng. 2004, 61,
407–412.
86. Tochampa, W.; Sirisansaneeyakul, S.; Vanichsriratana, W.; Srinophakun, P.; Baker, H.H.C.;
Chisti, Y. A model of xylitol production by the yeast Candida mogii. Bioprocess Biosyst. Eng.
2005, 28, 175–183.
87. Canilha, L.; Carvalho, W.; Silva, J.B.A.E. Xylitol bioproduction from wheat straw:
Hemicellulose hydrolysis and hydrolysate fermentation. J. Sci. Food Agric. 2006, 86,
1371–1376.
88. Carvalheiro, F.; Duarte, L.C.; Lopes, S.; Parajo, J.C.; Pereira, H.; Girio, F.M. Evaluation of
the detoxification of brewery’s spent grain hydrolysate for xylitol production by Debaryomyces
Downloaded by [University of Newcastle, Australia] at 10:53 05 January 2015

hansenii CCM1941. Process Biochem. 2005, 91, 93–100.


89. Vazquez, M.; Oliva, M.; Tellez-Luiz, S.J.; Ramirez, J.A. Hydrolysis of sorghum straw using
phosphoric acid: Evaluation of furfural production. Bioresour. Technol. 2007, 98, 3053–3060.
90. Mohagheghi, A.; Ruth, M.; Schell, D.J. Conditioning hemicellulose hydrolysate for fermen-
tation: Effects of overliming pH on sugar and ethanol yields. Process Biochem. 2006, 41,
1811–1896.
91. Roberto, I.C.; Mossato S.I.; Rodrigues, R.C.L. Dilute acid hydrolysis for optimization of xylose
recovery from rice straw in a semi-pilot reactor. Ind. Crops Prod. 2003, 17, 171–176.
92. Sues, A.; Millati, R.; Edebo, L.; Taherzadeh, M.J. Ethanol production from hexoses, pentoses
and dilute acid hydrolysate by Mucor indicus. FEMS Yeast Res. 2005, 5, 669–676.
93. Harerat, A.; Tellez-Luiz, S.J.; Ramirez, J.A.; Vazquez, M. Production of xylose from sorghum
straw using hydrochloric acid. J. Cereal Sci. 2003, 37, 267–274.
94. Dominguez, J.M.; Cao, N.; Gong, C.S.; Tsao, G.T. Dilute acid hemicellulosic hydrolysates from
corn cobs for xylitol production by yeast. Bioresource Technol. 1997, 6, 85–90.
95. Parajo, J.C.; Dominguez, H.; Dominguez, J.M. Charcoal adsorption of wood hydrolysates for
improving their fermentability: Influence of the operational conditions. Bioresource Technol.
1996, 57, 179–185.
96. Vazquez, M.; Oliva, M.; Tellez-Luis, S.J.; Ramirez, J.A. Hydrolysis of sorghum straw using
phosphoric acid: Evaluation of furfural production. Bioresource Technol. 2007, 98, 3053–3060.
97. Converti, A.; Borghi, M.D. Inhibition of the fermentation of oak hemicellulose acid-hydrolysate
by minor sugars. J. Biotechnol. 1998, 64, 211–218.

You might also like