You are on page 1of 11

G Model

PRBI-10195; No. of Pages 11 ARTICLE IN PRESS


Process Biochemistry xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Process Biochemistry
journal homepage: www.elsevier.com/locate/procbio

Review

Biotechnological production of xylitol from lignocellulosic wastes:


A review
Tiago Lima de Albuquerque a , Ivanildo José da Silva Jr. a , Gorete Ribeiro de Macedo b ,
Maria Valderez Ponte Rocha a,∗
a
Universidade Federal do Ceará, Departamento de Engenharia Química, Campus do Pici, Bloco 709, 60455-760 Fortaleza, CE, Brazil
b
Universidade Federal do Rio Grande do Norte, Departamento de Engenharia Química, Campus Universitário Lagoa Nova, CEP 59078-970 Natal, RN, Brazil

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

Article history: Several studies have sought to evaluate alternative routes for the production of xylitol, a polyol with
Received 20 May 2014 high employability in both the food and pharmaceutical industries. Large-scale production is typically
Received in revised form 25 June 2014 achieved by a chemical process of d-xylose hydrogenation. However, due to the conditions employed in
Accepted 15 July 2014
this procedure, it remains very costly. The biotechnological route has been the subject of intense research
Available online xxx
throughout the last decade because, apart from implying lower costs, many microorganisms can be used.
The employment of unconventional substrates as carbon sources for this production also stands out as a
Keywords:
very promising alternative. In addition to the possibility of being assimilated by many microorganisms,
Xylitol
Bioprocess
it becomes a good destination for the disposal of industrial waste. Thus, the aim of this review was to
Lignocellulosic wastes gather data on the biotechnological production of xylitol from agro-industrial waste, as well as to identify
Yeasts the best cultivation conditions (i.e. treatment of hydrolysate, pH, temperature and aeration) employed
Operating conditions in the process.
© 2014 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2. Xylitol applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3. Xylitol production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.1. Biotechnological production of xylitol from lignocellulosic waste . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.1.1. Waste utilization for xylitol production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.1.2. Microorganisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.2. Interferences obtained in the hydrolysis of lignocellulosic materials for xylitol production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.3. Growth conditions aimed at producing xylitol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.3.1. pH and temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.3.2. Aeration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.3.3. Carbon sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.3.4. Nitrogen sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.3.5. Bioreactor operation modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.4. Xylitol purification from fermented hemicellulosic hydrolysates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4. Future prospects and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

∗ Corresponding author at: Departamento de Engenharia Química, Universidade Federal do Ceará, Campus do Pici, Bloco 709, Fortaleza, CE 60455-760, Brazil.
Tel.: +55 85 3366 9611; fax: +55 85 3366 9610.
E-mail addresses: valderez.rocha@ufc.br, valponterocha@yahoo.com.br (M.V.P. Rocha).

http://dx.doi.org/10.1016/j.procbio.2014.07.010
1359-5113/© 2014 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010
G Model
PRBI-10195; No. of Pages 11 ARTICLE IN PRESS
2 T.L.d. Albuquerque et al. / Process Biochemistry xxx (2014) xxx–xxx

1. Introduction price. The study of these researchers concluded that if a substantial


portion of hemicellulose used is converted into xylitol, this product
Xylitol (C5 H12 O5 ) is a polyol that has a hydroxyl group attached would generate all the required income to maintain the process.
to each carbon atom in its chain, and possesses a high sweetening Furthermore, in addition to the use of lignocellulosic materi-
power, presenting 40% fewer calories than sucrose. It was discov- als for the production of xylitol and ethanol, some studies [16–18]
ered in 1891 by chemist Emil Fischer (German, 1852–1919) and his also show the use of sugars obtained from such materials for the
team, who obtained it in syrup form from the reaction of xylose production of polyunsaturated fatty acids with medicinal interest.
with sodium amalgam [1]. The main characteristic of this polyol Thus, the aim of this article was to review alternative processes
is its high cooling power inherent in its high endothermic heat for the biotechnological production of xylitol from agro-industrial
(34.8 cal g−1 ) [2]. It has been vastly cited in the literature for its waste. The diversity of microorganisms used in the process was
organoleptic characteristics and health benefits such as: high sol- studied, as well as the optimum conditions for polyol production.
ubility [3], low glycaemic rates [4], lack of carcinogenicity [5] and
cariostatic properties [6]. Furthermore, xylitol does not participate
2. Xylitol applications
in Maillard reactions (which can cause the formation of browning
compounds in food) and thus does not reduce the nutritional value
In the European Economic Community the safety of xylitol use
of the proteins present [7]. These characteristics confirm its appli-
has been recognized since 1984, and the agency that regulates food
cability in food production processes, such as candies, caramels,
and drugs in the United States (Food and Drug Administration –
chocolates, ice creams, jellies, marmalades and beverages [7].
FDA) has classified it as “Generally Recognized as Safe” (GRAS)
Industrial production of xylitol has been on-going for almost
since 1986 and “Safe for Teeth” since 1994. Xylitol is extremely
four decades and, since the early years, solutions of purified d-
well tolerated by the human organism when taken in doses over a
xylose typically undergo a catalytic hydrogenation process, under
certain time frame with a maximum amount of 20 g, with the total
conditions of high-temperature (80–140 ◦ C) and pressure (up to
amount consumed not exceeding 60 g per day, since the intake of
50 atm) [8], until polyol formation is achieved (chemical process).
higher doses presents laxative effects [2]. The literature includes
Due to the operating conditions and the need for purity in the
several studies on the use of xylitol as a replacement for sucrose
employed xylose, this traditional process becomes quite expensive.
in certain foods aimed at caloric reduction, such as in cupcakes
Several studies have sought alternatives to chemical routes, with
[19], cookies [20], chocolate [21] and chewing gums [22]. The use
particular attention paid to biotechnological processes, which led
of xylitol in the pharmaceutical industry is now widespread, mainly
them to become a major focus in scientific studies. Aiming to further
due to its proven anticariogenic properties, which can reduce tooth
reduce processing costs and, in addition, searching for solutions to
decay by up to 100%. Several studies have shown the effective-
the recycling of agro-industrial waste, the potential utilization of
ness of this sugar-alcohol against the formation of oral biofilms,
renewable raw materials is being evaluated, such as bamboo [9],
especially when confronting the bacterium Streptococcus mutans
vegetable residues [10] and sunflower stalks [11]. These wastes, in
[23,24]. Some authors claim that the cariogenic characteristics of
turn, can be hydrolysed, generally by diluted acids, with further
xylitol against this bacterium are due to a decrease in glucosyl-
release of d-xylose available for microbial xylitol conversion.
transferase expression in the microorganism [23]. Other authors
Xylitol production is expanding in a global level [12] and it is
state that it changes the cell structure by turning the polysaccharide
currently estimated a strong demand for this product in the global
that surrounds them into a diffuse and sparse cell wall [24]. Apart
market, of more than 125,000 tons per year, with a value that is
from S. mutans, other studies have shown the effectiveness of xylitol
relatively high ($4.50–5.50 per kg for bulk purchase by pharma-
employment against other microorganisms that are harmful to oral
ceutics and food companies and $20.00 per kg in supermarkets)
health, such as: Streptococcus pneumoniae and Haemophilus influen-
and therefore, encouraging the search for production processes
zae [24], Staphylococcus aureus [25] and Pseudomonas aerugionosa
with lower costs. The industrial production of the substance is
[26]. Besides fighting tooth decay, xylitol is recognized for being
performed chemically by a catalytic hydrogenation of the purified
able to contribute to tooth calcification, which, in most cases, is
xylose. This method utilizes specialized and expensive equipment
caused by ageing [27]. It has also been reported to be effective in the
to achieve the high pressure and temperature conditions necessary
treatment of several diseases, for instance: diabetes [4], anaemia
for the process [12–15]. Xylitol has a 12% share of the total polyols
[28], acute otitis media [29] and osteoporosis [30].
market with rapid growth worldwide due to an increasingly health
conscious consumer and fast increase in chewing gum sales [14,15],
and it is approved for food use in over 50 countries [15]. 3. Xylitol production
Studies that aim exclusively on economic evaluation of xylitol
production by bioprocesses routes are not found in the literature, as Xylitol is an intermediate product formed during carbohydrate
well as studies for comparing its costs to the chemical process. It is metabolism in mammals, including man, with a human being under
known that renewable carbon sources, such as industrial residues, normal conditions able to produce 5–15 g of xylitol per day [31].
can be converted into substances with higher added value, both The commercial value of xylitol restricts its employment to cer-
chemically and by biotechnological routes, and the choice of the tain foods and medicines, and due to this fact, new production
process should be based on a detailed assessment of sustainability, processes need to be devised in order to cheapen the costs inher-
resource availability and transport logistics [14]. ent in it [8]. Xylitol production may be performed by chemical
Franceschin et al. [15] assessed the bioconversion of rye straw means, by xylose-hydrogenation reactions, or through microbio-
into ethanol and xylitol, evaluating this integrated process econom- logical routes, in which specific microorganisms convert xylose to
ically. In that study, it was concluded that the capital required for xylitol. Apart from these, the enzymatic production process has
the execution of the bioprocess reached a total amount of about $40 raised the interest of scientists given the fact that it is a biotechno-
million, with 25.9% and 29.0% of this amount approximately corre- logical alternative that can convert up to 100% of xylose to xylitol,
sponding to the costs with the fermentation and recovery of xylitol, since no metabolic diversions for cell maintenance occur [32,33].
respectively. It is noteworthy that the economic analysis is strictly Industrial-scale production first started in 1975 in Finland, by
dependent on the marketing price of xylitol, and it was observed the Finnish Sugar Co. Ltda., in Helsinki. This process, which utilizes
that the market would only be saturated with xylitol after a long chemical routes, is carried out in high-pressure (31–40 atm) and
period, which would occasionally lead to a decrease in its selling high-temperature (100–130 ◦ C) reactors and lasts from three to 5 h,

Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010
G Model
PRBI-10195; No. of Pages 11 ARTICLE IN PRESS
T.L.d. Albuquerque et al. / Process Biochemistry xxx (2014) xxx–xxx 3

Fig. 1. Xylitol production by chemical route.

depending on the employed conditions. The currently used large-


scale production is divided into several steps (Fig. 1): Step 1. Acid
hydrolysis of the xylan-rich natural material; Step 2. Hydrolysate
purification up to the point where a solution of pure xylose is Fig. 2. Diagram of process for bioproduction of xylitol from lignocellulosic material.
obtained; Step 3. Catalytic hydrogenation of pure xylose to xyli-
tol with a catalyst (a Ni-Al2 alloy); Step 4. Purification of the xylitol and the process of converting xylose to xylitol has a signifi-
solution obtained; Step 5. Xylitol crystallization [34,35]. cant economic role regarding biomass employment [39]. Thus,
The high temperatures and pressures employed required spe- the development and optimization of methods both for obtain-
cialized and expensive equipment to achieve conditions necessary ing xylose from lignocellulosic wastes and for then converting this
for the process [12–15], not to mention the numerous purification sugar into products with higher added value, such as xylitol, is
steps required, render the process of obtaining xylitol expensive of great interest. Fig. 2 shows a simplified flowchart of the steps
[12,13]; therefore, new means of production that can be as effec- performed to produce xylitol using lignocellulosic materials.
tive as many of the traditional methods have been sought [36]. The employment of waste generated by agribusiness has been
Currently, Danisco (industry of DuPont Company) is a major global widely investigated for potential use of its raw materials in xyli-
supplier of xylitol, mainly from China, which it produces via catal- tol production. It is known that these materials, after undergoing a
ysis using xylose obtained from hardwood sources [15]. They have pretreatment, can release fermentable sugars such as d-xylose, the
recently developed an integrated process with the pulp and paper main carbon source for xylitol production. Table 1 shows the variety
industry. However conversion of xylose to xylitol is still achieved of waste that has been evaluated throughout the last decade for this
via a chemical route. process. A pretreatment or hydrolysis step is performed initially
on these raw materials to expose their constitutive sugars so they
3.1. Biotechnological production of xylitol from lignocellulosic can be used later as a carbon source by different microorganisms,
waste enabling these raw materials to be susceptible to biotechnological
usage. Many of the pretreatments or hydrolysis steps that release
3.1.1. Waste utilization for xylitol production xylose from the hemicellulosic structure employ acids at dilute
Xylose, along with glucose, is a major sugar derived from the concentrations, such as phosphoric acid [11] and sulphuric acid
hydrolysis of lignocellulosic biomass, among other sugars such as [38,40]. The waste is then subjected to these acidic solutions at
mannose, galactose, arabinose and rhamnose (all with countless temperatures generally above 100 ◦ C.
applications in both pharmaceutical and food industry) [37,38]. It is noteworthy that during the hydrolysis of lignocellulosic
This substance is the second most abundant sugar found in nature, materials, many compounds that can inhibit microorganisms can

Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010
4

PRBI-10195; No. of Pages 11


G Model
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010
Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.

Table 1
Xylitol production by different yeast strains and operational conditions using lignocellulosic waste as feedstock.

Microorganism Waste Culture conditions Maximum xylitol Fermentation mode Yield (g g−1 ) Productivity Reference
production (g L−1 ) (g L−1 h−1 )

pH Temperature (◦ C) Time (h)

T.L.d. Albuquerque et al. / Process Biochemistry xxx (2014) xxx–xxx

ARTICLE IN PRESS
Candida athensensis SB18 Vegetable waste 7.0 30.0 102 100.1 Batch mode (bioreactor) 0.81 0.98 [10]
Candida guilliermondii Sugar cane bagasse 5.5 30.0 120 50.5 Batch mode (Erlenmeyer flasks) 0.81 0.60 [37]
FTI20037
Candida guilliermondii Rice straw – 30.0 116 66.1 Batch mode (bioreactor) 0.84 0.17 [71]
FTI20037
Candida magnoliae Bamboo culm – 30.0 30 10.5 Batch mode (bioreactor) 0.59 0.42 [9]
Candida tropicalis As 2.1776 Corncob 5.5 30.0 120 96.5 Fed-batch mode (bioreactor) 0.83 1.01 [43]
Candida tropicalis BCRC 20520 Wood sawdust 5.0 30.0 96 41.4 Batch mode (Erlenmeyer flasks) 0.70 0.43 [77]
Candida tropicalis CCTCC Corncob 6.0 35.0 14 38.8 Fed-batch mode (bioreactor) 0.70 0.46 [36]
M2012462
Candida tropicalis HDY-02 Corncob 7.0 35.0. 78 58 Fed-batch mode (bioreactor) 0.73 0.74 [78]
Candida tropicalis JH030 Rice straw 6.0 30.0 80 31.1 Batch mode (Erlenmeyer flasks) 0.71 0.44 [90]
Candida tropicalis NBRC 0618 Olive pruning waste 5.0 30.0 25 53 Batch mode (bioreactor) 0.49 – [40]
Candida tropicalis W103 Corncobs 6.0 35.0 70 68.4 Fed-batch mode (bioreactor) 0.70 0.95 [91]
Debaryomyces hansenii Sugar cane bagasse 6.0 40.0 156 71.2 Batch mode (immobilized cells) 0.82 0.46 [39]
Debaryomyces hansenii CCMI Grain brewery 5.5 30.0 72 24.0 Batch mode (Erlenmeyer flasks) 0.57 0.51 [88]
941
Debaryomyces hansenii NRRL Vine waste 6.0 30.0 116 27.5 Batch mode (Erlenmeyer flasks) 0.54 – [89]
Y-7426
Hansenula polymorpha ATCC Sunflower stalks 5.5 30.0 169 0.31 Batch mode (bioreactor) 0.0023 – [11]
34438
Kluyveromyces marxianus CCA Cahew apple bagasse 6.0 30.0 96 6.76 Batch mode (Erlenmeyer flasks) – – [38]
510
Kluyveromyces marxianus CE Cashew apple bagasse 4.5 40.0 72 4.8 Batch mode (Erlenmeyer flasks) – – [87]
025
Pichia stipitis NRRL Y-30785 Corn stover 5.6 30.0 72 12.5 Batch mode (Erlenmeyer flasks) 0.61 0.18 [76]
G Model
PRBI-10195; No. of Pages 11 ARTICLE IN PRESS
T.L.d. Albuquerque et al. / Process Biochemistry xxx (2014) xxx–xxx 5

be formed. The main such compounds generated are furfural, intermediate metabolite during the metabolism of d-xylose. Xylose
hydroxymethylfurfural, aliphatic acids, phenolic compounds and reductase (XR) is a typical NADPH-dependent enzyme, whereas
vanillin [41–43]. Inhibitors present or formed from hydrolysates xylitol dehydrogenase (XDH) requires NADP+. Thus, conversion of
can limit consumption of the carbon source and may even stop the d-xylitol occurs in two steps, a reduction step followed by an oxi-
fermentation process from happening [41]. dation step. d-Xylose is first reduced to d-xylitol by NADPH and
In many bioprocesses some treatment of the hydrolysate is subsequently this metabolite is oxidized to d-xylulose by NADP+
mandatory in order to eliminate these fermentative interferences. (Fig. 3), with these two reactions considered limiting for d-xylose
This is mainly due to the fact that even in the smallest amounts fermentation and d-xylitol production [10,51].
some of them can prevent microbial growth. Several detoxifica- Numerous microbial species have a metabolic system with
tion methods are used to remove toxic compounds from plant NADPH-dependent XR and NAD+-dependent XDH as cofactors and
biomass hydrolysates, such as enzyme biological treatments (lac- these enzymes are induced by xylose [52]. Among the genera capa-
case), physical treatments, evaporation and extraction, chemical ble of producing xylitol, Candida is the most studied and stands as
treatments with alkaline agents (NaOH, CaO, Ca(OH)2 ) and physico- a promising microorganism. With this type of microorganism, oxy-
chemical treatments involving adsorption with activated carbon gen availability is the most important factor in xylitol production
and ion-exchange resins. Activated carbon [39] and ion-exchange from d-xylose. Under limited oxygen conditions, oxidative phos-
resins [37] have been used quite successfully to decrease or even phorylation cannot re-oxidize all the NADH generated. Thus, the
completely eliminate the major fermentation inhibitors. Addition- intracellular NADH concentration increases and leads to xylitol
ally, treatments employing microorganisms for detoxification of accumulation.
lignocellulosic hydrolysates have received considerable attention Yablochkova et al. [53] studied the activities of XR and XDH in 11
in recent years. Fonseca et al. [42] studied biological detoxifi- different species of yeast (including Candida, Kluyveromyces, Pichia,
cation of different hemicellulosic hydrolysates (corn fibre, straw Torulopsis and Pachysolen) and found that Candida tropicalis Y-456
and sugarcane bagasse, coffee bean husks) using yeast Issatchenkia had the highest specific activity of XR: 6.57 ␮mol min−1 mgprotein −1 .
occidentalis CCTCC M 206097. The greatest clearance of inhibitors According to literature reports, microorganisms with high XR activ-
together with the lowest loss of reducing sugars was observed ity and NADPH-dependent are potential producers of xylitol from
when the hydrolysate was concentrated fivefold before the bio- d-xylose.
logical treatment. The total reductions observed within 24 h of the Misra et al. [54] also investigated xylitol production by 18 strains
experiment were: syringaldehyde (66.67%), ferulic acid (73.33%), of yeasts and, coincidentally, the Candida yeasts showed better pro-
furfural (62%) and 5-HMF (82%). Mussatto and Roberto [2] treated duction performance. A strain of C. tropicalis achieved production
rice straw hydrolysate with five types of activated carbon and of 12.11 g L−1 of xylitol using 50 g L−1 initial xylose concentration
observed that with carbons presenting lower particle sizes, greater at 30 ◦ C, pH 5 and 200 rpm for 72 h.
reductions in toxic compounds could be achieved. Studies have shown variable values for xylitol production by
biotechnological processes using yeasts from industrial residues
3.1.2. Microorganisms (Table 1). It can be readily observed that the results are very
Different species of microorganisms, as well as several alter- discrepant and are related to the different microbial species and
native raw materials, have been evaluated for the production of growth conditions involved (carbon and nitrogen sources, pH, aer-
xylitol by biotechnological methods, but most studies are restricted ation, and so forth). Controlling these conditions is therefore of
to certain species of Candida. The biotechnological production of fundamental importance for optimizing the xylitol production pro-
xylitol, being less expensive, has the potential to be an advanta- cess.
geous alternative to the chemical process that is conventionally Zikou et al. [17] employed a Thamnidium elegans strain that pro-
employed. Various types of microorganisms, including bacteria duced a maximum xylitol quantity of 31 g L−1 during growth on
and filamentous fungi, can be used in the process. Rangaswamy blends on xylose and glucose.
and Agblevor [44], for example, selected 17 bacterial cultures of Zhang et al. [10] studied xylitol production using the yeast Can-
the genera Serratia, Cellulomonas and Corynebacterium for poten- dida athensensis SB18 in agitated flasks and bioreactor performed
tial xylitol production (with trials carried for 48 h at 30 ◦ C and in both batch and fed-batch systems. The concentration, produc-
130 rpm in Erlenmeyer flasks), and found out that Corynebac- tivity and yield of xylitol obtained in this study were 100 g L−1 ,
terium sp. B-4247 was the strain with the highest production rates 0.81 g g−1 and 0.98 g L−1 h−1 , respectively, using hydrolysed hor-
(10.05 g L−1 ). Cirino et al. [45] described the production of xyli- ticultural waste containing 200 g L−1 xylose.
tol using Escherichia coli W3110, a genetically modified strain, and Villarreal et al. [55] evaluated the xylitol production from euca-
achieved a production of up to 38 g L−1 of xylitol (30 ◦ C, 250 rpm lyptus hemicellulosic hydrolyzate by Candida guilliermondii and,
for 80 h). Other genetically modified bacteria are also promising under the best employed operating conditions, 32.7 g L−1 of xyli-
for xylitol production, such as Bacillus subtilis [46] and Corynebac- tol were produced after 48 h fermentation, which correspond to
terium glutamicum [47]. Studies with filamentous fungi for xylitol 0.68 g L−1 h−1 volumetric productivity.
production are scarce, but some researchers have reported positive In a study done by Rocha et al. [38], Kluyveromyces marxianus
results with Hypocrea jecorina [48] and Trichoderma reesei [49]. In CCA 510 produced ethanol and xylitol, using cashew apple bagasse
the latter case, the authors studied a genetically modified strain of T. hydrolysate as culture medium and it reaching a concentration of
reesei to produce xylitol from barley straw pre-treated with organic 11.89 g L−1 and 6.76 g L−1 , respectively, at 96 h. More study exam-
solvents, with the process being conducted for 168 h at 30 ◦ C. ples related will be cited in the following topics.
Iverson et al. [50] evaluated the increasing catabolic reducing
power output (NADH) of glucose catabolism for reduction of xylose 3.2. Interferences obtained in the hydrolysis of lignocellulosic
to xylitol by genetically engineered E. coli AI05. The obtained strain materials for xylitol production
could in theory simultaneously uptake glucose and xylose, and util-
izes glucose as a source of reducing power for the reduction of Lignocellulosic materials, including cellulose, hemicellulose and
xylose to xylitol, under anaerobic conditions. lignin, are the most abundant materials on Earth. Hemicellulose is
Yeasts are still the primary focus of studies that aim to produce the second most common component in plant cell walls (cellulose
xylitol from hemicellulosic hydrolysates and different species are being the first) and is a potential substrate for ethanol and xylitol
cited in Table 1. These microorganisms can produce xylitol as an production [56]. Hemicellulose degradation by chemical processes

Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010
G Model
PRBI-10195; No. of Pages 11 ARTICLE IN PRESS
6 T.L.d. Albuquerque et al. / Process Biochemistry xxx (2014) xxx–xxx

Fig. 3. d-Xylose metabolic pathway in yeasts.

yields xylose as the main fraction, as well as variable fractions of making this method unsuitable for pretreatment of olive pruning
arabinose, mannose, galactose and glucose. Apart from these afore- hydrolysates. As for the use of activated carbon in this hydrolysate,
mentioned sugars, chemical hydrolysis of lignocellulose can release it was shown to be an excellent alternative to reduce inhibitors (46%
35 other byproducts that act as microbial inhibitors [57]. of acetic acid, 81% of phenolic compounds and 98% of total furans),
Inhibitors can be divided into furan derivatives, such as fur- with the possibility of regenerating the carbon employed to reduce
fural and 5-hydroxymethylfurfural (HMF), phenolic compounds, the cost of the process.
weak acids (levulinic and formic acid) and ions of heavy metals
(nickel, aluminium, chromium, etc.) [58]. Furfural and HMF are two 3.3. Growth conditions aimed at producing xylitol
furan derivatives formed by the hydrolysis of sugars, hexoses and
pentoses, respectively. These furans are present at relatively high Several factors may influence microbial growth and, in order to
concentrations in the hydrolysate and can act as potential microor- optimizing these, conditions should be adjusted according to the
ganism inhibitors. Furans are probably the most important group specificity of each microorganism. The most extensively studied
of inhibitors, since the fermentability of hydrolysates obtained by factors in microbial growth that show some influence in biopro-
acidification is inversely related to the concentration of these com- cesses are temperature, pH, aeration rate, inoculum concentration
pounds. In addition, phenolic compounds cause loss of integrity in and nutrients (carbon, nitrogen, and vitamin sources, etc.). Proper
biological membranes, affecting their ability to serve as selective control of these variables is of great importance for good per-
barriers and enzyme matrices [59]. Acetic acid, the main aliphatic formance in obtaining high-quality products with high efficiency.
compound present, is released from acetyl-hemicellulose groups. There have been many studies on the optimal environmental con-
In turn, formic acid is a product of HMF degradation. ditions for yeast growth aimed at biotechnological production of
Several methods to detoxify the culture media can be employed, xylitol, and these are discussed below.
such as physical (detoxification mediated by evaporation), chem-
ical (addition of Ca(OH)2 ), adsorption (using ion-exchange resins 3.3.1. pH and temperature
and activated carbons) and biological (microbial and enzymatic) to Most studies employing yeasts in biotechnological processes
fight inhibitors of xylitol production [56]. have been carried out by adjusting the temperature in the range
Mateo et al. [60] studied several methods to detoxify acid of 30–37 ◦ C, which proves to be optimal for these microorganisms
hydrolysates of olive pruning residues (liming with NaOH, CaO and in xylitol production [36,61].
Ca(OH)2 and adsorption with activated carbon). When NaOH was In some cases, however, higher temperatures may be employed;
used, the decrease in phenolic compounds was minimal (28–31%) for instance, the thermotolerant characteristic of the yeast K.
for two different alkaline detoxification processes, and a slight marxianus makes it stand out in fermentation studies when com-
reduction in acetic acid (<7%) was achieved, although no sig- pared to other microorganisms. This is because, in addition to its
nificant sugar loss was observed. The calcium-based treatments, enzymatic reactions being accelerated, the risk of contamination
though, had a notable effect, removing more lignin-derived com- by other microorganisms is diminished [62]. Rodrussamee et al.
pounds (56–71%), acetic acid (about 50%) and total furan content [63] studied the potential of K. marxianus DMKU3-1042 for xyli-
(approximately 70%). However, a high loss of sugar also occurred, tol and ethanol production from sugars present in hemicellulose

Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010
G Model
PRBI-10195; No. of Pages 11 ARTICLE IN PRESS
T.L.d. Albuquerque et al. / Process Biochemistry xxx (2014) xxx–xxx 7

hydrolysates at high temperatures (30 ◦ C, 40 ◦ C and 45 ◦ C). Cell this process is redox imbalance, which is caused by the difference
growth and sugar consumption were observed at all temperatures in the enzymes’ preference for coenzymes, NADPH in the case of XR
studied, confirming that the strain can grow at high temperatures. and NAD+ in the case of XDH. This redox imbalance is the key rea-
It is noteworthy that cell growth showed diminished intensity son for the activity by which the yeast accumulates xylitol during
as the temperature increased, though without ceasing. At 30 ◦ C, anaerobic cultivation [69]. Fig. 3 shows the key reactions in d-xylose
ethanol production was 2.5 g L−1 at 72 h, and xylitol production metabolism during xylitol production using xylose as substrate.
was 4.3 g L−1 after 48 h. Xylitol production was actually favoured Sirisansaneeyakul et al. [70] studied xylitol production by Candida
at 40 ◦ C obtained a concentration of 7.0 g L−1 . magnoliae TISTR 5663 in a fed batch and under oxygen-limited
Three strains of K. marxianus (IMB2, IMB3 and IMB4), isolated conditions (1 vvm). Xylitol yields of 0.727 g g−1 , 0.719 g g−1 and
from an Indian distillery were characterized for their growth in 0.720 g g−1 were obtained with average biomass concentrations
xylose at high temperatures (40–45 ◦ C) and different pH values (4.5, of 21 g L−1 , 48 g L−1 and 50 g L−1 , respectively, under the evaluated
5.0 and 5.5). It was observed that the lower temperature (40 ◦ C) at conditions (stirring speed 300 rpm, pH 7.0 and 30 ◦ C).
pH 5.5 was better for ethanol and xylitol yields, where values of El-Baz et al. [68], in turn, evaluated the potential of C. tropicalis
2.08 g L−1 and 7.36 g L−1 , respectively, were reached using the IMB4 in a synthetic medium containing 20 g L−1 xylose as carbon source
strain after 96 h of cultivation [64]. In turn, the optimal tempera- and concluded that xylitol production was increased when medium
ture for Kluyveromyces sp. IIPE453 growth was at 50 ◦ C and pH 5.0, oxygenation was reduced (by doubling the volume of medium in
reaching a maximal production of ethanol and xylitol of 82 g L−1 the 250 mL Erlenmeyer flask from 20 mL to 40 mL). Mussatto and
and 11.5 g L−1 at 48 h, respectively [65]. Roberto [71] investigated xylitol production by C. guilliermondii FTI
Srivani and Setty [66] attempted to find the optimal environ- 20037 in a bioreactor from rice straw hemicellulosic hydrolysate
mental conditions for xylitol production by Candida parapsilosis at different agitation rates (200 rpm, 300 rpm and 500 rpm). These
NCIM-3323, studying variations in temperature (25–35 ◦ C) and ini- researchers concluded that xylose bioconversion to xylitol by the
tial pH (3–6). Over the range of initial pH studied, it was observed yeast depends on the stirring speed, reaching a maximum yield of
that xylitol productivity was maximal when the initial pH was 0.84 g g−1 at 300 rpm.
3.5. Above this, there was a decrease in production. At the tem-
perature ranges, it was further observed that the productivity of 3.3.3. Carbon sources
xylitol increased until the temperature reached 30 ◦ C and, from The use of lignocellulosic materials for the production of
this temperature on, a drop in production occurred. Ramesh et al. high-added value bioproducts is promising, as they represent an
[67] investigated xylitol production from corncob hemicellulosic abundant and renewable supply of carbon sources. Lignocellulosic
hydrolysate by Debaryomyces hansenii var. hansenii (MTTC 3034) materials such as corncobs, rice straw, sugarcane bagasse, cashew
using a statistical optimization of response surface approach. In bagasse, sawdust and oat hulls, among others (Table 1), represent
this study, the optimum temperature and pH for production of the an important and cheap source of microbial substrates. The hemi-
polyol were 31.8 ◦ C and 7.25 g L−1 , respectively. cellulosic fraction can be hydrolysed to xylose and then fermented
With C. tropicalis, several studies have considered temperatures to xylitol [9].
of 30–35 ◦ C for xylitol production. Ping et al. [36], for example, With respect to xylose utilization inside the cells, this compound
utilized C. tropicalis CCTCC M2012462 for xylitol production from can be either metabolized through the phosphoketolase reaction,
hydrolysed corncob at 35 ◦ C, reaching 38.8 g L−1 after 84 h of bio- which is the most efficient pathway yielding around 1.2 mole of
processing. Misra et al. [61] also used a corncob hydrolysate to acetyl-CoA per 100 g of xylose (∼0.66 mole) utilized, or the pentose
obtain 11.89 g L−1 xylitol at 30 ◦ C using a strain of C. tropicalis. El- phosphate pathway, where around 1.0 mole of acetyl-CoA is formed
Baz et al. [68] studied the production of xylitol in synthetic media per 100 g of xylose utilized [72].
at 30 ◦ C for C. tropicalis, obtaining a maximum concentration of Several studies have evaluated the use of corncob hydrolysate
36.25 g L−1 . as a carbon source for xylitol production [36,61,64]. Kamat et al.
[73] isolated a yeast strain, Cyberlindnera (Williopsis) saturnus, from
3.3.2. Aeration mangrove forests that was capable of producing xylitol (29.1 g L−1 )
Yeasts do not form a homogeneous group regarding their energy from detoxified corncob hydrolysate (composition: 65 g L−1 xylose,
efficiency. The destination of pyruvate, a glycolytic intermediate, 13 g L−1 glucose and 6.3 g L−1 arabinose). Xylitol production started
determines the type of metabolism. When all the pyruvate pro- at 48 h of fermentation and continued up to 144 h. Glucose, how-
duced is converted via the tricarboxylic acid cycle (respiration), ever, was consumed mostly in the first hours of microorganism
oxidative metabolism occurs and, if it is reduced to ethanol or other incubation, with total assimilation completed by 24 h. Misra et al.
compounds, oxy-reductive metabolism (fermentation) happens, [61] also studied the potential of corncob hydrolysate for xylitol
with the two pathways also able to occur simultaneously. Aeration production using a strain of C. tropicalis. These authors achieved a
is seen as a major factor in the experimental production of xylitol, maximum extraction of 20.92 g L−1 xylose using 1% (v/v) sulphuric
and the amount of oxygen available is one of the determining fac- acid in the corncob hydrolysis. Concentrating the hydrolysate up
tors for which pathway xylose will be shunted to, fermentation or to 52.71 g L−1 , in relation to xylose concentration, 15.19 g L−1 of
respiration, in turn regulating the carbon consumption balance for xylitol was obtained at 60 h of culture. Ping et al. [36] used a concen-
both growth and bioconversion [68]. trated non-detoxified corncob hydrolysate for xylitol production by
It is evident that yeasts require little oxygenation in the cul- C. tropicalis CCTCC M2012462, reaching a maximum of 38.8 g L−1
ture medium for d-xylose reduction and have specific rates for each xylitol.
species [68]. Barley straw was evaluated for its employment in xylitol pro-
To use and metabolize d-xylose, it must be converted to xylu- duction by a genetically modified T. reesei [49]. Two hydrolysis
lose. Xylulose is then phosphorylated to xylulose 5-phosphate by methods were tested in the raw material (one by adding organic
xylulokinase (XK), with the latter being metabolized by the pentose solvents and the other by addition of NaOH). The best xylitol pro-
phosphate pathway. In bacteria, d-xylose is converted directly to d- duction (13.22 g L−1 ) was achieved with the use of barley straw
xylulose by xylose isomerase (XI). In eukaryotes such as yeasts and treated with organic solvents plus the addition of 2% xylose.
filamentous fungi, d-xylose is converted to d-xylulose by means of Detoxified bamboo culm acid hydrolysate (19 g L−1 xylose) was
reduction and oxidation, which are catalyzed by xylose reductase fermented by C. magnoliae FERM P-16522 in a study by Miura
(XR) and xylitol dehydrogenase (XDH). The main limiting factor in et al. [9]. The authors [9] concluded that this raw material has the

Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010
G Model
PRBI-10195; No. of Pages 11 ARTICLE IN PRESS
8 T.L.d. Albuquerque et al. / Process Biochemistry xxx (2014) xxx–xxx

potential to be utilized in xylitol production, reaching a maximum and yeast extract had a significant influence, with optimal concen-
concentration of 10.5 g L−1 for the polyol. trations of these components of 5.0 g L−1 and 4.6 g L−1 , respectively.
Martinez et al. [11] investigated the production of xylitol
and ethanol by Hansenula polymorpha from sunflower stems 3.3.5. Bioreactor operation modes
hydrolysed with phosphoric acid. The results of this research Most studies about the xylitol production showed in the litera-
showed that, for an initial concentration of total reducing sugars ture are conducted in batch mode (flasks or bioreactors). However,
of 13.3 g L−1 obtained in the process, xylitol and ethanol yields some experiments approach the production of xylitol in other oper-
were 0.023 g g−1 and 0.14 g g−1 , respectively. Rocha et al. [38], ating modes, as in fed batch mode (Table 1), which can achieve
in a study on ethanol production from non-detoxified cashew higher yields.
bagasse hydrolysate (29.08 g L−1 glucose, 24.48 g L−1 xylose and Silva et al. [79], for example, used fluidized bed reactor operated
11.33 g L−1 arabinose), observed simultaneous xylitol production in semi-continuous mode for xylose-to-xylitol conversion by C.
by the yeast K. marxianus CE025. Xylitol production was high- guilliermondii immobilized on porous glass. The reactor was loaded
est at 72 h of fermentation, reaching a concentration of 4.77 g L−1 . with 200 g of immobilized system and operated at a volume of
Some recent researches have also been dedicated to evaluating 1.4 L in semi-continuous mode. The fermentation was performed
xylitol production in synthetic media containing xylose as the in seven cycles (total time of 672 h), using fresh medium for each
main carbon source. Srivani and Setty [66] studied the optimiza- one. The first and second cycles provided the highest values of
tion of several parameters (pH, temperature and initial xylose xylitol yield of 0.79 g g−1 and 0.57 g g−1 , respectively, and volumet-
concentration) for xylitol production from xylose fermentation ric productivity of 0.52 g L−1 h−1 and 0.60 g L−1 h−1 , respectively, a
by C. parapsilosis NCIM-3323. It was concluded that the maxi- decrease being observed afterwards.
mum production of xylitol (28.14 g L−1 ) was achieved with the Santos et al. [80] studied the influence of aeration rate and car-
following values for pH, temperature and initial xylose con- rier concentration on xylitol production from sugarcane bagasse
centration: 3.5, 30 ◦ C and 60 g L−1 , respectively. Vajzovic et al. hydrolysate in immobilized-cell fluidized bed reactor. The results
[74] used a synthetic medium (containing glucose or xylose, obtained showed that carrier concentration (Cs) had a negative
30 g L−1 ) to evaluate xylitol production in the presence of cer- influence on xylitol yield based in the consumption of xylose
tain inhibitors (furfural, 5-hydroxymethylfurfural and acetic acid) (YP/S ) and volumetric productivity (QP ), whereas aeration rate (AR)
and the results obtained by these authors showed that high con- had a positive influence on QP and a negative influence on YP/S .
centrations of inhibitors (above 3 g L−1 ) negatively affected xylitol With an AR = 0.093 min−1 and Cs = 62.5 g L−1 , the YP/S value was
production. low (0.25 g g−1 ), but the QP was highest (0.44 g L−1 h−1 ), probably
because cell metabolism was faster when more oxygen was avail-
able to the yeast. When the highest levels of the AR and Cs were
3.3.4. Nitrogen sources used, the xylitol yield and concentration were the lowest and the
Various nitrogen sources are being investigated in biotechno- cell concentration was the highest, suggesting that cell metabolism
logical studies to optimize the growth of microorganisms and the was directed preferentially towards biomass production instead of
production of metabolites of interest. Among the most studied xylitol accumulation.
organic sources are peptone, yeast extract and casamino acid [75]. Salgado et al. [81] evaluated the coupling of two sizes of
Among the inorganic sources are ammonium sulphate and phos- CSTR-type bioreactors for sequential lactic acid and xylitol pro-
phate, sodium nitrate and urea [9,41]. duction from hemicellulosic hydrolysates of vineshoot trimmings.
There have been many studies investigating the influence By connecting two reactors of 2 L and 10 L, operational condi-
of urea utilization as a nitrogen source for xylitol production. tions were set up for the sequential production of lactic acid and
Rodrigues et al. [76] investigated the effects of employing urea xylitol in continuous process, considering the dependence of the
and ammonium sulphate as nitrogen sources on xylitol pro- main metabolites and process parameters on the dilution rate.
duction by Pichia stipitis YS-30. The medium, containing corn The maximum production of xylitol (5.1 g L−1 ), volumetric xyli-
stover hydrolysate, was supplemented with either 5.0 g L−1 urea tol productivity (QP = 0.218 g L−1 h−1 ), volumetric rate of xylose
or 5.0 g L−1 ammonium sulphate. It was observed that when urea consumption (QS = 0.398 g L−1 h−1 ) and product yield (0.55 g g−1 )
was used instead of (NH4 )2 SO4 , xylose consumption and ethanol were achieved at an intermediate dilution rate of 0.043 h−1
production rates increased by 25% and 34%, respectively. (F = 3.55 mL min−1 ).
Ko et al. [77] studied xylitol production employing waste wood
fermentation. The use of urea as a nitrogen source in the fermenta- 3.4. Xylitol purification from fermented hemicellulosic
tion medium was tested, replacing the yeast extract with 10 g L−1 hydrolysates
urea or 10 g L−1 soybean meal. The yield obtained using urea was
1.3-fold higher than that obtained with yeast extract. Zhang et al. In recent years, xylitol crystallization has attracted the attention
[10] evaluated ethanol and xylitol production from C. athensensis of scientists around the world. Recovery and purification are com-
SB18 from vegetable waste, supplementing the production medium plicated steps in many fermentation processes, since they usually
with 2.0 g L−1 urea. Significant variations in the parameters of depend on both the nature of the product and the complex com-
biomass growth and xylitol production were observed. The strain position of the fermentation broth. Xylitol, sorbitol and arabinitol
preferably used urea as a nitrogen source and showed high conver- are difficult to separate on an industrial scale and a specific sepa-
sion of xylose to xylitol, which was completely consumed in 108 h ration technique would be of great value. Regarding polyols, very
of fermentation. The highest cell density and xylitol production little information is available on xylitol recovery, with reports more
were 12.07 g L−1 and 112.27 g L−1 , respectively. The results indi- oriented to the treatment of hemicellulosic hydrolysates, fermen-
cated that, among the nitrogen sources investigated, urea is the tation and metabolic bioconversion [82].
most promising for the studied strain. Crystallization is a first-order phase transition. In other words, at
In contrast, Hongzhi et al. [78] conducted an experimental sta- the transition point the solid and liquid phases are in equilibrium;
tistical design to optimize a culture medium for xylitol production a supersaturated or undersaturated liquid phase can be created
from corn bagasse hydrolysate. The nitrogen sources used in the by a change in concentration, temperature or pressure; the two
experiments were ammonium nitrate, peptone, urea, ammonium phases are separated by a crystal surface with an interfacial tension
sulphate and yeast extract. Among these sources, only (NH4 )2 SO4 higher than zero; and supersaturation is needed to overcome the

Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010
G Model
PRBI-10195; No. of Pages 11 ARTICLE IN PRESS
T.L.d. Albuquerque et al. / Process Biochemistry xxx (2014) xxx–xxx 9

nucleation barrier caused by the interfacial tension [83]. Crystal- literature presents few researches related to the economic viability
lization is a separation process where a solid phase is created from of the biotechnological production of xylitol from lignocellulosic
a liquid phase. About 70% of the industrial products sold as bulk waste, therefore, there is a need for further research to show
chemicals, intermediates, fine chemicals, biochemicals and food data referents to this topic. The recovery of xylitol from lignocel-
additives are solids [83]. lulosic media fermented by microorganisms is poorly explored
The literature reveals few studies on recovering xylitol from in scientific papers, as well as the possible viable techniques for
fermented broths, and the establishment of a methodology that downstream processes. Theses processes are one of the most
allows recovery and makes its biotechnological production viable essential aspects for development of integrated technological
is desired. Many studies have attempted to purify xylitol using a solution for production of second generation biorefinary products
crystallization process. The prevalence of crystallization as a purifi- like xylitol via biotechnological process at an economic industrial
cation method is due to the feasibility of obtaining high-purity and scale. Finally, there are limited data in the literature about the
uniformly sized crystals, which facilitates the following steps in the parameters of scaling-up the xylitol production and modes of
process such as filtration, washing and drying of the product. reactors operation that can be used to optimize this process.
In order to improve the purity and yield of xylitol crystals
some studies have focused on the purification process, using ion-
exchange resins and optimization of crystallization conditions. Acknowledgments
Xylitol purification from fermented broths is difficult because of
the low product concentration and the complex composition of The authors are grateful for the financial support provided by
the media. Even though crystallization has been employed to the Brazilian research agencies CNPq and CAPES.
recover xylitol from fermentation media formulated with commer-
cial xylose, the complex composition of cultivated hydrolysates
References
does not allow a direct crystallization, requiring previous purifi-
cation steps such as adsorption on activated charcoal and ethanol [1] Bär A, Xylito L. In: O’Breen Nabors L, Gelardi RC, editors. Alternative sweeteners.
precipitation [84]. 2nd ed. New York: Marcel Deckker; 1991. p. 341.
Wei et al. [85], for example, evaluated the crystallization and [2] Mussatto SI, Roberto IC. Xilitol Edulcorante com efeitos benéficos para a saúde
humana. Braz J Pharm Sci 2002;38:401–13.
purification of xylitol from fermented corncob hydrolysate. The fer- [3] Khalid E, Johan S, Helén J, Jan S. Calorimetric and relaxation properties of
mented broth was initially bleached with activated carbon (solid xylitol–water mixtures. J Chem Phys 2012;136:104508.
percentage 1%, 60 ◦ C, 165 rpm), desalted with a combination of [4] Islam S. Effects of xylitol as a sugar substitute on diabetes-related parameters
in nondiabetic rats. J Med Food 2011;14:505–11.
two ion-exchange resins (732 and D301x), and residual sugars [5] Uittamo J, Nieminen MT, Kaihovaara P, Bowyer P, Salaspuro M, Rautemaa R.
were separated using resin UBK-555 (Ca2+ ). Then, the solution Xylitol inhibits carcinogenic acetaldehyde production by Candida species. Int J
was concentrated in vacuum to supersaturation (750 g L−1 xylitol). Cancer 2011;129:2038–41.
[6] Ritter AV, Bader JD, Leo MC, Preisser JS, Shugars DA, Vollmer WM, Amaechi BT,
Afterwards, 1% xylitol crystals were added to the supersaturated Holland JC. Tooth-surface-specific effects of xylitol: randomized trial results. J
solution, which was then cooled to −20 ◦ C for 48 h. Crystalline Dent Res 2013;92:512–7.
xylitol with a purity of 95% and a yield of 60.2% was obtained [7] Monedero V, Pérez-Martínez G, Yebra MJ. Perspectives of engineering lactic
acid bacteria for biotechnological polyol production. Appl Microbiol Biotechnol
by crystallization from the clarified fermentation broth. Mussatto
2010;86:1003–15.
et al. [86] studied xylitol purification from hemicellulosic sugar- [8] Chen X, Jiang Z-H, Chen S, Qin W. Microbial and bioconversion production of
cane bagasse hydrolysate fermented by C. guilliermondii FTI 20037. d-xylitol and its detection and application. Int J Biol Sci 2010;6:834–44.
[9] Miura M, Watanabe I, Shimotori Y, Aoyama M, Kojima Y, Kato Y. Microbial
Undesirable impurities were extracted from the broth using ethyl
conversion of bamboo hemicellulose hydrolysate to xylitol. Wood Sci Technol
acetate, chloroform or dichloromethane, with the best results for 2013;47:515–22.
clarification of the fermentation broth obtained with ethyl acetate, [10] Zhang J, Geng A, Yao C, Lu Y, Li Q. Xylitol production from d-xylose and horticul-
also considering xylitol loss. When ethanol, acetone or tetrahydro- tural waste hemicellulosic hydrolysate by a new isolate of Candida athensensis
SB18. Bioresour Technol 2012;105:134–41.
furan were used to precipitate impurities, only tetrahydrofuran was [11] Martínez. ML, Sáncheza S, Bravo V. Production of xylitol and ethanol by
effective in clarifying the fermentation broth but a large loss in Hansenula polymorpha from hydrolysates of sunflower stalks with phosphoric
xylitol content was noted (about 30%). acid. Ind Crop Prod 2012;40:160–6.
[12] Kelloway A, Daoutidis P. Process synthesis of biorefineries: optimiza-
Rivas et al. [84] evaluated the purification of xylitol obtained by tion of biomass conversion to fuels and chemicals. Ind Eng Chem Res
fermentation of corncob hydrolysates by a strain of Debaryomices 2014;53:5261–73.
hansenii. Under selected conditions, 43.7% of xylitol contained in [13] Nigam P, Singh D. Processes for fermentative production of xylitol – a sugar
substitute. Process Biochem 1995;30:117–24.
the initial fermentation broth was recovered in well-formed, homo- [14] Koutinas AA, Vlysidis A, Pleissner D, Kopsahelis N, Garcia IL, Kookos IK,
geneous crystals, in which xylitol accounted for 98.9% of the total Papanikolaou S, Kwan TH, Lin CSK. Valorization of industrial waste and
oven-dry weight. by-product streams via fermentation for the production of chemicals and
biopolymers. Chem Soc Rev 2014;43:2587–627.
[15] Franceschin G, Sudiro M, Ingram T, Smirnova I, Brunner G, Bertucco A. Conver-
sion of rye straw into fuel and xylitol: a technical and economical assessment
4. Future prospects and conclusions based on experimental data. Chem Eng Res Des 2011;89:631–40.
[16] Ruan Z, Zanotti M, Zhong Y, Liao W, Ducey C, Liu Y. Co-hydrolysis of lig-
nocellulosic biomass for microbial lipid accumulation. Biotechnol Bioeng
This article reviews the production of xylitol and the interest 2013;110:1039–49.
in xylitol has increased considerably in recent years, due to many [17] Zikou E, Chatzifragkou A, Koutinas AA, Papanikolaou S. Evaluating glucose and
commercial applications in different industrial sectors like food, xylose as cosubstrates for lipid accumulation and ␥-linolenic acid biosynthesis
of Thamnidium elegans. J Appl Microbiol 2013;114:1020–32.
dental related products, and pharmaceuticals. Therefore, over the [18] Fakas S, Papanikolaou S, Batsos A, Galiotou-Panayotoua M, Mallouchos A, Agge-
past few decades much effort has been devoted to the development lis G. Evaluating renewable carbon sources as substrates for single cell oil
of cost-effective and environmentally friendly biotechnological production by Cunninghamella echinulata and Mortierella isabellina. Biomass
Bioenerg 2009;33:573–80.
processes by evaluating lignocellulosic substrates. Numerous [19] Edelstein S, Smith K, Worthington A, Gillis N, Bruen D, Kang SH, Gilpin K, Acker-
types of hydrolysates of low-value raw materials have been used man J, Guiducci G. Comparisons of six new artificial sweetener gradation ratios
for biotechnological production. Additionally, a wide variety of with sucrose in conventional-method cupcakes resulting in best percentage
substitution ratios. J Cul Sci Technol 2008;5:61–74.
microorganisms is currently being assessed for their ability to pro- [20] Winkelhausena E, Malinovskaa RJ, Velickovaa E, Kuzmanovaa S. Sensory and
duce the polyol from these culture media, especially yeasts from microbiological quality of a baked product containing xylitol as an alternative
several genera, achieving significant productivity. The scientific sweetener. Int J Food Prop 2007;10:639–49.

Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010
G Model
PRBI-10195; No. of Pages 11 ARTICLE IN PRESS
10 T.L.d. Albuquerque et al. / Process Biochemistry xxx (2014) xxx–xxx

[21] Sokmen A, Gunes G. Influence of some bulk sweeteners on rheological proper- [52] Sene L, Vitolo M, Felipe MGA, Silva SS. Effect of environmental conditions on
ties of chocolate. LWT – Food Sci Technol Int 2006;39:1053–8. xylose reductase and xylitol dehydrogenase production in Candida guillier-
[22] Fisker HO, Nissen V. Effect of gum base and bulk sweetener on release of specific mondii. Appl Biochem Biotechnol 2000;84:371–80.
compounds from fruit flavoured chewing gum. Dev Food Sci 2006;46:429–32. [53] Yablochkova EN, Bolotnikova OI, Mikhailova NP. The activity of xylose reduc-
[23] Lee SH, Choi BK, Kim YJ. The cariogenic characters of xylitol-resistant and tase and xylitol dehydrogenase in yeasts. Microbiology 2003;72:414–7.
xylitol-sensitive Streptococcus mutans in biofilm formation with salivary bacte- [54] Misra S, Raghuwanshi S, Gupta P, Dutt K, Saxena RK. Fermentation behavior
ria. Arch Oral Biol 2011;57:697–703. of osmophilic yeast Candida tropicalis isolated from the nectar of Hibis-
[24] Tapiainen T, Sormunen R, Kaijalainen T, Kontiokari T, Ikäheimo I, Uhari M. Ultra- cus rosa sinensis flowers for xylitol production. Antonie Van Leeuwenhoek
structure of Streptococcus pneumoniae after exposure to xylitol. J Antimicrob 2012;101:393–402.
Chemother 2004;54:225–8. [55] Villarreal MLM, Prata AMR, Felipe MGA, Almeida e Silva JB. Detoxification
[25] Akiyama H, Ono T, Huh WK, Yamasaki O, Ogawa S, Katsuyama M, Ichikawa procedures of eucalyptus hemicellulose hydrolysate for xylitol production by
H, Iwatsukia K. Actions of farnesol and xylitol against Staphylococcus aureus. Candida guilliermondii. Enzyme Microb Technol 2006;40:17–24.
Chemotherapy 2002;48:122–8. [56] Chandel AK, Silva SS, Singh OV. Detoxification of lignocellulose hydrolysates:
[26] Ammons MCB, Ward LS, Dowd S, James G. Combined treatment of Pseudomonas biochemical and metabolic engineering toward white biotechnology. Bioen-
aeruginosa biofilm with lactoferrin and xylitol inhibits the ability of bacteria to ergy Res 2012;6:388–401.
respond to damage resulting from lactoferrin iron chelation. Int J Antimicrob [57] Parawira W, Tekere M. Biotechnological strategies to overcome inhibitors in
Agents 2011;37:316–23. lignocellulose hydrolysates for ethanol production: review. Crit Rev Biotechnol
[27] Imazato S, Ikebe K, Nokubi T, Ebisu S, Walls AW. Prevalence of root caries in a 2011;31:20–31.
selected population of older adults in Japan. J Oral Rehabil 2006;33:137–43. [58] Palmqvist E, Hahn-Hägerdal B. Fermentation of lignocellulosic hydrolysates.
[28] Wang R, Li L, Zhang B, Gao X, Wang D, Hong J. Improved xylose fermentation II: Inhibitors and mechanism of inhibition review. Bioresour Technol
of Kluyveromyces marxianus at elevated temperature through construction of a 2000;74:25–33.
xylose isomerase pathway. J Ind Microbiol Biotechnol 2013;40:841–54. [59] Kang L, Lee YY, Yoon SH, Smith AJ, Krishnagopalan GA. Ethanol production from
[29] Danhauer JL, Kelly A, Johnson CE. Is mother–child transmission a possible vehi- the mixture of hemicellulose prehydrolysate and paper sludge. BioResources
cle for xylitol prophylaxis in acute otitis media? Int J Audiol 2011;50:661–72. 2012;7:3607–26.
[30] Mattila PT, Kangasmaa H, Knuuttila MLE. The effect of a simultaneous [60] Mateo S, Roberto IC, Sánchez S, Moya AJ. Detoxification of hemicellulosic
dietary administration of xylitol and ethanol on bone resorption. Metabolism hydrolyzate from olive tree pruning residue. Ind Crop Prod 2013;49:196–203.
2005;54:548–51. [61] Misra S, Raghuwanshi S, Saxena RK. Evaluation of corncob hemicellulosic
[31] Ylikahri R. Metabolic and nutritional aspects of xylitol. Adv Food Res hydrolysate for xylitol production by adapted strain of Candida tropicalis. Car-
1979;25:159–80. bohydr Polym 2013;92:1596–601.
[32] Branco RF, Silva SS. A novel use for sugarcane bagasse hemicellulosic fraction: [62] Nonklang AS, Banat BMA, Cha-Aim K, MoonjaI. N, Hoshida H, Limtong S, Yamada
xylitol enzymatic production. Biomass Bioenerg 2011;35:241–3246. M, Akada R. High-temperature ethanol fermentation and transformation with
[33] Rafiqul ISM, Sakinah AMM. A perspective: bioproduction of xylitol by enzyme linear DNA in the thermotolerant yeast Kluyveromyces marxianus DMKU3-
technology and future prospects. Food Technol 2012;19:405–8. 1042. Appl Environ Microbiol 2008;74:7514–21.
[34] Jaffe GM, Szkrybalo W, Weinert PH. Process for producing xylose. US Patent no. [63] Rodrussamee N, Lertwattanasakul N, Hirata K, Suprayogi Limtong S, Kosaka T,
3,784,408; 1974. Yamada M. Growth and ethanol fermentation ability on hexose and pentose
[35] Melaja AJ, Hamäläinen L. Process for making Xylitol. US Patent no. 4,008,285; sugars and glucose effect under various conditions in thermotolerant yeast
18 June 1975, publ. 15 February 1977. Kluyveromyces marxianus. Appl Microbiol Biotechnol 2011;90:1573–86.
[36] Ping Y, Ling H-Z, Song G, Ge J-P. Xylitol production from non-detoxified [64] Wilkins MR, Mueller M, Eichling S, Banat IM. Fermentation of xylose by the
corncob hemicellulose acid hydrolysate by Candida tropicalis. Biochem Eng J thermotolerant yeast strains Kluyveromyces marxianus IMB2, IMB4, and IMB5
2013;75:86–91. under anaerobic conditions. Process Biochem 2008;43:346–50.
[37] Arruda PV, Rodrigues, RDCLB, Silva DDV, Felipe, MDGDA. Evaluation of hexose [65] Kumar S, Singh SP, Mishra IM, Adhikari DK. Ethanol and xylitol production
and pentose in pre-cultivation of Candida guilliermondii on the key enzymes from glucose and xylose at high temperature by Kluyveromyces sp. IIPE453. J
for xylitol production in sugarcane hemicellulosic hydrolysate. Biodegradation Ind Microbiol Biotechnol 2009;36:1483–9.
2011;22:815–22. [66] Srivani K, Setty YP. Parametric optimization of xylitol production from xylose
[38] Rocha MVP, Rodrigues THS, Albuquerque TL, Gonçalves LRB, Macedo GR. Eval- by fermentation. Asia-Pac J Chem Eng 2012;7:280–4.
uation of dilute acid pretreatment on cashew apple bagasse for ethanol and [67] Ramesh S, Muthuvelayudham R, Kannan R, Rajesh, Viruthagiri T. Response
xylitol production. Chem Eng J 2014;243:234–43. surface optimization of medium composition for xylitol production by Debary-
[39] Prakash G, Varma AJ, Prabhune A, Shouche Y, Rao M. Microbial produc- omyces hansenii var. hansenii using corncob hemicellulose hydrolysate. Chem
tion of xylitol from d-xylose and sugarcane bagasse hemicellulose using Ind Chem Eng Q 2013;19:377–84.
newly isolated thermotolerant yeast Debaryomyces hansenii. Bioresour Technol [68] El-Baz AF, Shetaia YM, Elkhouli RR. Xylitol production by Candida tropicalis
2011;102:3304–8. under different statistically optimized growth conditions. Afr J Biotechnol
[40] García FJ, Sánchez S, Bravo V, Cuevas M, Rigal L, Gaset A. Xylitol production 2011;10:15353–63.
from olive-pruning debris by sulphuric acid hydrolysis and fermentation with [69] Zhang B, Zhang L, Wang D, Gao X, Hong J. Identification of a xylose reductase
Candida tropicalis. Holzforschung 2011;65:59–65. gene in the xylose metabolic pathway of Kluyveromyces marxianus NBRC1777.
[41] Bura R, Vajzovic A, Doty SL. Novel endophytic yeast Rhodotorula mucilaginosa J Ind Microbiol Biotechnol 2011;38:2001–10.
strain PTD3 I: production of xylitol and ethanol. J Ind Microbiol Biotechnol [70] Sirisansaneeyakul S, Wannawilai S, Chisti Y. Repeated fed-batch produc-
2012;2012(39):1003–11. tion of xylitol by Candida magnolia TISTR 5663. J Chem Technol Biotechnol
[42] Fonseca GG, Carvalho NMB, Gombert AR. Growth of the yeast Kluyveromyces 2013;88:1121–9.
marxianus CBS 6556 on different sugar combinations as sole carbon and energy [71] Mussatto SI, Roberto IC. Xylitol production from high xylose concentration:
source. Appl Microbiol Biotechnol 2013;97:5055–67. evaluation of the fermentation in bioreactor under different stirring rates. J
[43] Li M, Meng X, Diao Du F. Xylitol production by Candida tropicalis from corn Appl Microbiol 2003;95:331–7.
cob hemicellulose hydrolysate in a two-stage fed-batch fermentation process. [72] Papanikolaou S, Aggelis G. Lipids of oleaginous yeasts. Part I: Biochemistry of
J Chem Technol Biotechnol 2011;87:387–92. single cell oil production. Eur J Lipid Sci Technol 2011;113:1031–51.
[44] Rangaswamy S, Agblevor FA. Screening of facultative anaerobic bacteria uti- [73] Kamat S, Gaikwad S, Kumar AR, Gade WN. Xylitol production by Cyberlind-
lizing d-xylose for xylitol production. Appl Microbiol Biotechnol 2002;60: nera (Williopsis) saturnus, a tropical mangrove yeast from xylose and corn cob
88–93. hydrolysate. J Appl Microbiol 2013;115:1357–67.
[45] Cirino PC, Chin JW, Ingram LO. Engineering Escherichia coli for xylitol production [74] Vajzovic A, Bura R, Kohlmeier K, Doty SL. Novel endophytic yeast Rhodotorula
from glucose–xylose mixtures. Biotechnol Bioeng 2006;95:1167–76. mucilaginosa strain PTD3 II: production of xylitol and ethanol in the presence
[46] Povelainen M, Miasnikov AN. Production of xylitol by metabolically engineered of inhibitors. J Ind Microbiol Biotechnol 2012;39:1453–63.
strains of Bacillus subtilis. J Biotechnol 2010;128:24–31. [75] Wang YM, Patterson JH, Van Eys J. The potential use of xylitol in glucose-6-
[47] Sasaki M, Jojima T, Inui M, Yukawa H. Xylitol production by recombi- phosphate dehydrogenase deficiency anemia. J Clin Invest 1971;50:1421–8.
nant Corynebacterium glutamicum under oxygen deprivation. Appl Microbiol [76] Rodrigues RCLB, Kenealy WR, Jeffries TW. Xylitol production from DEO
Biotechnol 2010;86:1057–66. hydrolysate of corn stover by Pichia stipitis YS-30. J Ind Microbiol Biotechnol
[48] Berghäll S, Hilditch S, Penttilä M, Richard P. Identification in the mould Hypocrea 2011;38:1649–55.
jecorina of a gene encoding an NADP(+): d-xylose dehydrogenase. FEMS Micro- [77] Ko CH, Chiang PN, Chiu PC, Liu CC, Yang CL, Shiau IL. Integrated xylitol pro-
biol Lett 2007;277:249–53. duction by fermentation of hardwood wastes. J Chem Technol Biotechnol
[49] Dashtban M, Kepka G, Seiboth B, Qin W. Xylitol production by genetically engi- 2008;83:534–40.
neered Trichoderma reesei strains using barley straw as feedstock. Appl Biochem [78] Hongzhi L, Keke C, Jingping G, Wenxiang P. Statistical optimization of xylitol
Biotechnol 2013;169:554–69. production from corncob hemicellulose hydrolysate by Candida tropicalis HDY-
[50] Iverson A, Garza E, Zhao J, Wang Y, Zhao X, Wang J, Manow R, Zhou S. Increasing 02. Nat Biotechnol 2011;28:673–8.
reducing power output (NADH) of glucose catabolism for reduction of xylose [79] Silva SS, Santos JC, Carvalho W, Aracava KK, Vitolo M. Use of a fluidized
to xylitol by genetically engineered Escherichia coli AI05. World J Microbiol bed reactor operated in semi-continuous mode for xylose-to-xylitol conver-
Biotechnol 2013;29:1225–32. sion by Candida guilliermondii immobilized on porous glass. Proc Biochem
[51] Pal S, Choudhary V, Kumar A, Biswas D, Mondal AK, Sahoo DK. Studies on 2003;38:903–7.
xylitol production by metabolic pathway engineered Debaryomyces hansenii. [80] Santos JC, Converti A, Carvalho W, Mussatto SI, Silva SS. Influence of aer-
Bioresour Technol 2013;147:449–55. ation rate and carrier concentration on xylitol production from sugarcane

Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010
G Model
PRBI-10195; No. of Pages 11 ARTICLE IN PRESS
T.L.d. Albuquerque et al. / Process Biochemistry xxx (2014) xxx–xxx 11

bagasse hydrolyzate in immobilized-cell fluidized bed reactor. Proc Biochem [87] Rocha MVP, Rodrigues THS, Melo VMM, Gonçalves LRB, Macedo GR.
2005;40:113–8. Cashew apple bagasse as a source of sugars for ethanol production by
[81] Salgado JM, Rodríguez N, Cortés S, Domínguez JM. Coupling two sizes Kluyveromyces marxianus CE025. J Ind Microbiol Biotechnol 2011;38:1099–
of CSTR-type bioreactors for sequential lactic acid and xylitol production 107.
from hemicellulosic hydrolysates of vineshoot trimmings. Nat Biotechnol [88] Carvalheiro F, Duarte LC, Medeiros R, Gírio FM. Xylitol production by Debary-
2012;29:421–7. omyces hansenii in brewery spent grain dilute-acid hydrolysate: effect of
[82] Misra S, Gupta P, Raghuwanshi S, Dutt K, Saxena RK. Comparative study on dif- supplementation. Biotechnol Lett 2007;29:1887–91.
ferent strategies involved for xylitol purification from culture media fermented [89] García-Diéguez C, Salgado JM, Roca E, Domínguez JM. Kinetic modelling of the
by Candida tropicalis. Sep Purif Technol 2011;78:266–73. sequential production of lactic acid and xylitol from vine trimming wastes.
[83] Martínez EA, Giulietti M, Almeida e Silva JB, De Derenzo S, Almeida Felipe MDG. Bioprocess Biosyst Eng 2001;34:869–78.
Batch cooling crystallization of xylitol produced by biotechnological route. J [90] Huang C-F, Jiang Y-F, Guo G-L, Hwang W-S. Development of a yeast strain for
Chem Technol Biotechnol 2009;84:376–81. xylitol production without hydrolysate detoxification as part of the integration
[84] Rivas B, Torre P, Domínguez JM, Converti A, Parajó JC. Purification of xyli- of co-product generation within the lignocellulosic ethanol process. Bioresour
tol obtained by fermentation of corncob hydrolysates. J Agric Food Chem Technol 2011;102:3322–9.
2006;54:4430–5. [91] Cheng KK, Zhang JA, Ling HZ, Ping WX, Huang W, Ge J-P, Xu J-M. Optimiza-
[85] Wei J, Yuan Q, Wang T, Wang L. Purification and crystallization of xylitol from tion of pH and acetic acid concentration for bioconversion of hemicellulose
fermentation broth of corncob hydrolysates. Front Chem Sci Eng 2010;4:57–64. from corncobs to xylitol by Candida tropicalis. Biochem Eng J 2009;43:
[86] Mussatto SI, Santos JC, Filho WCR, Silva SS. A study on the recovery of xyli- 203–7.
tol by batch adsorption and crystallization from fermented sugarcane bagasse
hydrolysate. J Chem Technol Biotechnol 2006;81:1840–5.

Please cite this article in press as: Albuquerque TLd, et al. Biotechnological production of xylitol from lignocellulosic wastes: A review.
Process Biochem (2014), http://dx.doi.org/10.1016/j.procbio.2014.07.010

You might also like