Microbial Enzymes in Bioconversions of Biomass
Microbial Enzymes in Bioconversions of Biomass
Microbial Enzymes in Bioconversions of Biomass
Vijai Kumar Gupta Editor
Microbial
Enzymes in
Bioconversions
of Biomass
Biofuel and Biorefinery Technologies
Volume 3
Series editors
Vijai Kumar Gupta, Molecular Glycobiotechnology Group, Department of
Biochemistry, School of Natural Sciences, National University of Ireland
Galway, Galway, Ireland
Maria G. Tuohy, Molecular Glycobiotechnology Group, Department of
Biochemistry, School of Natural Sciences, National University of Ireland
Galway, Galway, Ireland
More information about this series at http://www.springer.com/series/11833
Vijai Kumar Gupta
Editor
Microbial Enzymes
in Bioconversions
of Biomass
123
Editor
Vijai Kumar Gupta
Department of Biochemistry, School of
Natural Sciences
National University of Ireland Galway
Galway
Ireland
v
vi Preface
Part I Cellulases
2 Cellobiohydrolases: Role, Mechanism, and Recent
Developments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
Neelamegam Annamalai, Mayavan Veeramuthu Rajeswari
and Nallusamy Sivakumar
3 Endo-1,4-β-glucanases: Role, Applications and Recent
Developments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Neelamegam Annamalai, Mayavan Veeramuthu Rajeswari
and Thangavel Balasubramanian
4 The Role and Applications of β-Glucosidases in Biomass
Degradation and Bioconversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
Hanlin Ouyang and Feng Xu
Part II Hemicellulases
5 Role and Applications of Feruloyl Esterases in Biomass
Bioconversion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
Constantinos Katsimpouras, Io Antonopoulou, Paul Christakopoulos
and Evangelos Topakas
6 Endo-β-1,4-xylanase: An Overview of Recent Developments . . . . . . 125
Alexandre Gomes Rodrigues
7 Microbial Xylanases: Sources, Types, and Their Applications . . . . . 151
Hesham Ali El Enshasy, Subeesh Kunhi Kandiyil, Roslinda Malek
and Nor Zalina Othman
vii
viii Contents
8 Mannanase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
Suttipun Keawsompong
9 The Role and Applications of Xyloglucan Hydrolase in Biomass
Degradation/Bioconversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
M. Saritha, Anju Arora, Jairam Choudhary, Vijaya Rani,
Surender Singh, Anamika Sharma, Shalley Sharma and Lata Nain
M.P. Raghavendra
Postgraduate Department of Microbiology, Maharani’s Science College for Women,
JLB Road, Mysore, Karnataka 570 005, India
S.C. Nayaka (&)
Department of Studies in Biotechnology, University of Mysore, Mysore,
Karnataka 570 005, India
e-mail: [email protected]
V.K. Gupta
Department of Biochemistry, School of Natural Sciences,
National University of Ireland, Galway, Ireland
ganisms and its array of enzymes need to be effectively screened, identified and
employed in developing effective strategies for converting biomass to biofuel.
1.1 Introduction
Due to over reliance on petrochemicals, rising oil prices, pollution associated with
its usage, world is looking forward for alternative renewable resources of energy
from different sources. Biofuel obtained from carbon sources are gaining
momentum as a source of biofuel but the concern over production of ethanol from
food crops is creating a imbalance and there is a search for a resources, which are
not being used as food and should be available plenty in the ecosystem so that the
problems associated with the starchy material can be addressed. In this connection
cellulosic feed stock is considered to be excellent source for ethanol production, but
there are problems associated with degradation of cellulose also due to its recal-
citrance to biodegradation at least at the early stages of degradation. Liquid fuel
derived from biomass feedstock has been considered as a low-carbon substitute for
fossil hydrocarbons in vehicles, but many of the existing approaches are encoun-
tering significant challenges.
Search for new energy supplies is an important agenda for several countries in the
coming years. They are developing new energy infrastructure capable of meeting
growing demands for electric power and transportation fuels. Even unrest in oil
producing countries will have great impact on national security and economic
strength of several oil dependent countries. India in particular currently imports
80 % of its oils and it is expected to increase to 91 % by 2020. This overreliance on
oil import is the greatest burden on Indian economy and its security. On the other
hand, India is an agrarian country and known for its biodiversity that can be
exploited to convert biomass to biofuel. Biofuels employ recycling of agricultural
byproducts and dedicated energy crops, which offer opportunities for mitigation of
greenhouse gas emission as growing these leads to C sequestration through
photosynthesis.
Various plants and plant-derived materials are used for biofuels manufacturing
including grains (1st generation) and lignocellulosic biomass (2nd generation). The
second generation of biofuels is more important as they are based on the cheap and
abundant lignocellulosic biomass and do not compete with food crops. Wheat straw
and bagasse are considered that best example, which are inexpensive and widely
available ligonocellulosic resources containing 75–80 % polysaccharides (cellulose
and hemicelluloses). The monosaccharides glucose and xylose obtained from
hydrolysis of these can be used as substrate for production of industrially important
1 Microbial Enzymes for Conversion of Biomass to Bioenergy 3
Lignin, Hydrolysates of
Cellulose hemicellulose
Enzymatic/acid hydrolysis
Green gasoline or
diesel Fermentation
products (Fig. 1.1). These products will have more potential economical, envi-
ronmental, and strategic advantages over traditional fossil-based products.
The era of advanced biofuels—cellulosic ethanol, biomass-based diesel,
biobutanol, bio-oil, green gasoline, and bio-based jet fuel—is also drawing nearer
and nearer. According to the International Energy Agency, biofuels have the
potential to meet more than a quarter of world demand for transportation fuels by
2050. Bioethanol is by far the most widely used biofuel for transportation world-
wide (Mohanram et al. 2013). It has been estimated that widespread use of biofuels,
biopower, and other bio-based products has the potential to conserve 1.26 billion
barrels of oil, 58 million tons of coal, and 682 million tons of carbon dioxide from
2020 to 2030.
The cell walls of plants represent an enormous nutrient source, yet highly variable
in terms of amount, diversity, and botanical source. Plant carbohydrates are
chemically and structurally highly complex, and are arranged in a
three-dimensional network that has evolved to be intrinsically resistant to enzymatic
4 M.P. Raghavendra et al.
breakdown (Fontes and Gilbert 2010). Thus high molecular weight crystalline
cellulose microfibrils are intertwinned with hemicelluloses and pectins, which are a
whole range of homo- and heteropolysaccharides composed of dozens of different
monosaccharide units linked in a multitude of ways to resist degradation. Ester
substituents or non-carbohydrate, polymers, such as lignin, proteins, cutin, and
suberin, add a further layer of complexity. As a result, a single vegetable contains
hundreds of different bonds that need to be cleaved in order to unlock the assim-
ilable carbon of the cell wall constituents.
In general cell walls are intricate assemblages of celluloses, hemicelluloses (i.e.,
xyloglucans, arabinoxylans, and glucomannans), pectins (i.e., homogalacturonan,
rhamnogalacturonan I and II, and xylogalacturonans), lignins, and proteoglycans
(e.g., arabinogalactan-proteins, extensins, and proline-rich proteins). Most mass in
the plant cell wall is in the form of polysaccharides (cellulose and hemicelluloses).
The next most abundant polymer is lignin, which is composed predominantly of
phenylpropane building blocks.
Among these different carbon sources lignocellulosic biomass is the most
abundant, low-cost and unexploited biomass that is considered as a source for
industrial production of fuel ethanol. Considering its availability over the last
decades, researchers have been devoted to converting lignocellulosic materials to
bioethanol. Lignocellulosics are composed of heterogeneous complex of carbohy-
drate polymers (cellulose, hemicelluloses, and lignin). Majority of plant biomass is
locked up in 5- and 6-carbon sugars, comprised of mainly cellulose (a glucose
homopolymer, 40–60 % by wt.); less so, hemicelluloses (a sugar heteropolymer,
20–40 % by wt.); and least of all lignin (a complex aromatic polymer, 10–30 % by
wt). The major component cellulose, which is organized into microfibrils, each
containing up to 36 glucan chains having thousands of glucose residues linked by
β-(1,4) glycosidic bonds, is largely responsible for the plant cell wall’s mechanical
strength. Hemicelluloses are built up by pentoses (D-xylose, D-arabinose), hexoses
(D-mannose, D-glucose, D-galactose) and sugar acids. These include β-glucan,
xylan, xyloglucan, arabinoxylan, mannan, galactomannan, arabinan, etc. Hardwood
contains mainly xylans, while in softwood glucomannans are most common. Both
the cellulose and hemicelluloses can be broken down enzymatically into the
component sugars which may be then fermented to ethanol.
Cellulose can be classified into two types based on the arrangement of chains. In
type I cellulose chains are arranged in parallel direction of the long axis of the
microfibril, whereas in type II chains are in antiparrellel position. Among these,
type I is considered to be dominant that can be converted to type II by alkali
treatment. Studies show that cellulose I actually contain two distinct crystal lattices:
cellulose Iα, with triclinic symmetry, and cellulose Iβ, with monoclinic symmetry.
These two forms differ in the organization of their intermolecular hydrogen bonds
and lead to further complexity of cellulose structure. The percentage crystallinity of
cellulose varies from very low to almost 100 % and it depends on cellulose source.
However, it is clear that cellulose microfibrils are not uniformly crystalline;
imperfections in packing or mechanical damage result in a proportion of substrate
1 Microbial Enzymes for Conversion of Biomass to Bioenergy 5
evident that isolated cellulosomes could attain an overall tight or loose conforma-
tion (Mayer et al. 1987; Lamed et al. 1983a). The cellulolytic enzymes are either
secreted into the substrate or attached to the cell wall of the microorganism. The
former system is called noncomplexed and the latter is complexed cellulase system.
Noncomplexed cellulase systems are mostly found in filamentous fungi and acti-
nomycete bacteria, because they can penetrate the lignocellulosic material with their
hyphal extensions. The enzymes of noncomplexed cellulase systems are then just
released into the substrate and the free enzymes start hydrolyzing the cellulose. The
glucose and cellodextrins of a length of maximal four glucose molecules are taken
up by the microorganism and either used directly or cleaved further via intracellular
hydrolases. Organisms that produce noncomplexed cellulase systems are most often
used in the industrial production of cellulolytic enzymes, because the secreted
enzymes can easily be harvested (Lynd et al. 2002).
Cellulosome contains cellulose-binding domains, also called carbohydrate
binding modules (CBM), which bind to the cellulosic material (Fontes and Gilbert
2010; Sweeney and Xu 2012; Shoseyov et al. 2006). CBMs can also be found in
noncomplexed cellulase systems, where they are especially important for binding
and processing of exoglucanases. The main function of the cellulosome is to hold
the microorganism tight to the cellulose, and therefore reduce diffusion distances
and losses and to arrange the different cellulolytic enzymes in a favorable way.
More recently (Garcia-Alvarez et al. 2013), ultrastructural studies of a homoge-
neous mini-cellulosome containing three cohesin modules attached to three
matching cellulases suggested that the flexibility of the linkers connecting con-
secutive cohesin modules could control structural transitions and thus regulate
substrate recognition and degradation (Vazana et al. 2013).
One of the most potent cellulose degrading microorganisms is the well-studied
bacterium Clostridium thermocellum. This anaerobic thermophilic cellulolytic
bacterium secretes a multienzymatic complex called the cellulosome, first discov-
ered in 1983 (Lamed et al. 1983a, b). Later there are several reports on cellulosomal
oraganisms which possess an array of cellulosomal architectures (Belaich et al.
1997; Gal et al. 1997; Ding et al. 1999, 2000; Ohara et al. 2000; Rincon et al. 2003,
2004, 2010; Xu et al. 2003, 2004; Bayer et al. 2004, 2013; Dassa et al. 2012).
environmental and industrial niches, which produce cellulolytic strains that are
extremely resistant to environmental stresses. These include strains that are ther-
mophilic or psychrophilic, alkaliphilic, or acidophilic, and, strains that are halo-
philic. Not only can these strains survive the harsh conditions found in the
bioconversion process, but also they can produce enzymes that are stable under
extreme conditions which may be present in the bioconversion process and this may
increase rates of enzymatic hydrolysis, fermentation, and, product recovery (Maki
et al. 2009). Along with these due to availability of novel techniques for gene
transfer, microbes can be genetically modified easily to suit the need of the
industrial processes and also the different substrates employed in the fermentation.
Considering these, decomposition of lignocelluloses by enzymes produced by
complex microbial communities represents a promising alternative for biomass
conversion. In particular, thermophilic consortia are a potential source of enzymes
adapted to adverse reaction conditions (Wongwilaiwalina et al. 2010; Gladden
et al. 2011). Members of the genus Caldicellulosiruptor can grow on and degrade
biomass containing high lignin content as well as highly crystalline cellulose
without conventional pretreatment (Blumer-Schuette et al. 2008; Yang et al. 2009)
and are also known to ferment all primary C5 and C6 sugars from plant biomass
and are the most thermophilic cellulolytic bacteria known, with growth temperature
optima between 78 and 80 °C raising the possibility of further economic
improvement of biofuel production from plant biomass by reducing or eliminating
the pretreatment step (Hamilton-Brehm et al. 2010).
In comparison with the anaerobic thermophilic bacteria few aerobic have been
described to produce cellulases and xylanses. The aerobic thermophiles R. marinus
produces a highly thermostable cellulose (Cel 2A), and three glycoside hydrolases
belonging to family GH10 of the carbohydrate active enzyme (CAZy) database
(http://www.cazy.org) (Cantarel et al. 2009). Aquifex aeolicus, isolated in the
Aeolic Islands in Sicily (Italy), represents one of the most thermophilic bacteria
since its growth temperature can reach 95 °C (Deckert et al. 1998).
Bacteria belonging to Clostridium, Ruminococcus, Bacteroides, Erwinia,
Acetovibrio, Microbispora, and Streptomyces can produce cellulases (Bisaria
1991). Cellulomonas fimi and Thermomonospora fusca have been extensively
studied for cellulase production. Similarly, other bacterial strains have the ability to
produce cellulase complexes aerobically as well as anaerobically. Cellulase pro-
ducing bacterial strains of Rhodospirillum rubrum, Cellulomonas fimi, Clostridium
stercorarium, Bacillus polymyxa, Pyrococcus furiosus, Acidothermus cellulolyticus,
and Saccharophagus degradans have been extensively reviewed (Kumar et al.
2008; Weber et al. 2001; Kato et al. 2005; Taylor et al. 2006). An extracellular
alkaline carboxy methyl cellulase (CMCase) from Bacillus subtilis strain AS3 has
been purified and characterized by Deka et al. (2013) for utilization of cellulosic
biomass. Although many cellulolytic bacteria, particularly the cellulolytic anaer-
obes such as Clostridium thermocellum and Bacteroides cellulosolvens produce
1 Microbial Enzymes for Conversion of Biomass to Bioenergy 9
cellulases with high specific activity, they do not produce high enzyme titres (Lynd
et al. 2002).
Apart from the cellulolytic fungus T. reesei, many other fungi produce cellulases
and degrade treated cellulosic material or soluble cellulose derivatives such as
carboxymethylcellulose. However, they are not very effective on crystalline cel-
lulosic substrates. Mesophilic strains producing cellulases like Fusarium oxyspo-
rum, Piptoporus betulinus, Penicillium echinulatum, P. purpurogenum, Aspergillus
niger and A. fumigatus have also been reported (Martins et al. 2008; Valaskova and
Baldrian 2006). The cellulases from Aspergillus usually have high β-glucosidase
activity but lower endoglucanase levels, whereas, Trichoderma has high endo and
exoglucanase components with lower β-glucosidase levels, and hence has limited
efficiency in cellulose hydrolysis. Thermophillic microorganisms such as
Sporotrichum thermophile, Scytalidium thermophilum, Clostridium straminisolvens
and Thermomonospora curvata also produce the cellulase complex and can degrade
native cellulose (Kato et al. 2005; Kaur et al. 2004).
Even human gut microbiota encodes a huge diversity of enzymes for the
digestion of all components of plant cell wall polysaccharides including cellulose.
Turnbaugh et al. (2010) have shown that the distal gut microbiota of humans also
encodes dockerin-containing cellulolytic enzymes that indicate the presence of
cellulosomes (Fontes and Gilbert 2010; Bayer et al. 2004). The presence of
dockerin-containing proteins is also observed in the oral, nasal and vaginal samples
(Cantarel et al. 2012). In nondigestive sites, however, these dockerin domains may
have another role than attaching cellulases to form a cellulosome.
The interaction between loads of bacteria in gut with other bacteria has lead to
development of novel CAZymes. The recent outcome of the research by Hehemann
et al. (2010) proved the novel CAZymes added due to flexible microbial interac-
tions. They characterized the first porphyranases from a member of the marine
Bacteroidetes, Zobellia galactanivorans, active on the sulphated polysaccharide
porphyran from marine red algae of the genus Porphyra. Furthermore, they showed
that genes coding for these porphyranases, agarases and associated proteins have
been transferred to the gut bacterium Bacteroides plebeius isolated from Japanese
individuals. Their comparative gut metagenome analyses showed that por-
phyranases and agarases are frequent in the Japanese population and that they are
absent in metagenome data from North American individuals.
White Rot Fungi (WRF) are microorganisms of great interest that secrete
complex suites of nonspecific extracellular ligninolytic enzymes, i.e., lignin per-
oxidase (EC 1.11.1.14), manganese peroxidase (EC 1.11.1.13), and laccase (EC
1.10.3.2) to biodegrade lignin or digest substrates required for their proliferation
and growth. β-glucosidase is among the suite of enzymes produced by WRF such as
Pleurotus ostreatus, Auricularia auricula, Polyporus squamosus, Trametes versi-
color, Lentinula edodes, and Grifola frondosa to biodegrade plant biomass. The
β-glucosidase family (EC 3.2.1.21) is a wide-spread group of enzymes that catalyze
10 M.P. Raghavendra et al.
the hydrolysis of a broad variety of glycosides (Berrin et al. 2003). While some
organisms secrete either endoglucanase or β-glucosidase, in other organisms;
β-glucosidase is either lacking or produced in insufficient quantities (Kumar et al.
2008). When β-glucosidase secretion is low, cellobiose accumulates instead of
glucose. Cellobiose accumulation acts as a feedback-inhibitor of cellulose
depolymerization by endo- and exoglucanases (Howell and Stuck 1975; Morais
et al. 2010), which is a critical factor in the industrial scale conversion of cellulose
to glucose (Cai et al. 1997). This situation can be alleviated during industrial scale
conversion of cellulosic biomass by exogenous incorporation of β-glucosidase
enzyme.
Among known cellulolytic microorganisms, cellulases of Trichoderma reesei or
T. viride are well studied and characterized. Cellulase of T. reesei comprises of two
major cellobiohydrolases, (CBHI and CBHII), two major endoglucanases, (EGI and
EGII) and at least two low molecular weight endoglucanases, EGIII, and EGV. The
mixture of these enzymes is capable of solubilizing native cellulose. Many other
aerobic filamentous fungi, including Agaricus bisporus, Humicola spp., Irpex
lacteus, and Sclerotium roysii, also use similar cellulase systems.
Table 1.1 Novel enzymes and other proteins involved in degradation of biomass
Enzymes Function Microorganisms
Endo-1,4 β-D- Hydrolysis of the internal Acidothermus cellulolyticus,
glucanglucanohydrolase glycosidic linkages in a Pectobacterium
(EC. 3. 2. 1.4) random fashion, generating chrysanthami, Thermobifida
oligosaccharides of varying fusca, Thermotoga maritime,
lengths Bacillus, Pseudomonas,
Ruminococcus, Fibrobacter,
Clostrdium, Halomonas,
Streptomyces, Cellulomonas,
Mycobacterium, Aspergillus,
Trichoderma, Anaeromyces,
Pestalotiopsis,
Phanerochaete, Fusarium,
Orpinomyces, Piromyce
Exoglucanase or 1,4-β-D- Hydrolysis of beta-D- Bacillus, Pseudomonas,
glucan cellobiohydrolase glucosidic linkages by Clostridium, Paenibacillus,
(EC3.2.1.91) releasing mainly cellobiose Thermobifida, Cellulomonas,
either from the reducing or Mycobacterium, Ralstonia,
nonreducing ends of the Trichoderma, Penicillium,
chains Aspergillus, Chaetomium,
Fusarium, Pestalotiopsis,
Orpinomyces, Piromyces,
Rhizopus
β-glucosidases or β-D- Hydrolysis of terminal, Clostridium, Cellulomonas,
glucoside glucohydrolase nonreducing β-D-glucosyl Aerobacter, Leuconostoc,
(EC 3.2.1.21) residues with release of β-D- Aspergillus, Monilia,
glucose Phanerochaete, Sclerotium,
Saccharomyces,
Kluyveromyces
Cellodextrin phosphorylase Catalysis of the reversible Clostridium, Cellvibrio
or (1 → 4)-β-D-glucan: phosphorolytic cleavage of
phosphate α-D- cellodextrins ranging from
glucosyltransferase (EC cellotriose to cellohexoses
2.4.1.49)
Cellobiose phosphorylase Catalysis of the reversible Cellulomonas, Clostridium,
or cellobiose: phosphate phosphorolytic cleavage of Ruminococcus, Thermotoga,
α-D-glucosyltransferase cellobiose Cellvibrio, Fomes annosus
(EC 2.4.1.20)
Endo-1,4-β-D-xylanase or Endohydrolysis of (1 → 4)- Bacillus circulans,
1,4-beta-xylan beta-D-xylosidic linkages in Thermoanaerobacterium,
xylanohydrolase xylans to release xylose Ruminococcus, Geobacillus,
(EC 3.2.1.8) Thermopolyspora,
Cellulomonas, Streptomyces,
Aspergillus, Trichoderma,
Thermomyces, Fusarium,
Anaeromyces,
Neocallimastix,
Dictyoglomus
Thermophilum, Streptomyces
halstedii
(continued)
1 Microbial Enzymes for Conversion of Biomass to Bioenergy 13
Enzyme stabilization and lifetime improvement has been achieved through point
substitution and chemical modification, of single catalytic domains, either through
rational, structure-based modeling, or through directed evolution. Another approach
to increase enzyme stability and enhance activity is insertion of protein domains
into stable scaffolds.
An effort was made to stabilize the cellulases by embedding cellulases of
extremophiles into hyperstable α-helical consensus. It was observed that catalytic
domains CelA (CA, GH8; Clostridium thermocellum) and Cel12A (C12A, GH12;
Thermotoga maritima) to be stable in the context of ankyrin scaffold, and to be
active against both soluble and insoluble substrates. The ankyrin repeats in each
fusion are folded, although it appears that for C12A catalytic domain (CD) (where
the N- and C-termini are distant in the crystal structure), the two flanking ankyrin
domains are independent, whereas for CA (where termini are close), the flanking
ankyrin domains stabilize each other. Although the activity of CA is unchanged in
the context of the ankyrin scaffold, the activity of C12A is increased between two
and sixfold (for RAC and CMC substrates) at high temperatures. For C12A, activity
increases by with the number of flanking ankyrin repeats. These results show
ankyrin arrays to be a promising scaffold for constructing designer cellulosomes,
preserving or enhancing enzymatic activity and retaining thermostability.
In this type of scaffolding strategy, the distance between the N and C-termini of
the inserted protein (cellulase catalytic domains) is an important parameter.
Cellulases catalytic domains belong to various Glycoside Hydrolase (GH) families,
as reported in the CAZY database, with different folds, and thus have variable
N-to-C distances (Cantarel et al. 2009).
Insertion into a scaffold protein requires appropriate selection of insertion sites to
avoid disruption of structure of the inserted protein or scaffold (Betton et al. 1997;
Cutler et al. 2009). Proteins with linear, repeated architectures, such as ankyrin
repeats, have structurally modular architectures that should be able to accommodate
insertions locally, without affecting more distant regions of the scaffold. Ankyrin
repeat proteins are roughly linear arrays of 33 residues repeating units (Sedgwick
18 M.P. Raghavendra et al.
and Smerdon 1999; Zweifel and Barrick 2001). Each unit consists of two short
helices connected by a short, conserved turn, and connects to neighboring repeats
through an extended beta-hairpin containing loop (Zweifel et al. 2003). Adjacent
repeats strongly stabilize each other through highly stabilizing interfaces. The
repeating unit spans about 11 Å, approximately the size of each cellobiose unit
(2-units of glucose 10.4 Å). Studies have found that ankyrin repeats are greatly
stabilized by consensus sequence substitutions (Wetzel et al. 2008; Aksel et al.
2011). For a five repeat consensus ankyrin protein, the thermal denaturation mid-
point (Tm) was around 90 °C (Aksel et al. 2011; Aksel and Barrick 2009).
Therefore, consensus ankyrin seems to be an ideal scaffold to study the effects of
inserting cellulase CDs with various N–C distances.
Catalytic activity of an enzyme is defined and mediated by its 3D structure and its
mobility. It is impossible to obtain complete comprehension of the enzymatic action
without detailed molecular studies which include structural, biophysical, bio-
chemical and genetic investigations, enzymatic assays, bioinformatics, and
molecular dynamics computations. It is not possible to innovate and to consistently
produce competitive enzymatic cocktails/mixtures for biomass transformation
without profound understanding of function and activity of glycoside hydrolases.
Therefore, we need to integrate our knowledge of biodiversity, genomics and
genetics studies, structural and molecular biology, biochemistry, enzymology, and
bioinformatics, focusing on bioindustrial applications.
Enzyme cocktails that hydrolyze plant cell wall polysaccharides are a critical
component of bioprocessing configurations designed to transform lignocellulosic
biomass into biofuels (Lynd et al. 2002; Gao et al. 2011). Specific characteristics of
an enzyme (activity, thermostability, etc.) is often based on homology to known
enzymes (Schnoes et al. 2009). Although free cellulases and cellulosomes employ
different physical mechanisms to break down recalcitrant polysaccharides, when
these system combined display dramatic synergistic enzyme activity on cellulose.
Two natural enzyme systems—one produced by fungi and the other by bacteria—
break down cellulose faster if used in combination. The resulting process shows
promise for less expensive biofuels (Resch et al. 2013).
Current commercial cocktails consist of preparations of fungus derived glycoside
hydrolases, primarily cellulases and hemicellulases (Bouws et al. 2008; King et al.
2009). However, fungal enzymes are often deactivated by elevated temperatures or
by residual chemicals from pretreatment (Bouws et al. 2008). For example, ionic
liquids (ILs), such as 1-ethyl- 3-methylimidazolium acetate ([C2mim][OAc]), can
dissolve lignocellulosic biomass and dramatically improve cellulose hydrolysis
kinetics, yet multiple studies have shown that fungal endoglucanases are deacti-
vated at low levels of ILs that may persist in the biomass after pretreatment (Dadi
et al. 2006; Turner et al. 2003). In contrast, thermophilic bacterial and archaeal
1 Microbial Enzymes for Conversion of Biomass to Bioenergy 19
The natural diversity of enzymes could provide a large reservoir that can be further
improved by engineering enzymes and strains for increased performance (King
et al. 2009). A number of designer enzymes called glycosynthases, including cel-
lulases and hemicellulases, have been engineered by replacing nucleophilic residues
resulting in higher yields of different oligosaccharides (Kumar et al. 2008).
Feruloyl esterases (Faes) represent a subclass of carboxyl esterases that can release
phenolic acids, such as ferulic acids or other cinnamic acids, from esterified
polysaccharides, especially xylan and pectin. Feruloyl esterases (Faes) constitute a
subclass of carboxyl esterases that specifically hydrolyze the ester linkages between
ferulate and polysaccharides in plant cell walls. Recently, a comprehensive set of
enzymes essential for decomposing plant cell walls, including feruloyl esterase
activity, in a newly described in yak ruminal bacterium, Cellulosilyticum
ruminicola H1. Strain H1 grew robustly on natural plant fibers, such as corn cob,
alfalfa, and ryegrass, as the sole carbon and energy sources, as well as on a variety
of polysaccharides, including cellulose, xylan, mannan, and pectin (Li et al. 2011).
Therefore, Faes are regarded as the key enzymes to loosen the internal crosslink of
plant cell walls by acting as the important accessory enzymes in synergy with
(hemi)cellulases in plant cell wall hydrolysis.
Urgent need of the alternate energy resources increased the novel ideas to achieve
success to convert lignocelluloses and other carbohydrates to ethanol. In this
connection development of systems biology approaches and integrated, predictive
20 M.P. Raghavendra et al.
References
Dadi AP, Varanasi S, Schall CA (2006) Enhancement of cellulose saccharification kinetics using
an ionic liquid pretreatment step. Biotech. Bioeng. 95:904–910
Dassa B, Borovok I, Lamed R, Henrissat B, Coutinho P, Hemme L, Huang Y, Zhou J, Bayer EA
(2012) Genome-wide analysis of Acetivibrio cellulolyticus provides a blueprint of an elaborate
cellulosome system. BMC Genom 13:210–223
Datta S, Holmes B, Park JI, Chen Z, Dibble DC, Hadi M, Blanch HW, Simmons BA, Sapra R
(2010) Ionic liquid tolerant hyperthermophilic cellulases for biomass pretreatment and
hydrolysis. Green Chem 12:338–345
Deckert G, Warren PV, Gaasterland T, Young WG, Lenox AL, Graham DE, Overbeek R,
Snead MA, Keller M, Aujay M, Huber R, Feldman RA, Short JM, Olsen GJ, Swanson RV
(1998) The complete genome of the hyperthermophilic bacterium Aquifex aeolicus. Nature 392
(6674):353–358
Deka D, Jawed M, Goyal A (2013) Purification and characterization of an alkaline cellulase
produced by Bacillus subtilis (AS3). Prep Biochem Biotechnol 43:256–270
Den HR, Rose SH, Lynd LR, van Zyl WH (2007) Hydrolysis and fermentation of amorphous
cellulose by recombinant Saccharomyces cerevisiae. Metab Eng 9(1):87–94
Ding SY, Bayer EA, Steiner D, Shoham Y, Lamed R (1999) A novel cellulosomal scaffoldin from
Acetivibrio cellulolyticus that contains a family 9 glycosyl hydrolase. J Bacteriol
181:6720–6729
Ding SY, Bayer EA, Steiner D, Shoham Y, Lamed R (2000) A scaffoldin of the Bacteroides
cellulosolvens cellulosome that contains 11 type II cohesins. J Bacteriol 182:4915–4925
Eriksson T, Borjesson J, Tjerneld F (2002) Mechanism of surfactant effect in enzymatic hydrolysis
of lignocelluloses. Enz Microb Technol 31:353–364
Fontes CM, Gilbert HJ (2010) Cellulosomes: highly efficient nanomachines designed to
deconstruct plant cell wall complex carbohydrates. Annu Rev Biochem 79:655–681
Forsberg Z, Vaaje-Kolstad G, Westereng B, Bunæs AC, Stenstrøm Y, MacKenzie A, Sørlie M,
Horn SJ, Eijsink VGH (2011) Cleavage of cellulose by a CBM33 protein. Prot Sci
20:1479–1483
Fujita Y, Takahashi S, Ueda M, Tanaka A, Okada H, Morikawa Y, Kawaguchi T, Arai M,
Fukuda H, Kondo A (2002) Direct and efficient production of ethanol from cellulosic material
with a yeast strain displaying cellulolytic enzymes. Appl Environ Microbiol 68:5136–5141
Fujita Y, Iro J, Ueda M, Fukuda H, Kondo A (2004) Synergistic saccharification and direct
fermentation to ethanol, of amorphous cellulose by use of an engineered yeast strain
codisplaying three types of cellulolytic enzyme. Appl Environ Microbiol 70(2):1207–1212
Gal L, Pages S, Gaudin C, Belaich A, Reverbel-Leroy C, Tardif C, Belaich JP (1997)
Characterization of the cellulolytic complex (cellulosome) produced by Clostridium cellu-
lolyticum. Appl Environ Microbiol 63:903–909
Gao D, Uppugundla N, Shishir P, Chundawat S, Yu X, Hermanson S, Gowda K, Brumm P,
Mead D, Balan V, Dale BE (2011) Hemicellulases and auxiliary enzymes for improved
conversion of lignocellulosic biomass to monosaccharides. Biotechnol Biofuels 4:5–16
Garcia-Alvarez B, Melero R, Dias FM, Prates JA, Fontes CM, Smith SP, Romao MJ, Vazana Y,
Barak Y, Unger T, Peleg Y, Shamshoum M, Ben-Yehezkel T, Mazor Y, Shapiro E, Lamed R,
Bayer EA (2013) A synthetic biology approach for evaluating the functional contribution of
designer cellulosome components to deconstruction of cellulosic substrates. Biotechnol
Biofuels 6:182–200
Gladden JM, Allgaier M, Miller CS, Hazen TC, Vander Gheynst JS, Hugenholtz P, Simmons BA,
Singer SW (2011) Glycoside hydrolase activities of thermophilic bacterial consortia adapted to
switchgrass. Appl Environ Microbiol 77:5804–5812
Hamilton-Brehm SD, Mosher JJ, Vishnivetskaya T, Podar M, Carroll S, Allman S, Phelps TJ,
Keller M, Elkins JG (2010) Caldicellulosiruptor obsidiansis sp. nov., an anaerobic, extremely
thermophilic, cellulolytic bacterium isolated from Obsidian Pool, Yellowstone National Park.
Appl Environ Microbiol 76:1014–1020
1 Microbial Enzymes for Conversion of Biomass to Bioenergy 23
Mansfield MC, Touhy M, Saddler J (1998) The effect of the initial pore volume and lignin content
on the enzymatic hydrolysis of softwoods. Biores Technol 64:113–119
Martins LF, Kolling D, Camassola M, Dillon AJ, Ramos LP (2008) Comparison of Penicillium
echinulatum and Trichoderma reesei cellulases in relation to their activity against various
cellulosic substrates. Bioresour Technol 99:1417–1424
Mayer F, Coughlan MP, Mori Y, Ljungdahl LG (1987) Macromolecular organization of the
cellulolytic enzyme complex of Clostridium thermocellum as revealed by electron microscopy.
Appl Environ Microbiol 53:2785–2792
McMillan JD, Jenning EW, Mohagheghi A, Zuccarello M (2011) Comparative performance of
precommercial cellulases hydrolyzing pretreated corn stover. Biotech Biofuels 4:29–46
McQueen-Mason S, Cosgrove DJ (1994) Disruption of hydrogen bonding between plant cellwall
polymers by proteins that induce wall extension. Proc Natl Acad Sci USA 91:6574–6578
Mohanram S, Amat D, Choudhary J, Arora A, Nain L (2013) Novel perspectives for evolving
enzyme cocktails for lignocellulose hydrolysis in biorefineries. Sustain Chem Process 1:15–27
Morais S, Barak Y, Caspi J, Hadar Y, Lamed R, Shoham Y, Wilson DB, Bayer EA (2010)
Cellulase-xylanase synergy in designer cellulosomes for enhanced degradation of a complex
cellulosic substrate. mBio 1:00285–00210
Moser F, Irwin D, Chen SL, Wilson DB (2008) Regulation and characterization of Thermobifida
fusca carbohydrate-binding module proteins E7 and E8. Biotechnol Bioeng 100:1066–1077
Nair NU, Denard CA, Zhao H (2010) Engineering of enzymes for selective catalysis. Curr Org
Chem 14:1870–1882
Nakatani Y, Yamada K, Ogino C, Kondo A (2013) Synergetic effect of yeast cell-surface
expression of cellulase and expansin-like protein on direct ethanol production from cellulose.
Microb Cell Fact 12:66–73
Ohara H, Karita S, Kimura T, Sakka K, Ohmiya K (2000) Characterization of the cellulolytic
complex (cellulosome) from Ruminococcus albus. Biosci Biotechnol Biochem 64:254–260
Phillips CM, Beeson WT, Cate JH, Marletta MA (2011) Cellobiose dehydrogenase and a
copper-dependent polysaccharide monooxygenase potentiate cellulose degradation by
Neurospora crassa. ACS Chem Biol 6:1399–1406
Quinlan RJ, Sweeney MD, Lo Leggio L, Otten H, Poulsen JCN, Johansen KS, Krogh K,
Jorgensen CI, Tovborg M, Anthonsen A, Tryfona CP, Dupree WP, Xu F, Davies GJ,
Walton PH (2011) Insights into the oxidative degradation of cellulose by a copper
metalloenzyme that exploits biomass components. Proc Natl Acad Sci USA 108:15079–15084
Ranjan KK (2014) Adsorption, diffusion and activity of polycatalytic cellulase-nanoparticle
conjugates. Ph.D Dissertations. University of Connecticut
Resch MG, Donohoe BS, Baker JO, Decker SR, Bayer EA, Beckham GT, Himmel ME (2013)
Fungal cellulases and complexed cellulosomal enzymes exhibit synergistic mechanisms in
cellulose deconstruction. Energy Environ Sci 6:1858–1867
Rincon MT, Ding SY, McCrae SI, Martin JC, Aurilia V, Lamed R, Shoham Y, Bayer EA, Flint HJ
(2003) Novel organization and divergent dockerin specificities in the cellulosome system of
Ruminococcus flavefaciens. J Bacteriol 185:703–713
Rincon MT, Martin JC, Aurilia V, McCrae SI, Rucklidge GJ, Reid MD, Bayer EA, Lamed R,
Flint HJ (2004) ScaC, an adaptor protein carrying a novel cohesin that expands the
dockerin-binding repertoire of the Ruminococcus flavefaciens 17 cellulosome. J Bacteriol
186:2576–2585
Rincon MT, Dassa B, Flint HJ, Travis AJ, Jindou S, Borovok I, Lamed R, Bayer EA, Henrissat B,
Coutinho PM, Antonopoulos DA, Berg Miller ME, White BA (2010) Abundance and diversity
of dockerin-containing proteins in the fiberdegrading rumen bacterium, Ruminococcus
flavefaciens FD-1. PLoS ONE 5:e12476
Saloheimo M, Paloheimo M, Hakola S, Pere J, Swanson B, Nyyssönen E, Bhatia A, Ward M,
Penttilä M (2002) Swollenin, a Trichoderma reesei protein with sequence similarity to the plant
expansins, exhibits disruption activity on cellulosic materials. Eur J Biochem 269:4202–4211
Schnoes AM, Brown SD, Dodevski I, Babbitt PC (2009) Annotation error in public databases:
misannotation of molecular function in enzyme superfamilies. PLoS Comput Biol 5:e1000605
1 Microbial Enzymes for Conversion of Biomass to Bioenergy 25
Scott BR, Hill C, Tomashek J, Liu C (2009) Enzymatic hydrolysis of lignocellulosic feedstocks
using accessory enzymes. United States Patent Application 2009/0061484, 5 Mar 2009
Sedgwick SG, Smerdon SJ (1999) The ankyrin repeat: a diversity of interactions on a common
structural framework. Trends Biochem Sci 24:311–316
Shoseyov O, Shani Z, Levy I (2006) Carbohydrate binding modules: biochemical properties and
novel applications. Microbiol Mol Biol Rev 70:283–295
Sun Y, Cheng J (2002) Hydrolysis of lignocellulosic materials for ethanol production. Bioresour
Technol 83:1–11
Sweeney MD, Xu F (2012) Biomass converting enzymes as industrial biocatalysts for fuels and
chemicals: recent developments. Catalysts 2:244–263
Taherzadeh MJ, Karimi K (2007) Enzyme-based hydrolysis processes for ethanol from
lignocellulosic materials: a review. BioResources 2(4):707–738
Taylor LE, Henrissat B, Coutinho PM, Ekborg NA, Hutcheson SW, Weiner RM (2006) Complete
cellulase system in the marine bacterium Saccharophagus degradans strain 2-40T. J Bacteriol
188:3849–3861
Turnbaugh PJ, Quince C, Faith JJ, McHardy AC, Yatsunenko T, Niazi F, Affourtit J, Egholm M,
Henrissat B, Knight R, Gordon JI (2010) Organismal genetic and transcriptional variation in
the deeply sequenced gut microbiomes of identical twins. Proc Natl Acad Sci USA
107:7503–7508
Turner MB, Spear SK, Huddleston JG, Holbrey JD, Rogers RD (2003) Ionic liquid salt-induced
inactivation and unfolding of cellulose from Trichoderma reesei. Green Chem 5:443–447
Vaaje-Kolstad G, Horn SJ, van Aalten DMF, Synstad B, Eijsinkm VGH (2005) The noncatalytic
chitin-binding protein CBP21 from Serratia marcescens is essential for chitin degradation.
J Biol Chem 280:28492–28497
Valaskova V, Baldrian P (2006) Degradation of cellulose and hemicelluloses by the brown rot
fungus Piptoporus betulinus production of extracellular enzymes and characterization of the
major cellulases. Microbiol 152:3613–3619
Van RP, Van ZWH (1998) Pretorius IS: engineering yeast for efficient cellulose degradation. Yeast
14:67–76
Vazana Y, Barak Y, Unger T, Peleg Y, Shamshoum M, Ben-Yehezkel T, Mazor Y, Shapiro E,
Lamed R, Bayer EA (2013) A synthetic biology approach for evaluating the functional
contribution of designer cellulosome components to deconstruction of cellulosic substrates.
Biotechnol Biofuels 6:182–200
Wang Y, Tang R, Tao J, Gao G, Wang X, Mu Y, Feng Y (2011) Quantitative investigation of
non-hydrolytic disruptive activity on crystalline cellulose and application to recombinant
swollenin. Appl Microbiol Biotechnol 91:353–1363
Weber S, Stubner S, Conrad R (2001) Bacterial populations colonizing and degrading rice straw in
anoxic paddy soil. Appl Environ Microbiol 67:1318–1327
Wen F, Sun J, Zhao H (2010) Yeast surface display of trifunctional minicellulosomes for
simultaneous saccharification and fermentation of cellulose to ethanol. Appl Environ Microbiol
76:1251–1260
Westereng B, Ishida T, Vaaje-Kolstad G, Wu M, Eijsink VGH, Igarashi K, Samejima M,
Ståhlberg J, Horn SJ, Sandgren M (2011) The putative endoglucanase PcGH61D from
Phanerochaete chrysosporium is a metal dependent oxidative enzyme that cleaves cellulose.
PLoS ONE 6(11):e27807
Wetzel SK, Settanni G, Kenig M, Binz HK, Pluckthun A (2008) Folding and unfolding
mechanism of highly stable full-consensus ankyrin repeat proteins. J Mol Biol 376:241–257
Whitney SE, Gidley MJ, McQueen-Mason SJ (2000) Probing expansin action using
cellulose/hemicellulose composites. Plant J 22:327–334
Wilson DB (2009) Cellulases and biofuels. Curr Opin Biotechnol 20:295–299
Wongwilaiwalina S, Rattanachomsria U, Laothanachareona T, Eurwilaichitra L, Igarashib Y,
Champredaa V (2010) Analysis of a thermophilic lignocellulose degrading microbial
consortium and multi-species lignocellulolytic enzyme system. Enzyme Microb Tech
47:283–290
26 M.P. Raghavendra et al.
Wooley R, Ruth M, Glassner D, Sheejan J (1999) Process design and costing of bioethanol
technology: a tool for determining the status and direction of research and development.
Biotechnol Prog 15:794–803
Wyman CE (2007) What is (and is not) vital to advancing cellulosic ethanol. Trends Biotechnol
25:153–157
Xu Q, Gao W, Ding SY, Kenig R, Shoham Y, Bayer EA, Lamed R (2003) The cellulosome system
of Acetivibrio cellulolyticus includes a novel type of adaptor protein and a cell surface
anchoring protein. J Bacteriol 185:4548–4557
Xu Q, Bayer EA, Goldman M, Kenig R, Shoham Y, Lamed R (2004) Architecture of the
Bacteroides cellulosolvens cellulosome: description of a cell surface anchoring scaffoldin and a
family 48 cellulase. J Bacteriol 186:968–977
Yamada R, Hasunuma T, Kondo A (2013) Endowing non-cellulolytic microorganisms with
cellulolytic activity aiming for consolidated bioprocessing. Biotechnol Adv 31(6):754–763
Yang SJ, Kataeva I, Hamilton-Brehm SD, Engle NL, Tschaplinski TJ, Doeppke C, Davis M,
Westpheling J, Adams MW (2009) Efficient degradation of lignocellulosic plant biomass,
without pretreatment, by the thermophilic anaerobe Anaerocellum thermophilum DSM 6725.
Appl Environ Microbiol 75:4762–4769
Zambare VP, Zambare A, Muthukumarappan K, Christopher LP (2011) Potential of thermostable
cellulases in the bioprocessing of switchgrass to ethanol. BioResources 6:2004–2020
Zweifel ME, Barrick D (2001) Studies of the ankyrin repeats of the Drosophila melanogaster
Notch receptor. 2. Solution stability and cooperativity of unfolding. Biochemistry
40:14357–14367
Zweifel ME, Leahy DJ, Hughson FM, Barrick D (2003) Structure and stability of the ankyrin
domain of the Drosophila Notch receptor. Protein Sci 12:2622–2632
Part I
Cellulases
Chapter 2
Cellobiohydrolases: Role, Mechanism,
and Recent Developments
2.1 Introduction
Considering the plant biomass, the main component from plant cell wall is the
cellulose, which is the more abundant carbohydrate constituted by glucopyranose
monomers linked by β-1,4 glycosidic bonds with two distinct regions such as
N. Annamalai (&)
Hawaii Natural Energy Institute, University of Hawaii at Manoa,
1680, East-West Road, Honolulu, HI 96822, USA
e-mail: [email protected]
M.V. Rajeswari
Centre for Ocean Research, Sathyabama University, Jeppiar Nagar,
Chennai, TN 600119, India
N. Annamalai N. Sivakumar
Department of Biology, College of Science, Sultan Qaboos University,
PO Box 36, PC 123 Muscat, Oman
crystalline and amorphous regions (Dashtban et al. 2009). Cellulosic biomass is the
largest amount of waste generated through human activities, which is the most
attractive substrate for ‘biorefinery strategies’ to produce high-value products such
as fuels, bioplastics and enzymes through fermentation processes (Mazzoli et al.
2012). The value of cellulose as a renewable source of energy has made cellulose
hydrolysis the subject of intense research and industrial interest (Bhat 2000). In
recent days, there is an increasing interest on use of biomass for biofuel production
is pointed-out as an important alternative for reduction of energetic and environ-
ment problems.
Cellulases are classified into three major categories based on their substrate
specificities and hydrolysis mechanism as (i) 1,4 β-D-glucanases or endoglucanases
(EC 3.2.1.4), (ii) exoglucanases or cellobiohydrolases (EC 3.2.1.91), and (iii) cel-
lobiase or β-glucosidases (EC 3.2.1.21). The most efficient hydrolysis of cellulose is
the result of combined synergistic actions of cellulases, whereby the enzymatic
activity of an enzyme mixture is significantly higher than individual enzyme
activity (Wood 1992; Irwin et al. 1993; Nidetzky et al. 1994).
CBHs mainly release cellobiose from cellulose derivatives, and the presence of
cellobiose competitively inhibits the rate of hydrolysis. The fungal CBHs belong to
glycosyl hydrolase families 6 and 7 (GH-6 and 7) and act most efficiently on highly
ordered crystalline cellulose, hydrolyzing from either reducing or non-reducing end
to liberate predominantly cellobiose (C2) with a minor amount of cellotriose (C3)
(Boer et al. 2000; Takahashi et al. 2010). There are two types of cellobiohydrolase:
2 Cellobiohydrolases: Role, Mechanism, and Recent Developments 31
CBH I and II that works processively from the reducing and non-reducing end of
the cellulose chain respectively. The cellobiohydrolases Cel6A (a CBH II) and
Cel7a (a CBH I) are much studied enzymes of fungal origin and are usually
expressed by the ascomycete Trichoderma reesei, accounts for 40 % of total pro-
tein and 70 % of cellulytic activity in the industrially relevant fungus Hypocrea
jecorina (T. reesei) which is used for industrial-scale production due to its ability to
produce more sugars from biomass through hydrolysis (Suominen et al. 1993).
The most significant topological feature of CBHs catalytic module is the tunnel
structure, formed by two surface loops that may cover the entirety (family 7 CBHs)
or part of the active site (family 48 CBHs) (Koivula et al. 2002; Vocadlo and
Davies 2008; Zhang and Zhang 2013). The CBH I (TrCBH I) is the major com-
ponent of the T. reesei cellulase system that belongs to the GH-7 family and
hydrolyzes the β-1,4-linkages of a cellulose chain from its reducing end (Claeyssens
et al. 1990). The catalytic core domain (CCD) structures of TrCBH I suggested that
a central structural motif constituted mainly by an antiparallel β-sandwich motif that
bears four surface loops, which decorate the concave face of the sandwich forming
the cellulose-binding tunnel (Divne et al. 1998; Stahlberg et al. 1996). During
enzymatic catalysis, the cellulose chain slides through the substrate-binding tunnel,
and every second glycosidic bond is correctly presented to the catalytic apparatus
located at the far end of the tunnel to progressively liberate disaccharide units
(cellobiose) from cellulose (Divne et al. 1994).
The three-dimensional (3D) structures of two GH-6 family members including
CBH of T. reesei and that of Humicola insolens in complex with glucose, cel-
looligosaccharide, and a non-hydrolyzable substrate analog (Rouvinen et al. 1990;
Varrot et al. 2002). It seems that the significant amino acids associated with cat-
alytic core domain, where the catalytic site is buried inside a tunnel-shaped cavity
and an enzyme–cello-oligosaccharide hydrogen bond network. This typical struc-
ture suggests that the mode of action proceeds in a processive manner as cel-
lobiohydrolase progresses along the cellulose chain (Boisset et al. 2000; Henrissat
1998; Reverbel-Leroy et al. 1997).
Cellobiohydrolases or exoglucanases is able to work effectively on microcrys-
talline cellulose, presumably peeling cellulose chains from the microcrystalline
structure (Teeri 1997). In general, CBHs are active on the crystalline regions of
cellulose; whereas, endoglucanases (1,4-β-glucanases) are typically active on the
more soluble amorphous region of the cellulose crystal (Rouvinen et al. 1990;
Divne et al. 1998). It seems that there is a high degree of synergy observed between
exo- and endoglucanases, which is required for efficient hydrolysis of cellulose
crystals (Abdel-Shakour and Roushdy 2009).
32 N. Annamalai et al.
2.7 Perspectives
References
Maki M, Leung KT, Qin W (2009) The prospects of cellulase-producing bacteria for the
bioconversion of lignocellulosic biomass. Int J Biol Sci 5(5):500–516
Mazzoli M, Lamberti C, Pessione E (2012) Engineering new metabolic capabilities in bacteria:
lessons from recombinant cellulolytic strategies. Trends Biotechnol 30(2):111–119
Medve J, Karlsson J, Lee D, Tjerneld F (1998) Hydrolysis of microcrystalline cellulose by
cellobiohydrolase I and endoglucanase II from Trichoderma reesei: adsorption, sugar
production pattern, and synergism of the enzymes. Biotechnol Bioeng 59(5):621–634
Nidetzky B, Steiner W, Hayn M, Claeyssens M (1994) Cellulose hydrolysis by the cellulases
from Trichoderma reesei: a new model for synergistic interaction. Biochem J 298:705–710
Reverbel-Leroy C, Page S, Belaich A, Belaich JP, Tardif C (1997) The processive endocellulase
CelF, a major component of the Clostridium cellulolyticum cellulosome: purification and
characterization of the recombinant form. J Bacteriol 179:46–52
Riske FJ, Eveleigh DE, Macmillan JD (1990) Double-antibody sandwich enzyme-linked
immunosorbent assay for cellobiohydrolase I. Appl Environ Microbiol 56(11):3261–3265
Rouvinen J, Bergfors T, Teeri TT, Knowles JKC, Jones TA (1990) Three-dimensional structure of
cellobiohydrolase II from Trichoderma reesei. Science 249:380–386
Stahlberg J, Divne C, Koivula A, Piens K, Claeyssens M, Teeri TT (1996) Activity studies and
crystal structures of catalytically deficient mutants of cellobiohydrolase I from Trichoderma
reesei. J Mol Biol 264:337–349
Suominen PL, Mantyla AL, Karhunen T, Hakola S, Nevalainen H (1993) High frequency one-step
gene replacement in Trichoderma reesei. II. Effects of deletions of individual cellulase genes.
Mol Gen Genet 241(5–6):523–530
Teeri TT (1997) Crystalline cellulose degradation: new insights into the function of cellobiohy-
drolases. Trends Biotechnol 15:160–167
van Tilbeurgh H, Pettersson G, Bhikabhai R, De Boeck H, Claeyssens M (1985) Studies of the
cellulolytic system of Trichoderma reesei QM 9414. Reaction specificity and thermodynamics
of interactions of small substrates and ligands with the 1,4-beta-glucan cellobiohydrolase II.
Eur J Biochem 148:329–334
Varrot A, Frandsen TP, Driguez H, Davies GJ (2002) Structure of the Humicola insolens cel-
lobiohydrolase Cel6A D416A mutant in complex with a non-hydrolysable substrate analogue,
methyl cellobiosyl-4-thio-β-cellobioside at 1.9 Å. Acta Crystallogr D Biol Crystallogr
58:2201–2204
Vocadlo DJ, Davies GJ (2008) Mechanistic insights into glycosidase chemistry. Curr Opin Chem
Biol 12:539–555
Wood TM, Bhat KM (1988) Methods for measuring cellulase activities. Methods Enzymol
160:87–117
Wood TM (1992) Fungal cellulases. Biochem Soc Trans 20:46–53
Zhang XZ, Zhang YHP (2013) Cellulases: characteristics, sources, production, and applications.
In: Yang S-T, El-Enshasy HA, Thongchul N (eds) Bioprocessing technologies in biorefinery
for sustainable production of fuels, chemicals, and polymers. Wiley, New York, pp 131–146
Chapter 3
Endo-1,4-β-glucanases: Role, Applications
and Recent Developments
Abstract The biomass conversion processes are highly dependent on the use of
efficient enzymes to degrade polymeric cellulose or hemicellulose into simple sac-
charides, sugars which can be fermented by microorganisms for the production of
valuable fuel and chemicals. The cellulose hydrolysis involves enzymes such as
endo-1,4-β-glucanases (EC 3.2.1.4), cellobiohydrolases (or exo-1,4-β-glucanases)
(EC 3.2.1.91) and β-glucosidases (EC 3.2.1.21). Endo-β-(1,4)-glucanases (EGs) or
β-(1,4)-D-glucan-4-glucanohydrolases (EC 3.2.1.4), which act randomly on soluble
and insoluble β-(1,4)-glucan substrates. EGs breakdown cellulose by attacking the
amorphous regions to produce more accessible new free chain ends for the action of
cellobiohydrolases. The EG-I, II, III, and V, respectively, GH7, 5, 12, and 45 are
most common in natural fungal cellulase mixes. EGs play an important role in
increasing yield of fruit juices, beer filtration, and oil extraction, as well as improving
the nutritive quality of bakery products and animal feed. Reducing sugar assay is the
most convenient and reliable method for EG estimation. However, it is very con-
ventional and time consuming. Recently, a specific and sensitive assay methods have
been developed using substrate mixture comprises of benzylidene end-blocked
2-chloro-4-nitrophenyl-β-cellotrioside (BzCNPG3) and 4,6-O-benzylidene-
4-methylumbelliferyl-β-cellotrioside (BzMUG3).
N. Annamalai (&)
Hawaii Natural Energy Institute, University of Hawaii at Manoa,
1680, East-West Road, Honolulu 96822, HI, USA
e-mail: [email protected]
M.V. Rajeswari
Centre for Ocean Research, Sathyabama University, Jeppiar Nagar,
Chennai 600119, TN, India
T. Balasubramanian
CAS in Marine Biology, Faculty of Marine Sciences, Annamalai University,
Parangipettai 608502, TN, India
3.1 Introduction
Cellulases are used in the textile industry, in detergents, pulp and paper industry,
improving digestibility of animal feeds and food industry. They account for a
significant fraction of industrial enzyme markets. Obviously, cellulase will be the
largest industrial enzyme as compared to the current market size for all industrial
enzymes ($ *2 billion) (Zhang and Zhang 2013). Among the cellulases,
endoglucanases play a key role in increasing the yield of fruit juices, beer filtration
and oil extraction, as well as improving the nutritive quality of bakery products and
animal feed (Bhat 2000). Endo-β-1,4-glucanase has been widely used in industries
to hydrolyze cellulosic materials to sugars in the biofuel production (Wilson 2009),
to increase the rate of wort filtration and reduce the possibility of β-glucan pre-
cipitation in beer and reduce the mash viscosity and turbidity in the brewing
industry (Celestino et al. 2006), to bioremediate pulp waste in the paper industry
3 Endo-1,4-β-glucanases: Role, Applications and Recent … 41
(Ohmiya et al. 1997), and to increase β-glucan digestibility and improve feed
conversion efficiency in the feed industry (Mathlouthi et al. 2002). The
endo-1,4-β-glucanase is also used in the poultry and animal feed industry to
degrade β-glucan and thus increase the digestibility of feeds rich in barley and also
overcome the anti-nutritive effects of 1,3-1,4-β-D-glucan (Choct 2001). Further,
endo-1,4-β-glucanase is used for denim finishing and cotton softening in textile
industries, cleaning and colour care in the detergent industries and to modify fibre
and improves drainage in the paper industries (Cherry and Fidantsef 2003).
Thus, the enormous industrial applications demands endoglucanases with
varying pH and temperature optima, stabilities and substrate specificities. Although
cellulolytic enzymes of Trichoderma reesei have been investigated thoroughly
(Miettinen-Oinonen et al. 2004), the amount of cellulase secreted by this fungus is
insufficient for effective conversion of cellulose to glucose. Moreover, the industrial
demand is increasing day by day especially because of the emergence of second
generation-advanced biofuel industries, which require tremendous amounts of
various enzymes in their processes (Wilson 2009; Yeoman et al. 2010).
Thermostable endoglucanases/cellulases are more favourable for industrial
application due to their high activity and stability at high temperatures to decrease
process costs and increase the efficiencies (Yeoman et al. 2010). However, most
cellulases are not stable at high temperatures, and a number of thermostable cel-
lulases have been purified or cloned and characterized in recent years, some of them
display maximum activities even at 100 °C, such as GH 12 endoglucanase from
Pyrococcus furiosus (100 °C; Bauer et al. 1999), and endoglucanase CelB from
Thermotoga neapolitana (106 °C; Bok et al. 1998). As the main microbial source
of thermophilic endoglucanases, thermophilic fungi like Talaromyces emersonii
and Thermoascus aurantiacus have been reported to produce cellulases of GH 3, 5
and 7 families with temperature optima of 65–80 °C (Gomes et al. 2000; Grassick
et al. 2004).
References
Adsul MG, Singhvi MS, Gaikaiwari SA, Gokhale DV (2011) Development of biocatalysts for
production of commodity chemicals from lignocellulosic biomass. Bioresour Technol
102:4304–4312
Baldrian P, Valaskova V (2008) Degradation of cellulose by basidiomycetous fungi. FEMS
Microbiol Rev 32:501–521
Bauer MW, Driskill LE, Callen W, Snead MA, Mathur EJ, Kelly RM (1999) An endoglucanase,
EglA, from the hyperthermophilic archaeon Pyrococcus furiosus hydrolyzes β-1, 4 bonds in
mixed-linkage (1–3), (1–4)-β-d-glucans and cellulose. J Bacteriol 181:284–290
3 Endo-1,4-β-glucanases: Role, Applications and Recent … 43
Bhat KM, Hay AJ, Claeyssens M, Wood TM (1990) Study of the mode of action and site-specificity
of the endo-(1-4)-beta-D-glucanases of the fungus Penicillium pinophilum with normal,
1-3H-labelled, reduced and chromogenic cello-oligosaccharides. Biochem J 266:371–378
Bhat MK (2000) Cellulases and related enzymes in biotechnology. Biotechnol Adv 18:355–383
Bok JD, Yernool DA, Eveleigh DE (1998) Purification, characterization, and molecular analysis of
thermostable cellulases CelA and CelB from Thermotoga neapolitana. Appl Environ
Microbiol 64:4774–4781
Celestino KRS, Cunha RB, Felix CR (2006) Characterization of a β-glucanase produced by
Rhizopus microsporus var. microsporus, and its potential for application in the brewing
industry. BMC Biochem 7:23
Cherry JR, Fidantsef AL (2003) Directed evolution of industrial enzymes: an update. Curr Opin
Biotechnol 14:438–443
Choct M (2001) Enzyme supplementation of poultry diets based on viscous cereals. In:
Bedford M, Partridge G (eds) Enzymes in farm animal nutrition. CABI Publishing, Oxon,
pp 145–160
Chundawat SPS, Lipton MS, Purvine SO, Uppugundla N, Gao D, Balan V, Dale BE (2011)
Proteomics-based compositional analysis of complex cellulase-hemicellulase mixtures.
J Proteome Res 10:4365–4372
Claeyssens M, Henrissat B (1992) Specificity mapping of cellulolytic enzymes: classification into
families of structurally related proteins confirmed by biochemical analysis. Protein Sci 1:
1293–1297
Du J, Shao Z, Zhao H (2011) Engineering microbial factories for synthesis of value-added
products. J Ind Microbiol Biotechnol 38:873–890
Foreman PK, Brown D, Dankmeyer L, Dean R, Diener S, Dunn-Coleman NS, Goedegebuur F,
Houfek TD, England GJ, Kelley AS, Meerman HJ, Mitchell T, Mitchinson C, Olivares HA,
Teunissen PJ, Yao J, Ward M (2003) Transcriptional regulation of biomass-degrading enzymes
in the filamentous fungus Trichoderma reesei. J Biol Chem 278:31988–31997
Gomes I, Gomes J, Gomes DJ, Steiner W (2000) Simultaneous production of high activities of
thermostable endoglucanase and β-glucosidase by the wild thermophilic fungus Thermoascus
aurantiacus. Appl Microbiol Biotechnol 53:461–468
Grassick A, Murray PG, Thompson R, Collins CM, Byrnes L, Birrane G, Higgins TM, Tuohy MG
(2004) Three-dimensional structure of a thermostable native cellobiohydrolase, CBH IB, and
molecular characterization of the cel7 gene from the filamentous fungus, Talaromyces
emersonii. Eur J Biochem 271:4495–4506
Henrissat B (1991) A classification of glycosyl hydrolases based on amino acid sequence
similarities. Biochem J 280:309–316
Himmel ME, Xu Q, Luo Y, Ding SY, Lamed R, Bayer EA (2010) Microbial enzyme systems for
biomass conversion: emerging paradigms. Biofuels 1(2):323–341
Kanauchi M, Bamforth CW (2008) The relevance of different enzymes for the hydrolysis of
β-glucans in malting and mashing. J Inst Brew 114(3):224–229
Lange JP (2007) Lignocellulose conversion: an introduction to chemistry, process and economics.
Biofuel Bioprod Biorefin 1:39–48
Levy I, Shani Z, Shoseyov O (2002) Modification of polysaccharides and plant cell wall by
endo-1,4-beta-glucanase and cellulose-binding domains. Biomol Eng 19:17–30
Li Y, Irwin DC, Wilson DB (2010) Increased crystalline cellulose activity via combinations of
amino acid changes in the family 9 catalytic domain and family 3c cellulose binding module of
Thermobifida fusca Cel9A. Appl Environ Microbiol 76:2582–2588
Libertini E, Li Y, McQueen-Mason SJ (2004) Phylogenetic analysis of the plant
endo-beta-1,4-glucanase gene family. J Mol Evol 58:506–515
Mangan D, McCleary BV, Liadova A, Ivory R, McCormack N (2014) Quantitative fluorometric
assay for the measurement of endo-1,4- β -glucanase. Carbohydr Res 395:47–51
Mathlouthi N, Mallet S, Saulinier L, Quemener B, Larbier M (2002) Effects of xylanase and
beta-glucanase addition on performance, nutrient digestibility and physic-chemical conditions
44 N. Annamalai et al.
in the small intestine contents and caecal microflora of broiler chickens fed a wheat and
barley-based diet. Anim Res 51:395–406
McCleary BV, McKie V, Draga A (2012) Measurement of endo-1,4-β-glucanase. Methods
Enzymol 510:1–17
McCleary BV, Mangan D, Daly R, Fort S, Ivory R, McCormack N (2014) Novel substrates for the
measurement of endo-1,4-β-glucanase (endo-cellulase). Carbohydr Res 385:9–17
Miettinen-Oinonen A, Londesborough J, Joutsjoki V, Lantto R, Vehmaanpera J (2004) Three
cellulases from Melanocarpus albomyces for textile treatment at neutral pH. Enzyme Microb
Technol 34:332–341
Ohmiya K, Sakka K, Karita S, Kimura T (1997) Structure of cellulases and their applications.
Biotechnol Genet Eng Rev 14:365–414
Rahman M, Bhuiyan SH, Nirasawa S, Kitaoka M, Hayashi KJ (2002) Characterization of an
endo-β-1,4-glucanase of Thermotoga maritima expressed in Escherichia coli. Appl Glycosci
49:487–495
Sandgren M, Stahlberg J, Mitchinson C (2005) Structural and biochemical studies of GH family 12
cellulases: improved thermal stability and ligand complexes. Prog Biophys Mol Biol 89:
246–291
Sarria-Alfonso V, Sierra JS, Morales MA, Rojas IG, Sarmiento NM, Pinales P (2013) Culture
media statistical optimization for bio-mass production of a ligninolytic fungus for future rice
straw degradation. Indian J Microbiol 53(2):199–207
Sateesh L, Rodhe A, Naseeruddin S, Yadav K, Prasad Y, Rao L (2012) Simultaneous cellulase
production, saccharification and detoxification using dilute acid hydrolysate of S. spontaneum
with Trichoderma reesei NCIM 992 and Aspergillus niger. Indian J Microbiol 52:258–262
Sathitsuksanoh N, Zhu Z, Ho T-J, Bai M-D, Zhang Y-HP (2010) Bamboo saccharification through
cellulose solvent based biomass pretreatment followed by enzymatic hydrolysis at ultra-low
cellulase loadings. Bioresour Technol 101:4926–4929
Sipos B, Benko Z, Dienes D, Reczey K, Viikari L, Siika-aho M (2010) Characterisation of specific
activities and hydrolytic properties of cell-wall-degrading enzymes produced by Trichoderma
reesei Rut C30 on different carbon sources. Appl Biochem Biotechnol 161:347–364
Somerville C, Youngs H, Taylor C, Davis SC, Long SP (2010) Feedstocks for lignocellulosic
biofuels. Science 13:790–792
Somogyi MJ (1952) Notes on sugar determination. Biol Chem 195:19–23
Sweeney MD, Xu F (2012) Biomass converting enzymes as industrial biocatalysts for fuels and
chemicals: recent developments. Catalysts 2:244–263
Urbanowicz BR, Bennett AB, Del Campillo E, Catala C, Hayashi T, Henrissat B, Hofte H,
McQueen-Mason SJ, Patterson SE, Shoseyov O, Teeri TT, Rose JKC (2007) Structural
organization and a standardized nomenclature for plant endo-1,4-beta-glucanases (cellulases)
of glycosyl hydrolase family 9. Plant Physiol 144:1693–1696
Vlasenko E, Schulein M, Cherry J, Xu F (2010) Substrate specificity of family 5, 6, 7, 9, 12, and
45 endoglucanases. Bioresour Technol 101:2405–2411
Wang K, Luo H, Bai Y, Shi P, Huang H, Xue X, Yao B (2014) A thermophilic
endo-1,4-β-glucanase from Talaromyces emersonii CBS394.64 with broad substrate specificity
and great application potentials. Appl Microbiol Biotechnol 98:7051–7060
Wilson DB (2008) Three microbial strategies for plant cell wall degradation. Ann N Y Acad Sci
1125:289–297
Wilson DB (2009) Cellulases and biofuels. Curr Opin Biotechnol 20(3):295–299
Yeoman CJ, Han Y, Dodd D, Schroeder CM, Mackie RI, Cann IK (2010) Thermostable enzymes
as biocatalysts in the biofuel industry. Adv Appl Microbiol 70:1–55
Yu L, Sun J, Li L (2013) PtrCel9A6, an endo-1,4-β-glucanase, is required for cell wall formation
during xylem differentiation in populous. Mol Plant 6(6):1904–1917
Zhang Y-HP (2008) Reviving the carbohydrate economy via multi-product biorefineries. J Ind
Microbiol Biotechnol 35:367–375
3 Endo-1,4-β-glucanases: Role, Applications and Recent … 45
Zhang XZ, Zhang Y-HP (2013) Cellulases: characteristics, sources, production, and applications.
In: Yang ST, El-Enshasy HA, Thongchul N (eds) Bioprocessing technologies in biorefinery for
sustainable production of fuels, chemicals and polymers. Wiley, New York, pp 131–146
Zhu Z, Sathitsuksanoh N, Zhang P (2009) Direct quantitative determination of adsorbed cellulase
on lignocellulosic biomass with its application to study cellulase desorption for potential
recycling. Analyst 134:2267–2272
Chapter 4
The Role and Applications
of b-Glucosidases in Biomass Degradation
and Bioconversion
4.1 Introduction
H. Ouyang (&)
Novozymes North America Inc., 77 Perrys Chapel Church Road, PO BOX 576, Franklinton,
NC 27525, USA
e-mail: [email protected]
F. Xu
Novozymes Inc., 1445 Drew Ave, Davis, CA 95618, USA
e-mail: [email protected]
Historically, BGLs have been categorized into three groups based on the substrates:
aryl-BGLs, true cellobiases, and broad substrate specificity BGLs. Canonical BGLs
that catalyze the hydrolysis of b-glucopyranosides are assigned with the Enzyme
Commission (E.C.) number E.C.3.2.1.21. However, given the variety of b-linkage
terminal nonreducing glycosides in nature, many E.C. numbers have been assigned
to a larger group of enzymes that hydrolyze b-linkage glycosidic bond. For glucan
substrates, glucan 1,3-b-glucosidase (3.2.1.58) and glucan 1,4-b-glucosidase
(3.2.1.74) remove successively Glc from the nonreducing end with 1,3 and 1,4
b-linkage to the polysaccharide chain. Enzymes with hydrolytic activity on
b-glucosides with aglycon moieties also fall into E.C.3.2.1 family. These include
glucosylceramidase (3.2.1.45), steryl-b-glucosidase (3.2.1.104), 3a(S)-strictosidine
b-glucosidase (3.2.1.105), amygdalin-b-glucosidase (3.2.1.117), prunasin
b-glucosidase (3.2.1.118), vicianin b-glucosidase (3.2.1.119), raucaffricine
b-glucosidase (3.2.1.125), coniferin b-glucosidase (3.2.1.126), and b-D-glucopyr-
anosyl abscisate b-glucosidase (3.2.1.175). More complicated substrates like
thio-linkage, 7-[b-D-apiofuranosyl-(1, 6)-b-D-glucopyranosyloxy]isoflavonoid, b-D-
galactopyranose-(1, 4)-a-L-galactopyranose-6-sulfate in porphyran, or even 3-O-(N-
acetyl-b-D-glucosaminyl)-L-serine/threonine in proteins can be hydrolyzed by
thioglucosidase (myrosinases) (3.2.1.147), b-apiosyl-b-glucosidase (3.2.1.161),
b-porphyranase (3.2.1.178) and protein O-GlcNAcase (3.2.1.169) respectively.
E.C. numbering system is a classification based on the chemical reactions which
enzymes catalyze. It has little information on structure, mechanism, or phylogenetic
4 The Role and Applications of b-Glucosidases in Biomass … 49
One of the functions for BGL is being a part of the synergistic cellulase system that
breaks down cellulose, deployed by a large population of organisms including
bacteria, fungi (molds, mushrooms, yeasts), and animals (e.g., termites). BGL is
50 H. Ouyang and F. Xu
4.3.2 Plants
Plant BGLs have been demonstrated with a wide functional diversity in defense,
symbiosis, cell wall catabolism and lignification (monolignol conversion), signaling
and secondary metabolism. The wide spectrum of biological activity is supported by a
large number of BGLs from different GH families. For instance, 48 GH1 genes are
found in Arabidopsis thaliana (Xu et al. 2004), while the number in rice is 40
(Opassiri et al. 2006). These GH1 enzymes cover a cluster of classical myrosinases
and another cluster of myrosinases and BGLs found in the endoplasmic reticulum
(ER) and peroxisome. Interestingly, GH1 enzymes from chloroplast of other plants
such as maize, sorghum and wheat are much more closely related to homologues in
thermophilic bacteria and archaea than rice and Arabidopsis enzymes (Thorlby et al.
2004). In addition, most plants also encode BGLs from GH families 3 and 5 (Opassiri
et al. 2007; Arthan et al. 2006). To defend themselves, plants have developed a wide
range of molecules. The first group is cyanogenic b-glycosides, including linamarin
from clover and cassava, dhurrin from sorghum and prunasin from cherry. These
cyanogenic glycosides can be hydrolyzed by BGL and release their a-hydroxynitrile
moiety, which then breakdown either enzymatically or spontaneously to yield cya-
nide (Poulton 1990; Morant et al. 2008a, b). The second group consists of non-
cyanogenic molecules, such as b- and c-hydroxynitriles (Morant et al. 2008a, b) and
isoflavones in legumes (Suzuki et al. 2006), flavonoids (Chuankhayan et al. 2005),
coumarins, hydroxaminic acids (e.g., 2,4-dihydroxy-7-methoxy-1,4-benzoxazin-
3-one) in maize and wheat (Esen 1992; Sue et al. 2006; Nikus et al. 2001), and
4 The Role and Applications of b-Glucosidases in Biomass … 51
saponins (Nisius 1988). These defensive molecules are also stored as b-D-glucosides
and unleashed by specific BGLs.
In the symbiosis between the endophytic fungus Piriformospora indica and
plant Arabidopsis, the fungus prevents Arabidopsis from overgrowing the roots and
ends up triggering a defense response (Sherameti et al. 2008; Nagano et al. 2005).
A special BGL in ER, AtBGLU23, is revealed to be critical in this symbiosis (Lipka
et al. 2005; Bednarek et al. 2009) by demonstrated activity on scopolin, the most
abundant coumarin glucoside in root (Ahn et al. 2010). Upon cell wall disruption,
plant defensive BGLs and myrosinases often bind to cytoplasmic aggregating
factors, which are believed to localize the otherwise soluble glucosidases at the site
and therefore generate the defense compounds on site (Nisius 1988; Esen and
Blanchard 2000; Kittur et al. 2007). In addition to the plant–microbe interaction,
BGL may also be involved in the plant–insect interaction.
BGLs also play an important role in cell wall metabolism. Several BGLs that
hydrolyze cell wall derived oligosaccharides have been identified. One BGL in
germinating barley seedlings showed activity toward b-1,3- and b-1,4-linked
oligosaccharides (Leah et al. 1995; Hrmova et al. 1998), preferentially man-
nooligosaccharides, which are found in barley endosperm cell walls (Hrmova et al.
2006). Rice seedling BGLs Os3BGlu7, Os3BGlu8, and Os3BGlu26 demonstrate an
increased activity as degree of polymerization (DP) increase from 2 to 6
(Kuntothom et al. 2009), while Os4BGlu12 (Opassiri et al. 2006) is not sensitive to
substrate DP, and Os3BGlu6 (Seshadri et al. 2009) is most efficient on disaccha-
rides. Os3BGlu7 and Os3BGlu8 are widely expressed in rice tissues and are thus
believed able to remodel cell wall by releasing Glc from oligosaccharides derived
from it. Some other BGLs play roles in lignification of secondary cell walls. Two
examples are coniferin BGL from lodgepole pine tree xylem found in differentiating
region of the xylem (Dhamawardhana et al. 1995) and BGL (AtBGLU45/46) from
Arabidopsis. The later has been found active in hydrolyzing coniferin, syringin and
coumaryl alcohol glucosides to release monolignols (Escamilla-Trevino et al.
2006).
Phytohormone activation in many cases also requires glucosides hydrolysis by
BGLs, such as the maize BGL Zm-p60.1 which hydrolyzes and activates cytokinin
b-glucosides (Brzobohaty et al. 1993), an extracellular BGL which hydrolyzes
abscisic acid glucosyl (Glc) ester transported from roots to leaves (Dietz et al.
2000), the BGL hydrolyzing auxin Glc ester 6-O-(4-O)-indole-3-ylacetyl-b-D-Glc
(Jakubowska and Kawalczyk 2005), and the Arabidopsis ER BGL AtBG1 that
hydrolyzes abscissic acid Glc ester in response to drought stress (Lee et al. 2006).
Mutation in this Arabidopsis AtBG1 leads to early germination and defective
stomata closing.
BGLs can also play metabolic roles to remove Glc protective group from
metabolic intermediates and consequently trigger the biosynthesis of various natural
products (phytochemicals) such as monoterpene alkaloids. For instance, one of the
downstream products of strictosidine is raucaffricine, a glucoside that could only be
further metabolized into ajmaline after deglucosylated by raucaffricine BGL
(Barleben et al. 2007).
52 H. Ouyang and F. Xu
Plant BGLs can also release volatile terpenols from glycosides as flower and
fruit scents (Reuveni et al. 1999) or attractants for parasitoid wasp to attack her-
bivores parasites (Mattiacci et al. 1995).
4.3.3 Animals
Some insects feeding on plants also developed GH1 enzymes in their larvae stage,
such as BGLs from Tenebrio molitor and Spodoptera frugiperda (Ferrieira et al.
2001; Marana et al. 2001) and myrosinases from cabbage aphid Brevicoryne
brassicae (Jones et al. 2001). These enzymes have hydrolysis capability on
gluco-oligosaccharides and plant cellobiose, gentiobiose and even amygdalin
(Marana et al. 2001) and share higher sequence similarity with vertebrate
lactase-phloridzin hydrolase (LPH) than with GH1 from plants, indicating their
further divergence from the same animal ancestor after their first divergence from
plants.
Mammalian cells contain BGLs for metabolism of glycolipids and dietary gluco-
sides. Human acid BGLs releases Glc ceramide; defects in this enzyme lead to gly-
coceramides accumulation and eventual Gaucher disease (Butters 2007). Human LPH
is involved in food digestion and displays activity on b-glucosides, b-galactosides and
phloridzin, a polyphenol carrying glycoside found in many fruits. Deficiency of this
enzyme ends up with the widespread lactose intolerance (Tribolo et al. 2007). Human
acid BGL GBA1 and bile acid BGL GBA2 belong to the GH30, and participate in the
metabolism of glycolipids and dietary glucosides (Berrin et al. 2002) (Fig. 4.1).
4.4 Structure
BGLs from various GH families can have amino acid sequences ranging from 420
to 840 units, some comprising signal peptides for secretion or other domains in
addition to the catalytic core. The structures of the catalytic core (Fig. 4.2) can be
divided into three groups. The first one is a (b/a)8-barrel found in GH1, GH5, and
GH30. The second one contains an (a/a)6-barrel found in GH9. The last one
consists of a (b/a)8-barrel and an a/b sandwich in which a 6-stranded b-sheet is
flanked by two sets of three a-helices, and is found in GH3. The (b/a)8-barrel group
features a catalytic domain ranging from 420 to 560 residues in length, depending
on the highly variable loop regions on the b-strands at the C terminus
(Sanz-Aparicio et al. 1998). Within the catalytic domain, two conserved Glu or Asp
from b-strands 4 and 7 respectively are identified as the catalytic acid/base and
nucleophilic residue (Ketudat Cairns and Esen 2010; Vallmitjana et al. 2001;
Henrissat et al. 1995). Monomers of GH1 enzymes can dimerize or form even
higher orders of aggregates. The GH5 BGLs have a (b/a)8-barrel core structure as
the GH1 BGLs (both belonging to the GH-A clan), and the rice GH5 BGL has a
4 The Role and Applications of b-Glucosidases in Biomass … 53
OH OH
OH OH HO OH
HO
HO HO O
O O HO O O
HO HO HO
O OH
O HO O
O OH
OH OH CN
O
CN OH
NC CN
OH
OH
O HO
HO HO O
HO O
O
O OH
OH OH
O N OH
HO
O
HO O O
R
O O HO Daidzin R = H
OH
OH
O Genistin R = OH
O
DIMBOA beta-glucoside Apigenin 7-O-beta-glucoside
OH OH
OH
OH
OH HO
OH HO O
HO O
O HO O
HO O OH
HO
O OH
OH O O
OH
OH
Coniferin Phloridzin
HO
O HO
OH
OH
HO
HO O HO
HO O
OH O NH
OH O
O
6
OH
OH
11
OH
HO O Salicin
O
Glucosyl ceramide
HO
O H
OH OH
OH OH
O O
Abscisic acid glucosyl ester
HO
O O OH
OH HO
H O O HO
OH N HO
O OH
HO H
HO N
HO O O Cellobiose
OH
HN
O
OH
OH
OH HO
HO
Indoxyl beta-glucoside Strictosidine HO O O
O
OH OH OH
Laminaribiose
Fig. 4.1 Structures of some BGLs substrates involved in various physiological processes. 1 Plant
defense: 1a cyanogenic glucosides: linamarin (clover/cassava), dhurrin (sorghum), prunasin (cherry)
and its precursor D-amygdalin. 1b noncyanogenic glucosides: 2,4-dihydroxy-7-methoxy-1,4-
benzoxazin-3-one 2-O-b-glucoside (DIMBOA-Glc, maize/wheat), apigenin 7-O-b-glucoside,
daidzin, genistin, phloridzin. 2 Plant cell wall metabolism: coniferin, p-coumaryl alcohol 4-O-
b-glucoside. 3 Plant phytohormone activation: abscisic acid glucosyl ester, indoxyl b-glucoside,
salicin. 4 Plant secondary metabolism: strictosidine, precursor to a wide range of monoterpene
alkaloids. 5 Fungal and bacterial degradation of plant primary cell walls: cellobiose, laminaribiose. 6
Animal (structural, signaling and metabolic functions): glucosylceramide
54 H. Ouyang and F. Xu
fascin-like domain (Opassiri et al. 2007). The (a/a)6-barrel group seen in GH9 is
dominated by a large number of endoglycosidases and leave only a few BGLs like
that from V. cholera. So far there is no conclusive identification of the catalytic
residues on the BGL from this family. However, the study on another GH9 exo-
cellulase has identified a Glu residue on the loop connecting a-helices 12 and 13 as
the catalytic acid/base, while a water molecule activated by two conserved,
hydrogen bond-forming Asp acts as the nucleophilic residue (Schubot et al. 2004).
A similar situation might be expected in BGLs from the same family. For the GH3
enzymes, the interface of the (b/a)8-barrel and a/b sandwich harbors the catalytic
nucleophilic residue (a conserved Asp) and acid/base, which may be variable in
terms of position and identity among different BGLs (Gabrisko and Janecek 2015;
Pozzo et al. 2010; Varghese et al. 1999). A fibronectin type III domain is found at
the C-termini of Thermotoga neapolitana and Aspergillus aculeatus BGLs, and the
domain’s function remains unclear.
Fig. 4.2 Structures of BGLs from different GH families. GH1: T. reesei BGL 2 (PDB: 3AHY),
GH3: T. reesei BGL 1 (PDB: 3ZYZ), GH5: Saccharomyces cerevisiae flavonoid BGL (PDB:
1H4P), GH9: Vibrio parahaemolyticus BGL-like protein (PDB: 3H7L), GH30: Homo sapiens acid
BGL (PDB: 1Y7 V)
4 The Role and Applications of b-Glucosidases in Biomass … 55
All BGLs mentioned so far are exoglycoside hydrolases working on the nonre-
ducing end of a cellodextrin or glucoside substrate. Glucosidases initiate
b-glucosidic bond cleavage with a nucleophilic attack on the anomeric carbon of
the Glc. This is carried out by either the side chain carboxyl of the conserved
catalytic Glu/Asp (Keresztessy et al. 1994), or a water molecule activated by a
conserved Asp. The oxygen in the b-glycosidic bond is activated by the conserved
proximal catalytic acid while the anomeric carbon and the ring oxygen form an
oxocarbenium-like ionic transition state (Fig. 4.3). When the water serves as the
nucleophile (as in the case of GH9), the transition moves forward as the aglycone
moiety leaves the anomeric carbon in concert with the hydroxyl (from the activated
water) displacement on the same carbon to release the Glc (Qi et al. 2008). This
leads to the inversion of chirality on the anomeric carbon and thus is referred as the
inverting mechanism. With the endogenous nucleophile, the first transition state
leads to the leaving of the aglycone moiety as well; however, the b-glycosidic bond
on the anomeric carbon is now displaced with the same carboxyl that initiates the
nucleophilic attack. This a-linked covalent enzyme-glycone intermediate then
undergoes another round of nucleophilic attack on the anomeric carbon. The cat-
alytic acid in the first step gets deprotonated after activating aglycone leaving group
and turns into a catalytic base. It then activates a water to attack the anomeric
carbon to release the Glc from the enzyme-glycone a-linkage. This retaining
mechanism allows the product keep the anomeric chirality. In myrosinases, the
conserved catalytic Glu in the second step is replaced by a Gln, and therefore an
ascorbic acid is required to fulfill the role of the catalytic base.
Unequivocal evidences of the covalent intermediates have been obtained through
incubation of Agrobacterium sp. GH1 BGL with 2,4-dinitrophenyl-2-deoxy-2-
fluoroglucoside (2F-DNPG) (Withers et al. 1987; Withers and Street 1988). The
electronegative property of fluoride destabilizes the transition state upon binding in
the active site, 2,4-dinitrophenyl aglycone moiety is a highly reactive leaving group;
both favor the formation of a lasting and stable enzyme-fluoroglycone intermediate.
The presence of fluoride was firstly supported by 19F-NMR (Withers and Street 1988)
and later observed in crystal structures (Hrmova et al. 2001; Chuenchor et al.
2008) and inactivated enzyme pepsin digest (Withers et al. 1990). In a similar
approach, [3H] conduritol B epoxide (CBE) forms a trapped, radioactively labeled
glycosyl-enzyme intermediate in Aspergillus wentii BGL A3. Radioactivity detected
in trypsin-cleaved fragments establishes the catalytic role of the nucleophilic Asp,
albeit CBE possesses less specificity than 2F-DNPG and labels nearby residues
(Bause and Legler 1980). Nevertheless, the majority of the catalytic residues have
been identified through homology, mutagenesis, chemical modification, and crystal
structure analyses. For instance, the catalytic acid/base residue and nucleophile
residue have been inactivated by mutagenesis, respectively, in the investigation of
retaining mechanism. Replacement of the nucleophile with a small non-nucleophilic
residue, e.g., Ala and Gly, inactivates the enzyme in attacking the anomeric carbon
56 H. Ouyang and F. Xu
O
O O
H O
OH OH OH
OH O
RO
HO O O O
HO HO
HO OR HO HO
R OH
OH OH OH
O H O H OH
H H
O O HO
O O O
ES E+P
TS
O O
OH
OH H O
OH
HO O
HO OR RO
HO O
OH HO
O OH O
R OH
O O
ES O
OH
O
TS
HO O H
HO O
OH H
O
O O O
OH
OH H O
OH
HO O Intermediate + P
HO OH O H
HO O
OH HO
O OH O
O O
E+P TS
Fig. 4.3 The inverting and retaining mechanisms of BGLs. The inverting mechanism in
Scheme A is found in the GH9 BGL where water-anomeric carbon bond is formed in a single step
of displacement reaction. However, in the retaining mechanism (Scheme B), water does not attack
the anomeric carbon in the first displacement step which is done in a similar way by the catalytic
nucleophile. This leads to a covalently bonded glycosyl intermediate with inverted anomeric
carbon conformation. The following hydrolysis step ends up with an invertion of anomeric carbon
conformation for the second time and a retaining glycoside product as a result. ES enzyme–
substrate complex, TS 6¼ transition state, E + P (resting) enzyme and product
4 The Role and Applications of b-Glucosidases in Biomass … 57
Compared with the thorough characterization of the catalytic residues and their
mechanism in b-glycosidic bond cleavage, less is known about how the interaction
between BGLs and their substrates gives rise to the tremendous substrate diversity
and how the interaction determines BGLs’ substrate specificity. Crystal structures
have revealed that the binding pockets are cavities composed of several subsites. The
subsite that binds the Glc (glycone) unit is assigned subsite −1, whereas the aglycone
moiety binds to the subsites +1, +2, +3 and so on. The substrate specificity of BGL is
a combination of both Glc and aglycone specificities (Marana 2006).
Besides glucosides and cellodextrins, BGLs may act on other glycosides or
oligosacchardies. Such non-BGL activities are related to the coexistence of other
glycosidases in the GH families that BGLs belong to, a likely divergent evolution
effect. The same BGL may display different activities toward substrates with dif-
ferent glycosyls or glycones. For instance, S. frugiperda BGL hydrolyzes fucosides
40 times more quickly than galactosides (Marana et al. 2001). Aspergillus oryzae
GH3 BGL binds Glc well, but mannose (2-epimer of Glc) much more weakly, and
no allose or galactose (3- and 4-epimer) (Langston et al. 2006). GH1 BGLs might
act on b-D-mannosides, b-D-galactosides, b-D-fucosides and b-D-glucuronides,
while GH3 BGLs might act on b-D-xylosides and a-L-arabinosides. Hydrogen
bonding plays a key role in the BGL-substrate interaction (Fig. 4.4). In a set of
studies, the hydrogen bonds involving the glycone hydroxyls and subsite −1 are
studied in three different GH1 BGLs: human LPH (Fernandez et al. 1995),
Agrobacterium sp. Agbglu (Namchuck and Withers 1995) and S. frugiperda Sfbglu
(Marana et al. 2002). For LPH and Agbglu, replacement of the glycone hydroxyls
by hydrogen and/or fluorine is evaluated using enzyme kinetics. In the case of
Sfbglu, a combination of site-directed mutagenesis and enzyme kinetics is deployed
to determine the energy of isolated interaction between subsite −1 and glycone OH
in the glycosylation transition state. The analysis leads to the conclusion that the
specificity of glycone is controlled by a hydrogen bond network involving at least
five amino acid residues and four substrate hydroxyls. A Glu bridging the glycone
hydroxyls 4 and 6 are key in differentiating fucosides, glucosides, and galactosides.
In addition, interactions with the hydroxyl 2 are essential to the BGL activity. In
another set of studies, one GH3 cellodextrinase and three b-xylosidases from rumen
58 H. Ouyang and F. Xu
bacterium Prevotella bryantii B14 are examined for their gylcone specificity by
comparing kinetic parameters on different substrates (Dodd et al. 2010). Mutations
of each conserved residues in subsite −1 severely decrease the catalytic efficiency
without changing the overall structure of the enzyme variants. Similar to GH1
enzymes, replacing the conserved Glu115 (Asp120 in barley glucan b-glucosidase)
leads to 1.1-fold increase in kcat/Km for para-nitrophenyl-b-D-glucopyranoside
(pNP-Glc) and 14-fold decrease for pNP-b-D-xylopyranoside. Glu115 is therefore
believed to contribute to the discrimination between b-xylosides and b-glucosides.
In BGL, the conserved Asp forms hydrogen bond bridging the hydroxyls 4 and 6 of
the Glc unit, while in b-xylosidase, the conserved Glu forms hydrogen bond with
the hydroxyl 4 (Fig. 4.4b).
Fig. 4.4 Elements which are important in glycone (Dodd et al. 2010) and aglycone binding
subsites (Czjzek et al. 2000; Verdoucq et al. 2004). a GH3 active site residues for the b-glucan
exohydrolase (ExoI) from barley bound to Glc are shown (PDB: 1EX1). b Models for GH3
enzyme’s xylosides and glucosides recognition (Dodd et al. 2010). Asp120 from barley ExoI
forms hydrogen bond with the hydroxyls 4 and 6 of Glc. Glu115 from P. bryantii B14 Xyl3B
forms hydrogen bond with the hydroxyl 4 of xylose and may discriminate between Glc and xylose
on the basis of steric interactions. c, d Active site configurations from aglycone side of maize
BGL1 (ZMGlu1, c) and sorghum dhurrinase 1 (Dhr1, d), which are shown for the structures of the
ZMGlu1 Glu189Asp mutant in complex with DIMBOA-Glc (PDB: 1E56) and Dhr1 Glu191Asp
mutant in complex with dhurrin (PDB: 1V03). Phe261 residue narrows the active site in Dhr1 and
is also shown in front of the catalytic nucleophile Glu404
4 The Role and Applications of b-Glucosidases in Biomass … 59
The BGL’s interaction with the aglycone part of the substrate may be more
diverse or complex, as shown by various structure, mutagenesis and chimera
studies. In contrast to the highly conserved amino acid residues at the subsite −1, a
different though overlapping set of residues is involved in interacting with the
aglycone part of the substrate in BGL. An archetypical case is found in the maize
ZmGlu1, a GH1 enzyme, which is active on a wide range of aglycones including
DIMBOA. The structure of one of its inactive variant in complex with
DIMBOA-Glc, DIMBOA and the inhibitor dhurrin shows that the aglycone is
sandwiched by Phe198, Phe205 and Phe466 on one side and Trp378 on the other
(Czjzek et al. 2000). In addition, DIMBOA has a hydrophobic contact with Ala467.
All these residues, except Trp378, are variable among BGLs. For instance, sorghum
dhurrinase SbDhr1 conserves this Trp on the same side of the algycone subsite and
use Val196, Leu203 and Ser462 on the other side in replacement of the three Phe
residues for a tighter and more polar interaction (Verdoucq et al. 2004). The
mutagenesis of the aglycone binding residues in the Zm60.8 isoform of the ZmGlu1
further confirms the importance of these residues (Zouhar et al. 2001). It is thus
believed the conserved Trp is required to bind any aglycone and three or four
homologous positions on the other side would shape the pocket and therefore define
the aglycone specificity. Other factors may contribute to the aglycone binding as
well. It is discovered that the conserved Trp may also differentiate substrates like
dhurrin and strictosidine via its positional variation (Barleben et al. 2007; Verdoucq
et al. 2004; Cicek et al. 2000). Closely related wheat BGL (TaGlu) also hydrolyzes
DIMBOA-Glc, although a different set of residues are present in the binding pocket
(Sue et al. 2006). In the case of rice BGlu1, two of the three corresponding Phe
positions are replaced by Leu and Asn, which makes either no or indirect inter-
action with oligosaccharide substrates (Chuenchor et al. 2008). Instead, Asn245
plays a key role in binding to the third Glc moiety in the substrate, while the
corresponding residues in SbDhr1 and rice Os3BGlu6 appear to block the entrance
to their active sites, with no binding of a trimeric substrate (Opassiri et al. 2003;
Verdoucq et al. 2004). In general, the GH3 BGLs have fewer restraints on the
aglycone part of the substrate than the GH1 BGLs. For instance, A. oryzae
GH3 BGL has similar activities toward cellobiose, sophorose, laminaribiose, and
b-gentiobiose, in which the activated Glc unit (at the subsite −1) is linked to the
‘leaving’ Glc unit (corresponding to the aglycones of 4-glucosides) via 1,4-, 1,2-,
1,3-, and 1,6-glucosidic bond, respectively (Langston et al. 2006).
The GH1 BGLs have narrow and deep wells-like active sites, whose topology
(including space restricted subsites +1 or higher), electrostatic and hydrophobic
interaction properties impact the BGLs’ substrate specificity and product
(Glc) inhibition tolerance (de Giuseppe et al. 2014; Tribolo et al. 2007). The GH3
BGLs have crater-shaped active sites, whose depth, shape, and local amino acids
determine the BGLs’ substrate specificity. The topological difference is believed to
contribute to the difference in substrate reactivity and product inhibition between
the GH1 and GH3 BGLs (de Giuseppe et al. 2014).
60 H. Ouyang and F. Xu
When the substrate concentration is relatively low, BGLs’ main activity is hydro-
lytic, cleaving cellodextrins or oligosaccharides into smaller carbohydrate mole-
cules such as Glc, or glucosides into carbohydrate and aglycone subunits. Various
substrates may be used to assay BGLs’ properties. Chromogenic and fluorogenic
substrates, exemplified by pNP-Glc and 4-methylumbelliferyl-b-D-Glc, are quite
reactive and convenient for spectrophotometric detections, which are suitable for
microplate-based high-throughput or robotized studies (Perry et al. 2007). Because
the chromo- or fluorogenic aglycones are often unnatural, these substrates might not
yield data applicable for quantitative or comprehensive understanding of BGLs’
active sites, especially the subsites +1, +2, etc. Cellodextrins, other oligosaccha-
rides, and glucosides (such as that in Fig. 4.1), are able to reveal BGLs properties
under physiological or real applications conditions. BGLs’ action on these sub-
strates may be monitored with a range of detections including HPLC, mass spec-
trometry, coupled secondary chemical or enzyme reactions for the BGL reaction
products (such as a glucose oxidase or hexokinase assay on Glc), activity-overlay
electrophoresis (Saqib and Whitney 2006), or thermochemistry (Jeoh et al. 2005).
In general, BGLs’ hydrolase activity follows the Michaelis–Menten kinetics.
With cellobiose as the substrate, kcat, Km, and kcat/Km in the order of 101–102 s−1,
100–101 mM, and 101–102 mM−1 s−1 are often observed, while with pNP-Glc as
the substrate, kcat, Km, and kcat/Km in the order of 102 s−1, 100 mM, and
102 mM−1 s−1 are often observed (Ferrara et al. 2014; Uchiyama et al. 2013;
Frutuoso and Marana 2013; Sørensen et al. 2013; Teugjas and Väljamäe; 2013;
Singhania et al. 2013; Kuntothom et al. 2009; Mendonca and Marana 2008;
Eyzaguirre et al. 2005). Significant variations (tenfold or higher) may exist for these
parameters, depending on the BGL active sites (Tsukada et al. 2008) and assay
conditions (buffer, pH, temperature, enzyme and substrate concentrations), and no
significant statistical differences are observed between the GH1 and GH3 BGLs
(p = 0.19, 0.26, 0.59, 0.07 for the kcat (cellobiose), Km (cellobiose), kcat (pNP-Glc),
Km (pNP-Glc), from two-tail t Tests) (Fig. 4.5).
BGLs’ hydrolase activity has an optimal pH range from weakly acidic to neutral,
mainly determined by the actions of the catalytic Asp and/or Glu as the nucleophile
and general acid/base, but also with contributions from other components of the
active site (Tsukada et al. 2008). Most of the BGLs (from the mesophilic organisms)
have optimal temperatures at 50–70 °C, but the BGLs from extreme thermophiles
can have optimal temperatures close to or above 100 °C (Park et al. 2005).
Many molecules may inhibit or inactivate BGLs. Nonspecific, general
protein-denaturing inhibitors include ionic surfactants, heavy metal ions such as Hg+
and Ag+, and high concentration salts (including NaCl) which exert the typical
Hofmeister (salt-out) effect (on the enzymes’ apparent pKa, not the second-order rate
constant) (Bowers et al. 2007). Redox-active metal ions such as Cu2+ and Fe3+ may
4 The Role and Applications of b-Glucosidases in Biomass … 61
1.E+04 1.E+04
kcat, s-1, on cellobiose Km, mM, on cellobiose
1.E+03 1.E+03
1.E+02 1.E+02
1.E+01 1.E+01
1.E+00 1.E+00
1.E-01 1.E-01
1.E-02 1.E-02
GH1 (29) GH3 (57) All (101) GH1 (29) GH3 (59) All (105)
1.E+04 1.E+02
Km, mM, on pNP-Glc
1.E+03
1.E+01
1.E+02
1.E+00
1.E+01
1.E-01
1.E+00
1.E-02
1.E-01
kcat, s-1, on pNP-Glc
1.E-02 1.E-03
GH1 (71) GH3 (61) All (163) GH1 (72) GH3 (70) All (175)
Fig. 4.5 Box-and-whisker plots of the kcat and Km of BGLs, including those from the GH1 and
GH3 family, for the hydrolysis of cellobiose or pNP-Glc. In the parentheses are numbers of BGLs
analyzed. Data are from Brenda enzyme database (www.brenda-enzymes.org) and other literature
sources
inhibit BGL by oxidation (Tejirian and Xu 2010; Jeng et al. 2011; Sharmila et al.
1998). Tris, a commonly used ‘biological’ buffer, may also inhibit BGL by binding
to the active site (Jeng et al. 2011). BGLs may be deactivated by various protein
adsorbents that include soil minerals (Lammirato et al. 2010), oligo/polyphenols
(Tejirian and Xu 2011) and lignin, whose effect can become significant in the
enzymatic biomass hydrolysis (Haven and Jørgensen 2013; Xu et al. 2007).
Specific BGL inhibitors include its substrate, products, and the transition-state
mimics. Because of BGL active sites’ topological features, the products (Glc,
aglycone, or shortened oligosaccharide), which may be regarded as equivalent to
the subunits of a substrate, can bind to certain subsites and impede the enzymatic
reaction. Structural analogs of the products or substrates, or even competing sub-
strates with low-reactivity (e.g., fluoroglucosides), can also inhibit BGLs. The
transition-state mimics can lock up the active site with their structural features
highly similar to that of the activated substrate, and alter the enthalpy–entropy for
the binding, hence exerting competitive inhibitions more potent than the other
62 H. Ouyang and F. Xu
1.E+04 1.E+02
Ki(Glc), mM
1.E+01
1.E+03
1.E+00
1.E+02 1.E-01
1.E-02
1.E+01 1.E-03
1.E-04
1.E+00
Ki(GlcL), mM
1.E-05
1.E-01 1.E-06
GH1 (15) GH3 (44) All (76) GH1 (12) GH3 (25) All (44)
Fig. 4.6 Box-and-whisker plots of the Ki of Glc and GlcL on the hydrolysis of pNP-Glc by BGLs,
including those from the GH1 and GH3 family. In the parentheses numbers of BGLs analyzed.
Data are from Brenda enzyme database (www.brenda-enzymes.org) and other literature sources
When the substrate concentration is high, BGLs (except the GH9 ones) may show
significant glycosyltransferase or transglycosylase activity to synthesize oligosac-
charides or glucosidases, opposing the hydrolase activity (Uchiyama et al. 2013).
The (trans)glycosylation activity is allowed by the reversibility of the retaining
mechanism, when the water is replaced by a carbohydrate or another acceptor with
higher activity for the deacylation of the enzymatic intermediate (donor)
(Fig. 4.3b), as in the cases of other retaining glycosidases. The (trans)glycosylation
may yield sophorosyl, laminaribiosyl, or even cellobiosyl, products via 1,2-, 1,3-, or
1,4-glucosidic bond, or various glucosides (de Giuseppe et al. 2014).
4 The Role and Applications of b-Glucosidases in Biomass … 63
BGLs may have significant side activities on non-Glc substrates (Brenda enzyme
database www.brenda.org, Uchiyama et al. 2013; Sørensen et al. 2013; Singhania
et al. 2013; Kuntothom et al. 2009; Harnipcharnchai et al. 2009; Mendonca and
Marana 2008; Tribolo et al. 2007; Kim et al. 2006; Seidle et al. 2005; Park et al.
2005; Eyzaguirre et al. 2005; Opassiri et al. 2004; Berrin et al. 2002; Hrmova et al.
2002; Vallmitjana et al. 2001; Marana et al. 2001; Perry et al. 2007). With pNP-b-D-
fucopyranoside as substrate, kcat, Km, and kcat/Km of 101–102 s−1, 100–101 mM, and
102 mM−1 s−1 are often observed for BGL’s b-fucosidase activity. With pNP-b-D-
xylopyranoside as substrate, kcat, Km, and kcat/Km of 100–101 s−1, 100 mM, and 100–
103 mM−1 s−1 are often observed for BGL’s b-xylosidase activity. With pNP-a-L-
arabinofuranoside as substrate, kcat, Km, and kcat/Km of 100–101 s−1, 100–101 mM,
and 100–102 mM−1 s−1 are often observed for BGL’s a-arabinosidase activity. With
pNP-b-D-galactopyranoside as substrate, kcat, Km, and kcat/Km of 100–101 s−1, 100–
102 mM, and 10−1–101 mM−1 s−1 are often observed for BGL’s b-galactosidase
activity. Close range of BGL’s activities on these substrates with different gly-
cones indicate that the interaction between the active subsite −1 and the substrate’s
nonreducing glycone subunits 4- and 6-OH (which are changed in the galactosyl,
fucosyl, xylosyl, and arabinosyl subunits of the non-Glc substrates) does not dom-
inate BGL’s activity. In GH1, BGLs coexist with b-fucosidase, b-xylosidase, and
b-galactosidase, as well as b-mannosidase, b-glucuronidase. In GH3, BGLs coexist
with b-xylosidase and a-arabinofuranosidase. Therefore, the side activities might
arise from divergent evolution.
4.7.4 Engineering
the composition and topology of the active site affect the substrate specificity, pH or
temperature profile, or substrate/product inhibition (Jeng et al. 2012; Khan et al.
2011; Liu et al. 2011 Chuenchor et al. 2011; Tsukada et al. 2008; Mendonca and
Marana 2008; Chuenchor et al. 2008; Vallmitjana et al. 2001).
BGL variants with improved activity, stability, or specificity have been made by
either rational design or directed evolution (Lee et al. 2012; Khan et al. 2011; Hong
et al. 2006; Hancock et al. 2005). Engineering the active site (especially the subsites
that bind the aglycone of a substrate) has been pursued for the abatement of the
product (Glc) inhibition; however, the increase in the Glc tolerance is often
accompanied by a decrease in the hydrolase activity, due to the contribution made
by the BGL-Glc interaction to both the catalysis and inhibition. Domain swap
between mesophilic and thermophilic BGLs may lead to improved thermal or other
properties for BGLs originated from mesophiles, such as the chimera constructed
from Agrobacterium tumefaciens or Cellvibrio gilvus and Thermotoga maritime
(Deog and Kiyoshi 2011; Kim et al. 2006).
Most BGLs have no distinct functional domains other than the catalytic core
(although some GH3 BGLs have a C-terminal FnIII domain). The enzymes may be
engineered to possess domains with targeted functions, such as the carbohydrate-binding
module potentially useful for the enzymatic biomass hydrolysis (Gundllapalli et al.
2007).
The transglycosylase activity of BGLs may be improved by engineering
(Frutuoso and Marana 2013; Jeng et al. 2012; Lundemo et al. 2013; Shim et al.
2012; Choi et al. 2008; Feng et al. 2005). For instance, the Trp49 and Trp262 of
A. niger GH3 BGL are found regulating the preference to the transglycosylase and
hydrolase activity, and substitutions of the two Trp result in BGL variants with
abated or enhanced relative transglycosylase activity (Seidle et al. 2005, 2006). The
regioselectivity of a transglycosylating BGL may also be modulated by site-directed
mutagenesis (Choi et al. 2008).
4.8 Application
Biomass conversions lay the basis for the bioenergy, biorefinery, and other
biomass-sustained industries. The enzymatic hydrolysis of lignocellulose is a key
step of the processes, and its development has benefited greatly from the basic
biochemical studies on the natural lignocellulose degradation (Payne et al. 2015;
Xie et al. 2014; Sweeney and Xu 2012; van den Brink and de Vries 2011; Quinlan
et al. 2010; Xu 2004, 2010a, b; McFarland et al. 2007; Teter et al. 2006). The
industrial enzymatic biomass conversion can differ significantly from the natural
process in various aspects, especially the requirement for high efficiency: the much
shortened reaction time (days instead of years), enhanced turnover (to minimize the
cost of enzymes), and close to 100 % yield (to maximize the output), need for
enzymes to tolerate high levels of end products and insoluble substrates that are
66 H. Ouyang and F. Xu
free forms include the freedom a BGL has to act close to other enzymes or away
from them, pursuing diffused cellobiose (or other oligosaccharides) or avoiding
common deactivators.
BGLs may also be applied in associated, part of supramolecular assembly forms
as in the cellulosomes of Clostridium thermocellum and most of the anaerobic
cellulolytic bacteria. The molar ratio of the BGL in biomass-active cellulosomes
may be finely adjusted by specific dockerin–cohesin pairing. Non-native or non-
cellulosomal BGLs may be incorporated into designer cellulosomes (with selected
enzyme composition and molar ratio) with fused dockerins (Kellermann and
Rentmeister 2014; Hyeon et al. 2013). Working in the same supramolecular
assembly allows BGL to cooperate with other enzymes in close vicinity.
BGLs may also be applied in cell surface display forms. A BGL can be dis-
played, alone or with other biomass-active enzymes, on the surface of a fermen-
tative but noncellulolytic yeast or bacterium. The displayed BGL and other
enzymes can hydrolyze biomass to Glc and other fermentable sugars, which are
then fermented by the yeast or bacterium in situ. In the so-called consolidated
bioprocessing process, enzyme production and biomass-to-ethanol conversion are
done in a single unit process step, by engineered microbes such as the thermophilic
Kluyveromyces marxianus displaying A. aculeatus BGL and T. reesei EG
(Kellermann and Rentmeister 2014; Yamada et al. 2013; van Zyl et al. 2007).
References
Ahn YO, Shimizu B, Sakata K, Gantulga D, Zhou Z, Bevan DR, Esen A (2010)
Scopulin-hydrolyzing b-glucosidases in the roots of Arabidopsis. Plant Cell Physiol
51:131–143
Ali BR, Zhou L, Graves FM, Freedman RB, Black GW, Gilbert HJ, Hazelwood GP (1995)
Cellulases and hemicellulases of the anaerobic fungus Piromyces constitute a multiprotein
cellulose-binding complex and encoded by multigene families. FEMS Microbiol Lett
125:15–22
Arthan D, Kittakoop P, Esen A, Svasti J (2006) Furostanol glycoside 26-O-b-glucosidase from the
leaves of Solanum torvum. Phytochemistry 67:27–33
Barleben L, Panjikar S, Ruppert M, Koepke J, Stockigt J (2007) Molecular architecture of
strictosidine glucosidase: the gateway to the biosynthesis of the monoterpenoid indole alkaloid
family. Plant Cell 19:2886–2897
Barnett CC, Berka RM, Fowler T (1991) Cloning and amplification of the gene encoding and
extracellular b-glycosidase from Trichoderma reesei: evidence for improved rates of
saccharification of cellulosic substrates. Nat Biotechnol 9:552–567
Bause E, Legler G (1980) Isolation and structure of a tryptic glycopeptide from the active site of
b-glucosidase A3 from Aspergillus wentii. Biochim Biophys Acta 626:459–465
Bayer EA, Belaich JP, Shoham Y, Lamed R (2004) The cellulosomes: multienzyme machines for
degradation of plant cell wall polysaccharides. Annu Rev Microbiol 58:521–554
Bednarek P, Pislewska-Bednarek M, Svatos A, Schneider B, Doubsky J, Mansurova M,
Humphry M, Consonni C, Panstruga R, Sanchez-Vallet A, Molina A, Schulze-Lefert P (2009)
A glucosinolate metabolism pathway in living plant cells mediates broad-spectrum antifungal
defense. Science 323:101–106
68 H. Ouyang and F. Xu
Berrin JG, McLauchlan WR, Needs P, Williamson G, Puigserver A, Kroon PA, Juge N (2002)
Functional expression of human liver cytosolic b-glucosidase in Pichia pastoris. Insights into
its role in the metabolism of dietary glucosides. Eur J Biochem 269:249–258
Bhatia Y, Mishra S, Bisaria VS (2002) Microbial b-glucosidases: cloning, properties, and
applications. Crit Rev Biotechnol 22:375–407
Bowers EM, Ragland LO, Byers LD (2007) Salt effects on b-glucosidase: pH-profile narrowing.
Biochim Biophys Acta 1774:1500–1507
Brzobohaty B, Moore I, Kristoffersen P, Bako L, Campos N, Schell J, Palme K (1993) Release of
active cytokinin by a b-glucosidase localized to the maize root meristem. Science 262:
1051–1054
Butters TD (2007) Gaucher disease. Curr Opin Chem Biol 11:412–418
Choi W, Park KM, Jun SY, Park CS, Park KH, Cha J (2008) Modulation of the regioselectivity of
a Thermotoga neapolitana b-glucosidase by site-directed mutagenesis. J Microbiol Biotechnol
18:901–907
Chuankhayan P, Hua Y, Svasti J, Sakdarat S, Sullivan PA, Ketudat Cairns JR (2005) Purification
of an isoflavonoid 7-O-bapiosyl-glucoside b-glycosidase and its substrates from Dalbergia
nigrescens Kurz. Phytochemistry 66:1880–1889
Chuankhayan P, Rimlumduan T, Svasti J, Ketudat Cairns JR (2007) Hydrolysis of soybean
isoflavonoid glycosides by Dalbergia b-glucosidases. J Agric Food Chem 55:2407–2412
Chuenchor W, Pengthaisong S, Robinson RC, Yuvaniyama J, Svasti J, Ketudat Cairns JR (2011)
The structural basis of oligosaccharide binding by rice BGlu1 beta-glucosidase. J Struct Biol
173:169–179
Chuenchor W, Pengthaisong S, Robinson RC, Yuvaniyama J, Oonanant W, Bevan DR, Esen A,
Chen CJ, Opassiri R, Svasti J, Ketudat Cairns JR (2008) Structural insights into rice BGlu1
b-glucosidase oligosaccharide hydrolysis and transglycosylation. J Mol Biol 377:1200–1215
Cicek M, Blanchard D, Bevan DR, Esen A (2000) The aglycone specificity-determining sites are
different in 2,4-dihydroxy-7-methoxy-1,4-benzoxazin-3-one (DIMBOA)-glucosidase (maize
beta-glucosidase) and dhurrinase (sorghum beta-glucosidase). J Biol Chem 275:20002–20011
Czjzek M, Cicek M, Zamboni V, Bevan DR, Henrissat B, Esen A (2000) The mechanism of
substrate (aglycone) specificity in b-glucosidases is revealed by crystal structures of mutant
maize b-glucosidase-DIMBOA, -DIMBOAGlc, and dhurrin complexes. Proc Natl Acad
Sci USA 97:13555–13560
Dan S, Marton T, Dekol M, Bravdo BA, He S, Withers SG, Shoserov O (2000) Cloning,
expression, characterization and nucleophile identification of family 3 Aspergillus niger
b-glycosidase. J Biol Chem 275:4973–4980
Deog KJ, Kiyoshi H (2011) Construction and characterization of novel chimeric b-glucosidases
with Cellvibrio gilvus (CG) and Thermotoga maritima (TM) by overlapping PCR. In: Carpi A
(ed) Progress in molecular and environmental bioengineering—from analysis and modeling to
technology applications. InTech, Rijeka, pp 633–646
de Giuseppe PO, Souza Tde A, Souza FH, Zanphorlin LM, Machado CB, Ward RJ, Jorge JA,
Furriel Rdos P, Murakami MT (2014) Structural basis for glucose tolerance in GH1
b-glucosidases. Acta Crystallogr D Biol Crystallogr 70:1631–1639
Dhamawardhana DP, Ellis BE, Carlson JE (1995) A b-glucosidase from lodgepole pine xylem
specific for the lignin precursor coniferin. Plant Physiol 107:331–339
Dietz KJ, Sauter A, Wichert K, Messdaghi D, Hartung W (2000) Extracellular b-glucosidase
activity in barley involved in the hydrolysis of ABA glucose conjugate in leaves. J Exp Bot
51:937–944
Dodd D, Kiyonari S, Mackie RI, Cann IK (2010) Functional diversity of four glycoside hydrolase
family 3 enzymes from the rumen bacterium Prevotella bryantii B14. J Bacteriol 192:
2335–2345
Escamilla-Trevino LL, Chen W, Card ML, Shih MC, Cheng CL, Poulton JE (2006) Arabidopsis
thaliana b-glucosidases BGLU45 and BGLU46 hydrolyse monolignol glucosides.
Phytochemistry 67:1651–1660
4 The Role and Applications of b-Glucosidases in Biomass … 69
Esen A (1992) Purification and partial characterization of maize (Zea mays L.) b-glucosidase.
Plant Physiol 98:174–182
Esen A (2003) b-Glucosidases. In: Whitaker JR, Voragen AGJ, Wong DWS (eds) Handbook of
food enzymology. Marcel Dekker, New York, pp 791–804
Esen A, Blanchard DJ (2000) A specific b-glucosidase-aggregating factor (BGAF) is responsible
for the b-glucosidase null phenotype in maize. Plant Physiol 122:563–572
Eyzaguirre J, Hidalgo M, Leschot A (2005) b-Glucosidases from filamentous fungi: properties,
structure, and applications. In: Yarema KJ (ed) Handbook of carbohydrate engineering. Taylor
& Francis, Boca Raton, pp 645–685
Faure D, Desair J, Keijers V, Bekri MA, Proost P, Henrissat B, Vanderleyden J (1999) Growth of
Azospirillum irakense KBC 1 on the aryl b-glycosidase salicin requires either Sal A or Sal B.
J Bacteriol 181:3003–3009
Feng HY, Drone J, Hoffmann L, Tran V, Tellier C, Rabiller C, Dion M (2005) Converting a
b-glycosidase into a b-transglycosidase by directed evolution. J Biol Chem 280:37088–37097
Fernandez P, Canada FJ, Jimenez-Barbero J, Martın-Lomas M (1995) Substrate specificity of
small-intestinal lactase: study of the steric effects and hydrogen bonds involved in
enzyme-substrate interaction. Carbohydr Res 271:31–42
Ferrara MC, Cobucci-Ponzano B, Carpentieri A, Henrissat B, Rossi M, Amoresano A, Moracci M
(2014) The identification and molecular characterization of the first archaeal bifunctional
exo-beta-glucosidase/N-acetyl-beta-glucosaminidase demonstrate that family GH116 is made
of three functionally distinct subfamilies. Biochim Biophys Acta 1840:367–377
Ferrieira AHP, Marana SR, Terra WR, Ferreira C (2001) Purification, molecular cloning, and
properties of a b-glycosidase isolated from midgut lumen of Tenebrio molitor (Coleoptera)
larvae. Insect Biochem Mol Biol 31:1065–1076
Frutuoso MA, Marana SR (2013) A single amino acid residue determines the ratio of hydrolysis to
transglycosylation catalyzed by b-glucosidases. Protein Peptide Lett 20:102–106
Gabrisko M, Janecek S (2015) Novel family GH3 b-glucosidases or b-xylosidases of unknown
function found in various animal groups, including birds and reptiles. Carbohydr Res 408:
v44–50
Geerlings A, Ibanez MML, Memelinks J, Heijden RVD, Verpoorte R (2000) Molecular cloning
and analysis of strictosidine b-D-glycosidase, an enzyme in terpenoid indole alkaloid
biosynthesis in Catharanthus roseus. J Biol Chem 275:3051–3056
Gilbert HJ, Stalbrand H, Brumer H (2008) How the walls come tumbling down: recent structural
biochemistry of plant polysaccharide degradation. Curr Opin Plant Biol 11:338–348
Gloster TM, Davies GJ (2010) Glycosidase inhibition: assessing mimicry of the transition state.
Org Biomol Chem 8:305–320
Gloster TM, Meloncelli P, Stick RV, Zechel D, Vasella A, Davies GJ (2007) Glycosidase
inhibition: an assessment of the binding of 18 putative transition-state mimics. J Am Chem Soc
129:2345–2354
Gundllapalli SB, Pretorius IS, Cordero Otero RR (2007) Effect of the cellulose-binding domain on
the catalytic activity of a b-glucosidase from Saccharomycopsis fibuligera. J Ind Microbiol
Biotechnol 34:413–421
Hancock SM, Corbett K, Fordham-Skelton AP, Gatehouse JA, Davis BG (2005) Developing
promiscuous glycosidases for glycoside synthesis: residues W433 and E432 in Sulfolobus
solfataricus beta-glycosidase are important glucoside- and galactoside-specificity determinants.
ChemBioChem 6:866–875
Harnipcharnchai P, Champreda V, Sornlake W, Eurwilaichitr L (2009) A thermotolerant
b-glucosidase isolated from an endophytic fungi, Periconia sp. with a possible use for biomass
conversion to sugars. Prot Express Purif 67:61–69
Haven MØ, Jørgensen H (2013) Adsorption of b-glucosidases in two commercial preparations
onto pretreated biomass and lignin. Biotechnol Biofuels 6:165
Henrissat B, Callebaut I, Fabrega S, Lehn P, Mornon JP, Davies G (1995) Conserved catalytic
machinery and the prediction of a common fold for several families of glycosyl hydrolases.
Proc Natl Acad Sci USA 92:7090–7094
70 H. Ouyang and F. Xu
Marri L, Valentini S, Venditti D (1995) Cloning and nucleotide sequence of bglA from Erwinia
herbicola and expression of b-glycosidase activity in Escherichia coli. FEMS Microbiol Lett
128:135–138
Matsui I, Sakai Y, Matsui E, Kikuchi M, Kawarabayasi Y, Honda K (2000) Novel substrate
specificity of a membrane-bound b-glycosidase from the hyperthermophilic archeon
Pyrococcus horikoshii. FEBS Lett 467:195–200
Mattiacci L, Dicke M, Posthumus MA (1995) Beta-glucosidase: an elicitor of herbivore-induced
plant odor that attracts hostsearching parasitic wasps. Proc Natl Acad Sci USA 92:2036–2040
McFarland KC, Ding H, Teter S, Vlasenko E, Xu F, Cherry J (2007) Development of improved
cellulase mixtures in a single production organism. In: Eggleston G, Vercellotti JR
(eds) Industrial application of enzymes on carbohydrate based materials. American Chemical
Society & Oxford University Press, New York, pp 19–31
Mendonca LMF, Marana SR (2008) The role in the substrate specificity and catalysis of residues
forming the substrate aglycone-binding site of a b-glycosidase. FEBS J 275:2536–2547
Morant AV, Bjarnholt N, Kragh ME, Kjaergaard CH, Jørgensen K, Paquette SM, Piotrowski M,
Imberty A, Olsen CE, Møller BL, Bak S (2008a) The beta-glucosidases responsible for
bioactivation of hydroxynitrile glucosides in Lotus japonicus. Plant Physiol 147:1072–1091
Morant AV, Jørgensen K, Jørgensen C, Paquette SM, Sanchez-Perez R, Møller BL, Bak S (2008b)
b-Glucosidases as detonators of plant chemical defense. Phytochemistry 69:1795–1813
Nagano AJ, Matsushima R, Hara-Nishimura I (2005) Activation of an ER-body-localized
b-glucosidase via a cytosolic binding partner in damaged tissues of Arabidopsis thaliana. Plant
Cell Physiol 46:1140–1148
Namchuck NM, Withers SG (1995) Mechanism of Agrobacterium b-glucosidase: kinetic analysis
of the role of noncovalent enzyme/substrate interactions. Biochemistry 34:16194–16202
Nikus J, Daniel G, Jonsson LM (2001) Subcellular localization of beta-glucosidase in rye, maize
and wheat seedlings. Plant Physiol 111:466–472
Nisius A (1988) The stroma centre in Avena plastids: an aggregation of b-glucosidase responsible
for the activation of oat-leaf saponins. Planta 173:474–481
Opassiri R, Hua Y, Wara-Aswapati O, Akiyama T, Svasti J, Esen A, Ketudat Cairns JR (2004)
b-Glucosidase, exo-b-glucanase and pyridoxine transglucosylase activities of rice BGlu1.
Biochem J 379:125–131
Opassiri R, Ketudat Cairns JR, Akiyama T, Wara-Aswapati O, Svasti J, Esen A (2003)
Characterization of a rice b-glucosidase highly expressed in flower and germinating shoot.
Plant Sci 165:627–638
Opassiri R, Pomthong B, Okoksoong T, Akiyama T, Esen A, Ketudat Cairns JR (2006) Analysis
of rice glycosyl hydrolase family 1 and expression of Os4bglu12 b-glucosidase. BMC Plant
Biol 6:33
Opassiri R, Pomthong B, Akiyama T, Nakphaichit M, Onkoksoong T, Ketudat Cairns M, Ketudat
Cairns JR (2007) A stress-induced rice b-glucosidase represents a new subfamily of glycosyl
hydrolase family 5 containing a fascin-like domain. Biochem J 408:241–249
Park TH, Choi KW, Park CS, Lee SB, Kang HY, Shon KJ, Park JS, Cha J (2005) Substrate
specificity and transglycosylation catalyzed by a thermostable b-glucosidase from marine
hyperthermophile Thermotoga neapolitana. Appl Microbiol Biotechnol 69:411–422
Payne CM, Knott BC, Mayes HB, Hansson H, Himmel ME, Sandgren M, Ståhlberg J,
Beckham GT (2015) Fungal cellulases. Chem Rev 115:1308–1448
Perry JD, Morris KA, James AL, Oliver M, Gould FK (2007) Evaluation of novel chromogenic
substrates for the detection of bacterial b-glucosidase. J Appl Microbiol 102:410–415
Poulton JE (1990) Cyanogenesis in plants. Plant Physiol 94:401–405
Pozzo T, Pasten JL, Karlsson EN, Logan DT (2010) Structural and functional analyses of
beta-glucosidase 3B from Thermotoga neapolitana: a thermostable three-domain representative
of glycoside hydrolase 3. J Mol Biol 397:724–739
Qi M, Jun HS, Forsbert CW (2008) Cel9D, an atypical 1, 4-b-D-glucan glucohydrolase from
Fibrobacter succinogenes: characteristics, catalytic residues, and synergistic interactions with
other cellulases. J Bacteriol 109:1976–1984
4 The Role and Applications of b-Glucosidases in Biomass … 73
Quinlan RJ, Teter S, Xu F (2010) Towards the development of cellulases: approaches, obstacles
and outlook. In: Waldron K (ed) Bioalcohol production: biochemical conversion of
lignocellulosic biomass. Woodhead Publishing, CRC Press, Boca Raton, pp 178–204
Rani V, Mohanram S, Tiwari R, Nain L, Arora A (2014) Beta-glucosidase: key enzyme in
determining efficiency of cellulase and biomass hydrolysis. J Bioprocess Biotechnol 5:197
Reuveni M, Sagi Z, Evnor D, Hetzroni A (1999) b-Glucosidase activity is involved in scent
production in Narcissus flowers. Plant Sci 147:19–24
Sanz-Aparicio J, Hermoso JA, Martinez-Ripoll M, Lequerica JL, Polaina J (1998) Crystal structure
of beta-glucosidase A from Bacillus polymyxa: insights into the catalytic activity in family 1
glycosyl hydrolases. J Mol Biol 275:491–502
Saqib AAN, Whitney PJ (2006) Esculin gel diffusion assay (EGDA): a simple and sensitive
method for screening b-glucosidases. Enzyme Microb Technol 39:182–184
Schubot FD, Kataeva IA, Chang J, Shah AK, Ljungdahl LG, Rose JP, Wang BC (2004) Structural
basis for the exocellulase activity of the cellobiohydrolase CbhA from Clostridium
thermocellum. Biochemistry 43:1163–1170
Seidle HF, Allison SJ, George E, Huber RE (2006) Trp-49 of the family 3 beta-glucosidase from
Aspergillus niger is important for its transglucosidic activity: creation of novel
beta-glucosidases with low transglucosidic efficiencies. Arch Biochem Biophys 455:110–118
Seidle HF, Huber RE (2005) Transglucosidic reactions of the Aspergillus niger family 3
b-glucosidase: qualitative and quantitative analyses and evidence that the transglucosidic rate is
independent of pH. Arch Biochem Biophys 436:254–264
Seidle HF, McKenzie K, Marten I, Shoseyov O, Huber RE (2005) Trp-262 is a key residue for the
hydrolytic and transglucosidic reactivity of the Aspergillus niger family 3 b-glucosidase:
substitution results in enzymes with mainly transglucosidic activity. Arch Biochem Biophys
444:66–75
Seshadri S, Akiyama T, Opassiri R, Kuaprasert B, Ketudat Cairns J (2009) Structural and
enzymatic characterization of Os3BGlu6, a rice b-glucosidase hydrolyzing hydrophobic
glycosides and (1 ! 3)- and (1 ! 2)-linked disaccharides. Plant Physiol 151:47–58
Sharmila T, Sreeramulu G, Nand K (1998) Purification and characterization of b-1-4,-glucosidase
from Clostridium papyrosolvens. Biotechnol Appl Biochem 27:175–179
Sherameti I, Venus Y, Drzewiecki C, Tripathi W, Dan VM, Nitz I, Varma A, Grundler F,
Oelmuller R (2008) PYK10, a b-glucosidase located in the endoplasmatic reticulum, is crucial
for the beneficial interaction between Arabidopsis thanliana and the endophytic fungus
Piriformospora indica. Plant J 54:428–439
Shim JH, Chen HM, Rich JR, Goddard-Borger ED, Withers SG (2012) Directed evolution of a
b-glycosidase from Agrobacterium sp. to enhance its glycosynthase activity toward
C3-modified donor sugars. Protein Eng Des Sel 25:465–472
Singhania RR, Patel AK, Sukumaran RK, Larroche C, Pandey A (2013) Role and significance of
beta-glucosidases in the hydrolysis of cellulose for bioethanol production. Bioresour Technol
127:500–507
Sørensen A, Lübeck M, Lübeck PS, Ahring BK (2013) Fungal beta-glucosidases: a bottleneck in
industrial use of lignocellulosic materials. Biomolecules 3:612–631
Souza FHM, Nascimento CV, Rosa JC, Masui DC, Leone FA, Jorge JA, Furriel RPM (2010)
Purification and biochemical characterization of a mycelial glucose- and xylose-stimulated
b-glucosidase from the thermophilic fungus Humicola insolens. J Process Biochem 45:
272–278
Srivastava KK, Verma PK, Srivastava R (1999) A recombinant cellulolytic Escherichia coli:
cloning of the cellulase gene and characterization of a bifunctional cellulase. Biotechnol Lett
21:293–297
Sue M, Yamazaki K, Yajima S, Nomura T, Matsukawa T, Iwamura H, Miyamoto T (2006)
Molecular and structural characterization of hexameric beta-D-glucosidases in wheat and rye.
Plant Physiol 141:1237–1247
Suzuki H, Takahasi S, Watanabe R, Fukushima Y, Fujita N, Noguchi A, Yokoyama R,
Nishitani K, Nishino T, Nakayama T (2006) An isoflavone conjugate-hydrolyzing
74 H. Ouyang and F. Xu
b-glucosidase from the roots of soybean (Glycine max) seedlings. J Biol Chem 281:30251–
30259
Sweeney MD, Xu F (2012) Biomass converting enzymes as industrial biocatalysts for fuels and
chemicals: recent developments. Catalysts 2:244–263
Takano M, Moriyama R, Ohmiya K (1992) Structure of a b-glycosidase gene from Ruminococcus
albus and properties of the translated product. J Ferment Bioeng 73:79–88
Takashima S, Nakamura A, Hidaka M, Masaki H, Uozumi T (1999) Molecular cloning and
expression of the novel fungal b-glycosidase genes from Humicola grisea and Trichoderma
reesei. J Biochem 125:728–736
Tejirian A, Xu F (2010) Inhibition of cellulase-catalyzed lignocellulosic hydrolysis by iron and
oxidative metal ions and complexes. Appl Environ Microbiol 76:7673–7682
Tejirian A, Xu F (2011) Inhibition of enzymatic cellulolysis by phenolic compounds. Enzyme
Microb Technol 48:239–247
Teter S, Xu F, Nedwin GE, Cherry JR (2006) Enzymes for biorefineries. In: Kamm B, Gruber PR,
Kamm M (eds) Biorefineries—industrial processes and products: status quo and future
directions. Wiley-VCH, Weinheim, pp 357–384
Teugjas H, Väljamäe P (2013) Selecting b-glucosidases to support cellulases in cellulose
saccharification. Biotechnol Biofuels 6:105
Thorlby G, Fourier N, Warren G (2004) The SENSITIVE TO FREEZING2 gene, required for
freezing tolerance in Arabidopsis thaliana, encodes a beta-glucosidase. Plant Cell 16:
2192–2203
Tiwari R, Singh S, Nain PK, Rana S, Sharma A, Pranaw K, Nain L (2013) Harnessing the
hydrolytic potential of phytopathogenic fungus Phoma exigua ITCC 2049 for saccharification
of lignocellulosic biomass. Bioresour Technol 150:228–234
Tribolo S, Berrin J-G, Kroon PA, Czjzek M, Juge N (2007) The structure of human cytoplasmic
b-glucosidase unravels substrate aglycone specificity of a family 1 glycoside hydrolase. J Mol
Biol 370:964–975
Trimbur DE, Warren RAJ, Withers SG (1992) Region-directed mutagenesis of residues
surroundingthe active site nucleophile in b-glucosidase from Agrobacterium faecalis. J Biol
Chem 267:10248–10251
Tsukada T, Igarashi K, Fushinobu S, Samejima M (2008) Role of subsite +1 residues in pH
dependence and catalytic activity of the glycoside hydrolase family 1 b-glucosidase BGL1A
from the Basidiomycete Phanerochaete chrysosporium. Biotechnol Bioeng 99:1295–1302
Tsukada T, Igarashi K, Yoshida M, Samejima M (2006) Molecular cloning and characterization of
two intracellular b-glucosidases belonging to glycoside hydrolase family 1 from the
basidiomycete Phanerochaete chrysosporium. Appl Microbiol Biotechnol 73:807–814
Uchiyama T, Miyazaki K, Yaoi K (2013) Characterization of a novel b-glucosidase from a
compost microbial metagenome with strong transglycosylation activity. J Biol Chem
288:18325–18334
Vallmitjana M, Ferrer-Navarro M, Planell R, Abel M, Ausín C, Querol E, Planas A, Pérez-Pons JA
(2001) Mechanism of the family 1 beta-glucosidase from Streptomyces sp.: catalytic residues
and kinetic studies. Biochemistry 40:5975–5982
van den Brink J, de Vries RP (2011) Fungal enzyme sets for plant polysaccharide degradation.
Appl Microbiol Biotechnol 91:1477–1492
van Zyl WH, Lynd LR, den Haan R, McBride JE (2007) Consolidated bioprocessing for
bioethanol production using Saccharomyces cerevisiae. Adv Biochem Eng Biotechnol
108:205–235
Varghese JN, Hrmova M, Fincher GB (1999) Three-dimensional structure of a barley b-D-glucan
exohydrolase; a family 3 glycosyl hydrolase. Structure 7:179–190
Verdoucq L, Moriniere J, Bevan DR, Esen A, Vasella A, Henrissat B, Czjzek M (2004) Structural
determinants of substrate specificity in family 1 b-glucosidases: novel insights from the crystal
structure of sorghum dhurrinase-1, a plant b-glucosidase with strict specificity, in complex with
its natural substrate. J Biol Chem 279:31796–31803
4 The Role and Applications of b-Glucosidases in Biomass … 75
Vroemen S, Heldens J, Boyd C, Henrissat B, Keen NT (1995) Cloning and characterization of the
bgxa genefrom Erwinia chrysanthemi D1 that encodes a b-glycosidase/xylosidase enzyme.
Mol Gen Genet 246:465–477
Wang Q, Graham RW, Trimbur D, Warren RAJ, Withers SG (1994) Changing enzymic reaction
mechanisms by mutagenesis: conversion of a retaining glucosidase to an inverting enzyme.
J Am Chem Soc 116:11594–11595
Wang Q, Trimbur D, Graham R, Warren RA, Withers SG (1995) Identification of the acid/base
catalyst in Agrobacterium faecalis beta-glucosidase by kinetic analysis of mutants.
Biochemistry 34:14554–14562
Wilson CA, Wood TM (1992) The anaerobic fungus Neocallimastix frontalis: isolation and
properties of a cellulosome-type enzyme fraction with the capacity to solubilize
hydrogen-bond-ordered cellulose. Appl Microbiol Biotechnol 37:125–129
Withers SG, Street IP, Bird P, Dolphin DH (1987) 2-Deoxy-2-fluoroglucosides: a novel class of
mechanism based inhibitors. J Am Chem Soc 109:7530–7531
Withers SG, Street IP (1988) Identification of a covalent a-D-glucopyranosyl enzyme intermediate
formed on a b-glucosidase. J Am Chem Soc 110:8551–8553
Withers SG, Warren RAJ, Street IP, Rupitz K, Kempton JB, Aebersold R (1990) Unequivocal
demonstration of the involvement of a glutamate residue as a nucleophile in the mechanism of
a retaining glycosidase. J Am Chem Soc 112:5887–5889
Wulff-Strobel CR, Wilson DB (1995) Cloning, sequencing, and characterization of a membrane
associated Prevotella ruminicola B1 4-b-glycosidase with cellodextrinase and cyanoglycosi-
dase activities. J Bacteriol 177:5884–5890
Xie S, Syrenne R, Sun S, Yuan JS (2014) Exploration of natural biomass utilization systems
(NBUS) for advanced biofuel—from systems biology to synthetic design. Curr Opin
Biotechnol 27:195–203
Xu F (2010a) Enzymatic degradation of lignocellulosic biomass. In: Tao A, Kazlauskas R
(eds) Biocatalysis for green chemistry and chemical process development. Wiley, Hoboken,
pp 361–390
Xu F (2010) Biomass-converting enzymes and their bioenergy applications. In: Baltz RH,
Demain AL, Davies JE (ed-in-chief) Bull AT, Junker B, Katz L, Lynd LR, Masurekar P,
Reeves CD, Zhao H (eds) The manual of industrial microbiology and biotechnology, 3rd edn.
American Society for Microbiology Press, Washington, DC, pp 495–508
Xu F (2004) Enhancing biomass conversion to fermentable sugars: a progress report of a joint
government-industrial project. In: Ohmiya K, Sakka K, Karita S, Kimura T, Sakka M,
Onishi Y (eds) Biotechnology of lignocellulose degradation and biomass utilization. Uni
Publishers, Tokyo, pp 793–804
Xu F, Ding H, Osborn D, Tejirian A, Brown K, Albano W, Sheehy N, Langston J (2007) Partition
of enzymes between the solvent and insoluble substrate during the hydrolysis of lignocellulose
by cellulases. J Mol Catal B Enz 51:42–48
Xu Z, Escamilla-Trevino LL, Zeng L, Lalgondar M, Bevan DR, Winkel BSJ, Mohamed A,
Cheng C, Shih M, Poulton JE, Esen A (2004) Functional genomic analysis of Arabidopsis
thaliana glycoside hydrolase family 1. Plant Mol Biol 55:343–367
Yamada R, Hasunuma T, Kondo A (2013) Endowing non-cellulolytic microorganisms with
cellulolytic activity aiming for consolidated bioprocessing. Biotechnol Adv 31:754–763
Zouhar J, Vevodova J, Marek J, Damborsky J, Su X-D, Bryzobohaty B (2001) Insights into the
functional architecture of the catalytic center of a maize b-glucosidase Zm-p60.1. Plant Physiol
127:973–985
Part II
Hemicellulases
Chapter 5
Role and Applications of Feruloyl
Esterases in Biomass Bioconversion
5.1 Introduction
The combustion of fossil fuels constitutes a great percentage of our current energy
supply, and with depleting crude oil reserves and climate change due to the
greenhouse gas (GHG) emissions, it is imperative that we focus on more renewable
energy sources such as biofuels (Damásio et al. 2013). Biofuels from lignocellu-
losic biomass, also referred to as second-generation biofuels, are considered to be
the key due to their potential to mitigate GHG emissions compared with conven-
tional fuels, reduce oil dependency, and eradicate public concerns about a trade-off
between foods and fuels.
Fig. 5.1 Action of FAEs for the hydrolysis of ester bonds in plant cell wall arabinoxylan, as
indicated by black arrows. However, most of the characterized FAEs are acting on soluble
feruloylated fractions created by the action of endo-β-1,4-xylanases and not directly on the sugar
polymer showing strong synergistic interaction
5 Role and Applications of Feruloyl Esterases … 81
L-arabinans (Topakas et al. 2007). Besides FAEs’ role in hydrolytic reactions, they
are also applicable for the synthesis of bioactive compounds with antioxidant and/or
antimicrobial activity, and other compounds of high industrial interest. The present
review is an attempt to elucidate the role and applications of FAEs in lignocellu-
losic biomass degradation and as biosynthetic tools for the modification of plant
derived antioxidants.
Plant cell walls, besides their role as structural support to the plant body, also
restrict enzyme access rendering this complex polysaccharide network recalcitrant
to biochemical degradation. Cellulose, hemicellulose, lignin, and pectin are the
main components of lignocellulosic biomass along with proteins and aromatic
compounds at proportions that differ among plant species (de Oliveira et al. 2014).
In plants, two major types of cell walls could be distinguished: the primary cell
walls, which are composed of cellulose, hemicelluloses, pectins, and proteins, and
the secondary cell walls, consisting of cellulose, hemicelluloses, and lignin that are
formed after the end of cell expansion (Endler and Persson 2011). Cellulose is a
linear natural biopolymer consisting of glucose units linked by β-1,4-glycosidic
bonds forming crystalline microfibrils via hydrogen bonding and van der Waals
interactions (Horn et al. 2012).
After cellulose, hemicelluloses are the second most abundant polysaccharides on
earth with xylans to be their main representative. Hemicelluloses are a group of
heterogeneous polysaccharides consisting of a β-1,4-linked backbone of pentoses,
hexoses, and sugar acids, which is usually decorated with side branches (Saha
2003). The exact structure and abundance of hemicelluloses diversify widely
among different plant species and cell types with xyloglucans occurring mostly in
the primary walls of dicots and conifers, while arabinoxylans dominate in com-
melinid monocots (Scheller and Ulvskov 2010). Through interactions with cellu-
lose, and sometimes lignin, hemicelluloses have the ability to strengthen the cell
walls (Endler and Persson 2011).
Pectins are a group of heterogeneous and branched polysaccharides that can be
found only in the primary cell wall and consist of a backbone of α-1,4-linked D-
galacturonic acid residues that can be methyl-esterified or decorated with acetyl
groups. Homogalacturonans, xylogalacturonan, and rhamnogalacturonans (I & II)
are the main groups in which pectins could be classified according to their basic
structure (Wong 2008). There is also increasing evidence that pectin interacts
covalently with hemicelluloses providing structural and functional complexity to
the plant cell wall (Caffal and Mohnen 2009).
Cell walls are also fortified by the deposition of the phenylpropanoid polymer
lignin, which is one of the main contributors in recalcitrance of cell walls to enzy-
matic saccharification (Chen and Dixon 2007). The term lignin is employed to
describe a group of aromatic and nonsoluble heteropolymers consisting mainly of
82 C. Katsimpouras et al.
as diferulates. In pectins isolated from sugar beet pulp, 8.8 % of the ferulates were
mostly 8-8- and 8-O-4-coupled dehydrodimers (Oosterveld et al. 1997).
Fry et al. (2000) suggested that trimers or even larger products contribute highly
to cross-linking between polysaccharides in culture maize cells. The first FA
dehydrotrimer was isolated from maize bran insoluble fiber and found to be a 5-5/8-
O-4-coupled dehydrotrimer (Bunzel et al. 2003). Since then, more dehydrotrimers
and dehydrotetramers have been identified and characterized (Rouaou et al. 2003;
Bunzel et al. 2005, 2006; Funk et al. 2005; Barron et al. 2007; Hemery et al. 2009).
The complex and extended polysaccharide networks in plant cell walls, due to the
covalent cross-linking mediated by FA, raise hurdles in utilizing lignocellulosic
biomass for biofuel production, as it contributes to biomass recalcitrance to enzy-
matic degradation (Wong et al. 2013). In order to surmount these hurdles, along
with cellulases and hemicellulases, accessory enzymes such as FAEs should be
employed, enhancing this way the fermentable sugars yield from lignocellulosic
biomass (Faulds 2010). FAEs action depends highly on the type of xylanases that is
being used with, having an effect not only on the amount but also in the form of FA
released, with family 11 xylanases being more efficient in the hydrolysis of FA
whereas family 10 xylanases present a more synergistic effect on the release of
diferulic acid (Faulds et al. 2006). The synergism of a recombinant FAE isolated
from a rumen microbial metagenome (RuFae2), in association with GH10 and
GH11 endoxylanases was investigated by Wong et al. (2013). The results obtained
from the release of FA from several natural substrates such as corn fiber, corn bran,
wheat bran, wheat-insoluble arabinoxylan, and switchgrass indicated that the GH10
xylanase showed a greater synergistic effect than that of the GH11 xylanase (Wong
et al. 2013). The synergistic action between xylanases and FAEs on the release of
FA from feruloylated polysaccharides seems to render the biomass more vulnerable
to glycoside hydrolase enzymes (Yu et al. 2003).
Tabka et al. (2006) reported a synergistic effect of a recombinant FAE from
Aspergillus niger when combined with cellulases and xylanases from Trichoderma
reesei in the hydrolysis of wheat straw. Significant increase in the release of reducing
sugars (34.8 % increase in total reducing sugars) from oat hulls was also a result of
synergistic interaction between A. niger FAE and T. reesei xylanase (Yu et al. 2003).
The enzymatic saccharification of pretreated corn stover by a cellobiohydrolase from
T. reesei was significantly enhanced by the addition of small quantities of an
endoxylanase, a FAE, and an acetyl xylan esterase. FAE addition to cellobiohydrolase
achieved a 37 % synergistic improvement in glucan conversion to cellobiose, while a
substantial increase of 85 % was shown when all enzymes were added (Selig et al.
2008). Supplementation of a commercial mixture of cellulase and β-glucosidase with
pectinase and FAE preparations resulted in higher arabinose and xylose yields from
84 C. Katsimpouras et al.
pretreated dried distiller’s grains with solubles, a co-product of corn ethanol pro-
duction (Dien et al. 2008). Kim et al. (2008) reported a 15 % increase in glucose yields
and 2–4 times enhancement for xylose yields in the enzymatic hydrolysis of pretreated
distiller’s grain when cellulases were supplemented with xylanase and FAE activities.
Two recombinant FAEs from Penicillium funiculosum expressed in Aspergillus
awamori (FaeA and FaeB) were able to enhance the cellulolytic activity of purified T.
reesei on pretreated corn stover by releasing 19 and 7 % more cellobiose, respectively
(Knoshaug et al. 2008). The synergistic action among cellulases, xylanases,
β-glucosidases, and FAEs produced by T. reesei and A. awamori was investigated for
the hydrolysis of steam-pretreated sugarcane bagasse, as reported by Gottschalk et al.
(2010). Addition of FAE and xylanase activities, produced by Aspergillus oryzae, in a
commercial enzyme preparation resulted in increased bioconversion of sugarcane
bagasse by 36 % (Braga et al. 2014).
Enzymatic release of FA from agroindustrial waste materials with the use of
FAEs raises considerable interest due to its antioxidant and anti-inflammatory
properties. FA is considered to alleviate oxidative stress in neurodegenerative
disorders such as Alzheimer’s disease (Sultana et al. 2005). Furthermore, FA could
be employed as a source for the bioconversion into value-added products such as
vanillin, which is one of the most used flavoring agents in the food, pharmaceutical,
and cosmetic industries (Priefert et al. 2001).
Two FAEs from Streptomyces sp. were tested on their ability to release FA from
corn bran, wheat bran, and defatted rice bran in combination with xylanase and α-L-
arabinofuranosidase activities from Streptomyces with the results suggesting that the
amount of FA released from biomass increased due to a synergistic effect between
these enzymes (Uraji et al. 2014). A recombinant FAE from Aspergillus usamii in
combination with a recombinant GH 11 xylanase from the same microorganism
exhibited a 27 % increase in FA release from destarched wheat bran compared to that
of the FAE alone (Gong et al. 2013). The amount of FA released from steam exploded
corn stalk, when a FAE from Aspergillus flavus (AfFaeA) combined with a GH 10
xylanase from Geobacillus stearothermophilus, was 13-fold higher than that released
by AfFaeA alone (Zhang et al. 2013). Wu et al. (2012) reported the characterization of
two FAEs of myxobacterial origin (Sorangium cellulosum) capable of releasing FA
from grass biomass, such as triticale bran, and from wheat bran yielding up to 85 % of
total alkali-extractable FA in presence of a Trichoderma viride xylanase.
A recombinant FAE (MtFae1a) from the thermophilic fungus Myceliophthora ther-
mophila (synomym Sporotrichum thermophile) was able to release up to 41 % of the
total alkali-extractable FA from wheat bran only with the presence of a xylanase,
indicating this way the synergistic interaction between MtFae1a and M3 Trichoderma
longibrachiatum xylanase (Topakas et al. 2012). Synergistic interaction between a
FAE (Tx-Est1) and a GH11 xylanase (Tx-Xyl11) both from Thermobacillus
xylanilyticus was also reported by Rakotoarivonina et al. (2011), where esterase’s
efficiency in releasing FA and diferulic acid from destarched wheat bran was seven-
to eightfold higher than that of Tx-Est1 alone. A maximum of 67 % of total FA was
released from destarched wheat bran by the combined action of a type C FAE from
Fusarium oxysporum and a M3 T. longibrachiatum xylanase (Moukouli et al. 2008).
5 Role and Applications of Feruloyl Esterases … 85
FAEs appear to be a very diverse set of enzymes, with little unifying sequence and
physical characteristics to link them. In order to categorize and classify these car-
bohydrate esterases into different types sharing the same substrate specificity and/or
the same structure, several attempts have been made starting in 2004. Under this
scope, the use of multiple alignments of sequences or domains demonstrating FAE
activity, as well as related sequences, helped to construct a neighborhood-joining
phylogenic tree (Crepin et al. 2004a). The outcome of this genetic comparison
supported substrate specificity data and allowed FAEs to be subclassified into 4
types: A, B, C, and D, a classification that is still used in literature by many
researchers. Thus, there does appear to be an evolutionary relationship between
FAEs, acetyl xylan esterases, tannases and certain lipases. The substrate specificity
was based on their specificity toward mono- and diferulates, on substitutions on the
phenolic ring, and on their amino acid sequence identity (Crepin et al. 2004a). An
unofficial nomenclature for describing FAEs was followed based on this classifi-
cation, following both the source of the enzyme and the type of the esterases (e.g.,
the type-A FAE produced by F. oxysporum is termed FoFaeA).
It is extremely common for esterases to act on a broad range of substrates.
Type A FAEs exhibit relatively high sequence identity with fungal lipases, however
lipase activity is not detected. For example, the Type A FAE from A. niger
(AnFaeA; Faulds et al. 1997) shows homology with lipases from Thermomyces
lanuginosus TLL (30 % sequence identity) and Rhizomucor miehei (37 % sequence
identity). This subclass show preference for the phenolic moiety of the substrate
containing methoxy substitutions, especially at meta- position(s), as occurs in fer-
ulic and sinapic acids, while type B FAEs shows complementary activity, showing
preference to substrates containing one or two hydroxyl substitutions as found in p-
coumaric or caffeic acid. Type B FAEs, such as M. thermophila MtFae1a (Topakas
et al. 2012), P. funiculosum PfFaeB (Kroon et al. 2000) and Neurospora crassa
Fae-1 (Crepin et al. 2003a), show high sequence identity with acetyl xylan esterases
and are the only type of FAEs that are members of carbohydrate esterase
(CE) family 1 of the Carbohydrate Active enZymes (CAZy) database (http://www.
cazy.org/CE1.html; Lombard et al. 2013). In contrast to type B esterases, type A
FAEs appear to prefer hydrophobic substrates with bulky substituents on the
benzene ring (Kroon et al. 1997; Topakas et al. 2005a). One of the famous Τype of
FAEs, due to their synthetic potential, is type C that exhibit high sequence identity
with fungal tannases. Together with type D FAEs, these esterases exhibit broad
specificity against synthetic hydroxycinnamic acids (ferulic, p-coumaric, caffeic and
sinapic acid) showing difference only in the ability to release 5-5′ diFA (Crepin
et al. 2004a, b; Vafiadi et al. 2006b). Lastly, Type D FAEs have been found only in
bacteria, such as XYLD from Pseudomonas fluorescens (reclassified as Cellvibrio
japonicus) (Ferreira et al. 1993) that show sequence similarity with xylanases
(Crepin et al. 2004a). These four different types of FAEs have different evolutionary
origin, as found by phylogenetic analysis resulting in five major clades I to V
86 C. Katsimpouras et al.
profile of Type B FAEs with weak or no activity against methyl sinapate. Such
discrepancy is well documented in literature with typical examples of Type C FAEs
with B type mode of action, such as AnFaeB (Kroon and Williamson 1996), AoFaeΒ
(Suzuki et al. 2014) and FoFaeC (Moukouli et al. 2008) from Aspergillus and
Fusarium species. Based on sequence homology, the structure determination of
AoFaeB esterase underpins the elucidation of the first structure of Type C FAEs.
These Type B and C FAEs are also members of the tannase family of ESTHER
database, where most of them do not exhibit tannase activity, however, Tan410
discovered from a soil metagenomic library exhibits both FAE and tannase activities
(Yao et al. 2013). The different Types of fungal and bacterial FAEs that have been
structural determined to date, is shown in Fig. 5.2.
The aforementioned FAEs have a common α/β-hydrolase fold that is
well-known in literature for the diverse activities of the enzymes that belong to this
structural superfamily, such as hydrolases, lipases, dehalogenases and peroxidases
(Holmquist 2000). In addition, these enzymes also share a conserved catalytic triad
(Ser–His–Asp) composed of a nucleophile (Ser, Cys or Asp), a conserved histidine
Fig. 5.2 The structures of fungal FAEs, a AnFaeA from A. niger (Hermoso et al. 2004), b AoFaeB
from A. oryzae (Suzuki et al. 2014) and bacterial FAEs c EstE1 from B. proteoclasticus (Goldstone
et al. 2010), d LJ0536 from L. johnsonii (Lai et al. 2011), domains, e XynY (Prates et al. 2001) and
f XynZ (Schubot et al. 2001) of C. thermocellum cellulosome complex. The catalytic α/β hydrolase
domain (green) and the lid domain (magenta) are shown. GlcNAc residues of N-glucans and the
catalytic triad are shown as cyan and yellow sticks, respectively. A calcium ion is represented as an
orange sphere. The 3D structures were prepared using PyMOL (http://pymol.org)
88 C. Katsimpouras et al.
and an acidic residue. These residues are sequenced in the order of nucleophile,
acidic and histidine residues, with the nucleophile to be located between the fifth
β-strand and the following helix of the nucleophile elbow (Nardini and Dijkstra
1999). The tetrahedral oxyanion intermediate formed during catalysis is stabilized
by the oxyanion hole that is formed by the contribution of the nucleophilic elbow.
The catalytic triad of the Type A FAE AnFaeA and the active-site cavity is confined
by a lid and a loop that confers plasticity to the substrate binding site in analogy
with lipases (Fig. 5.2). What is surprising is the fact that FAE’s lid exhibits a high
ratio of polar residues keeping it in an open conformation that gives the typical
preference of catalyzing the hydrolysis of hydrophilic substrates compared
to lipases. In addition, the open conformation is further stabilized by a
N-glycosylation site (Hermoso et al. 2004). These subtle aminoacid and confor-
mational changes are a result of an evolutionary divergence within the FAE family,
as further discussed by Levasseur et al. (2006). A functional shift followed by a
duplication event was suggested, which was occurred within the ancestral lipase
genes. Such a “neofunctionalization” was a result of positive selection during the
functional shift, possibly due to the drastic environmental changes happened by the
colonization of land by terrestrial plants, giving the selective advantage to
Euascomycetes (Aspergilli), as hypothesized by the authors (Levasseur et al. 2006).
The Type A FAE from A. niger together with Est1E from B. proteoclasticus exhibit
a small lid domain (16 and 46 amino acid residues, respectively) compared with the
AoFaeB (159 amino acid residues), as shown by the first solved FAE structure of
tannase family (Suzuki et al. 2014). In addition, the lid domain of BpEst1E has a
unique fold forming a flexible β-sheet structure around a small hydrophobic core,
while on the other hand, FAE domains of C. thermocellum do not possess a lid
domain (Fig. 5.2). This newly discovered lid underpins the continuing diversity of
insertions that decorate the common α/β fold of hydrolases that point toward a
dynamic mechanism for binding and release of substrates (Goldstone et al. 2010).
A. nidulans (S. cerevisiae) 2 % (w/v) V = 500 mL, 30 °C, 20 h, MFA – 14,900 Shin and Chen (2007)
Galactose 250 rpm
A. nidulans (P. pastoris) Methanol 30 °C, 3 days MFA – 21,700 Debeire et al. (2012),
Bauer et al. (2006)
A. niger NRRL3 1 % CB V = 100 mL, 30 °C, MFA/CB 13.9/0 – Shin and Chen 2006
5 days
A. niger CFR 1105 WB 30 °C, 96 h/V = 100 mL, 4NPF – 12,800/11700a Hegde and Muralikrishna
30 °C, 105 rpma (2009)
A. niger (S. cerevisae 2 % Glycerol V = 1 L, 30 °C 4NPF – 8200 Wong et al. (2011)
DY150)
A. oryzae RIB40 Koseki et al. (2009)
(P. pastoris GS115)
(continued)
89
Table 5.1 (continued)
90
1-thiogalactopyranoside, MFA methyl ferulate, MpCA methyl p-coumarate, EFA ethyl ferulate, MCA methyl caffeate, 4NPF 4-nitrophenyl ferulate, pNFA
p-nitrophenyl ferulate, pNPA p-nitrophenyl acetate, pNPB p-nitrophenyl butyrate, CNPF 2-chloro-4-nitrophenyl ferulate, SNL standard nutrition liquid
a
ssf/smf
b
BSG/WB
c
FAE activity was expressed on a dry weight basis
93
94 C. Katsimpouras et al.
During the last decade, FAEs have gained increased interest acquiring considerable
role in biotechnological processes. Except for their role in biofuel industry as
accessory enzymes for the degradation of lignocellulose, FAEs’ potential applica-
tions cover a broad spectrum, including the pulp and paper industry as bleaching
agents, the cosmetic and pharmaceutical industry for the synthesis of bioactive
compounds with antioxidant and/or antimicrobial activity or for the production of
flavor and fragrance precursors, the food industry as food additives or as feed
additives in animal feeds (Fig. 5.3).
The utilization of the FAEs’ hydrolytic activity lies on the release of FA from
hemicellulose present in plant cells walls. In ruminal digestion, FAEs are important
as they deesterify dietary fibers releasing hydroxycinnamates but are also take part
in colonic fermentation where their activities in gut ruminal microorganisms
content and give high yields (Zeuner et al. 2011). On the contrary, FAEs exhibit
higher substrate specificity constituting them a potential biosynthetic tool with high
efficiency.
As with lipases, a reaction medium with low or non-water content is essential in
order to boost the FAEs’ synthetic activity to the detriment of the hydrolytic one.
The first synthetic reaction catalyzed by FAE was achieved in a water-in-oil
microemulsion system for the synthesis of 1-pentyl-ferulate (Giuliani et al. 2001).
Since then, detergentless microemulsions, which are ternary systems consisted of a
hydrocarbon, a short-chained alcohol and water, have mostly been employed for the
synthesis of various alkyl and sugar esters (Table 5.3). They are low water content
media representing thermodynamically stable dispersions of aqueous microdroplets
in the hydrocarbon solvent. The spherical droplets are stabilized by alcohol
molecules adsorbed at their surface, while the enclosed in water enzyme is pro-
tected from contacting the outer organic phase and from denaturation due to a
water-rich layer (Khmelnitsky et al. 1988). An important advantage of these mix-
tures as reaction systems is that they allow the separation of reaction products and
enzyme reuse, while the solubility of relatively polar phenolic acids is high owing
to the presence of large amount of polar alcohol. In these media, transesterification
reactions using activated esters of FA as acyl donors have been proven to be faster
and more efficient compared to the direct esterifications (Olsson et al. 2011).
Another approach for efficient transesterification is the use of ionic liquid mixtures
(ILs), which comprise organic salts that remain liquid under ambient temperature
(Table 5.4). The interest in ILs as reaction media is mainly urged by their lack of
vapor pressure as they have the potential to replace volatile organic solvents
resulting to the development of greener processes. A major advantage of using ILs
is their tailorability as they exhibit increased selectivity and ability to solubilize
both polar and nonpolar substrates and products, as well as they can adjust other
physicochemical properties such as density, viscosity and solvating power through
alternating the cation and anion type to meet the needs of each system (Zeuner et al.
2011). The lack of volatility together with their chemical and thermal stability allow
simple recycling and reuse therefore could reduce significantly the cost of the
process with respect to organic-based ones. When referring to transesterification
with saccharides, the advantage of using ILs, as compared to organic solvents is the
increased substrate solubility. Furthermore higher yields in shorter reaction times
including regioselectivity may also be achieved in enzymatic acylation of saccha-
rides in ILs compared to organic solvents. Most of the research has used
second-generation ILs, which may be too expensive for commercial applications.
The availability of advanced ionic liquids- greener, inexpensive and biodegradable-
increases the likelihood that they will find commercial use in biocatalysis (Gorke
et al. 2010). Although synthesis in organic solvents has been widely studied in
lipases, their use in FAEs has been rarely reported in literature due to stability
restraints (Table 5.5). An important advantage is that they allow the implementation
of one phase reaction and the separation of reaction products, although they do not
consist a green approach for transesterification.
Table 5.3 Enzymatic synthesis of hydroxycinnamic acid esters in detergentless microemulsions reported in the literature
Hydroxycinnamate Acyl donor Enzyme Solvent system Yield (h) T (°C) References
Synthesis of aliphatic
esters
Methyl ferulate 1-Butanol StFaeC n-Hexane:1-butanol: MES-NaOH pH Up to 20 % 35 Topakas
6.0 (53.4:43.4:3.2 v/v/v) (120) et al. (2005a, b)
Methyl sinapate Up to 10 %
(*144)
Methyl p-coumarate Up to 8 %
(*144)
Methyl caffeate Up to 5.5 %
(*144)
Methyl ferulate 1-Butanol StFae-A n-Hexane:1-butanol: Up to 8 % 35 Topakas et al.
buffer (47.2:50.8:2.0 v/v/v) (144) (2004)
Methyl p-coumarate Up to 50 %
5 Role and Applications of Feruloyl Esterases …
(144)
Methyl caffeate Up to 25 %
(144)
Methyl p-coumarate 1-Butanol FoFae-I n-Hexane/1-butanol/MES-NaOH pH Up to 70 % 35 Topakas et al.
6.0 (47.2:50.8:2.0 v/v/v) (144) (2003a)
Methyl caffeate Up to 22 %
(144)
Methyl ferulate Up to 13 %
(144)
Methyl sinapinate Up to 1 %
(144)
(continued)
105
Table 5.3 (continued)
106
Hydroxycinnamate Acyl donor Enzyme Solvent system Yield (h) T (°C) References
p-hydroxyphenylacetic 1-Propanol FoFae-II n-Hexane/1-propanol/MES-NaOH pH 75 (224) 30 Topakas et al.
acid 6.6 (47.2:50.8:2.0 v/v/v) (2003b)
p- 70 (224)
Hydroxyl-phenylpropionic
acid
Methyl sinapate 1-Butanol AnFaeA+1 n-Hexane:1- or 2-butanol: 56 % (120) 35 Vafiadi et al.
MES-NaOH pH 6.0 (47.2: 50.8: 2.0 (2008a)
v/v/v)
78 %a (120)
Methyl ferulate AnFaeA 42 % (120)
Methyl p-coumarate 2 % (120)
Methyl sinapate 2-Butanol 9 % (120)
Methyl ferulate 1-Butanol Ultraflo L+1 n-Hexane/1-butanol/water 97 % (144) 37 Vafiadi et al.
(47.2:50.8:2.0 v/v/v) (2008b)
Ultraflo 3.6 % (144)
Methyl ferulate 1-Butanol Depol n-Hexane/1-butanol/water 87 % (144) 37 Vafiadi et al.
740L+1 (47.2:50.8:2.0 v/v/v) (2008b)
Depol 740L 2.6 % (144)
Methyl ferulate 1-Butanol Depol 670L+ n-Hexane/1-butanol/water 5 % (144) 37 Vafiadi et al.
(47.2:50.8:2.0 v/v/v) (2008b)
Synthesis of sugar esters
Methyl ferulate L-Arabinose StFaeC n-Hexane:t-butanol:piperazine- HCl 40 % (160) 35 Topakas et al.
pH 6.0 (53.4:43.4:3.2 v/v/v) (2005b)
Methyl ferulate Up to 50 %a Vafiadi et al.
(120) (2005)
(continued)
C. Katsimpouras et al.
Table 5.3 (continued)
Hydroxycinnamate Acyl donor Enzyme Solvent system Yield (h) T (°C) References
Ethyl ferulate 6.3 % (–)
n-Propyl ferulate 3.8 % (–)
n-Butyl ferulate 3.4 % (–)
Isopropyl ferulate 3.3 % (–)
2-Butyl ferulate 2.7 % (–)
Isobutyl ferulate 4.2 % (–)
Methyl ferulate D-Arabinose 45 % (–)
D-Glucose n.q. (120) Vafiadi et al.
(2007a)
D-Xylose n.q. (120) Vafiadi et al.
(2005)
D-Mannose n.q. (120)
5 Role and Applications of Feruloyl Esterases …
Hydroxycinnamate Acyl donor Enzyme Solvent system Yield (h) T (°C) References
L-Arabinotetraose n-Hexane:2-methyl-2-propanol: n.q. (–)
piperazine HCl pH 6.0 (47.2:50.8:2.0
v/v/v)
Methyl ferulate L-Arabinopentaose 24 %a (96)
Ethyl ferulate L-Arabinohexaose 3 % (96)
Propyl ferulate L-Arabinose 4 % (96) 35 Vafiadi et al.
(2006a)
Methyl ferulate L-Arabinose TsFaeC n-Hexane: t-butanol: MOPS-NaOH 21.2 %a (96) 40 Vafiadi et al.
pH 6.0 (53.4:43.4:3.2 v/v/v) (2006b)
Ethyl ferulate 1.4 % (96)
Ferulic acid D-Arabinose Multifect n-Hexane/1-butanol or 11.3– 35 Couto et al.
P 3000 2-butanone/MES-NaOH pH 6.0 36.7 % (2010)
(51:46:3 v/v/v) (144)
D-Galactose 5.6–8.6 %
(144)
D-Xylose 12.1–
30.8 %
(144)
Ferulic acid D-Arabinose Ceremix n-Hexane/1-butanol or 15.3– 35 Couto et al.
2-butanone/MES-NaOH pH 6.0 32.5 % (2010)
(51:46:3 v/v/v) (144)
Ferulic acid D-Galactose RP-1 n-Hexane/1-butanol or 10.2– 35 Couto et al.
2-butanone/MES-NaOH pH 6.0 19.8 % (2010)
(51:46:3 v/v/v) (144)
D-Xylose 4.4–16.3 %
(144)
(continued)
C. Katsimpouras et al.
Table 5.3 (continued)
Hydroxycinnamate Acyl donor Enzyme Solvent system Yield (h) T (°C) References
Ferulic acid D-Arabinose Depol 670 n-Hexane/1-butanol or 7.2–17.7 % 35 Couto et al.
2-butanone/MES-NaOH pH 6.0 (144) (2010)
(51:46:3 v/v/v)
D-Galactose 3.4–15.8 %
(144)
D-Xylose 20.9–
26.5 %
(144)
Ferulic acid D-Arabinose Flavourzyme n-Hexane/1-butanol or 21.9– 35 Couto et al.
2-butanone/MES-NaOH pH 6.0 36.7 % (2010)
(51:46:3 v/v/v) (144)
D-Galactose 36.2–
41.9 %
(144)
5 Role and Applications of Feruloyl Esterases …
D-Xylose 20.1–
21.7 %
(144)
Ferulic acid Raffinose Depol 740L n-Hexane/2-butanone/MES-NaOH pH 11.9 %a 35 Couto et al.
6.0 (51:46:3 v/v/v) (168) (2011)
Xylobiose 9.4 % (144)
Arabinose 7.9 % (144)
Galactobiose 5.4 % (144)
(continued)
109
Table 5.3 (continued)
110
Hydroxycinnamate Acyl donor Enzyme Solvent system Yield (h) T (°C) References
Sucrose 13.2 % (–)
Lactose 4.4 % (–)
XOS 2.8 % (–)
FOS n-Hexane/1,4-dioxane/MES-NaOH 9.6 % (–)
pH 6.0 (51:46:3 v/v/v)
Raffinose 3.1 % (144)
Xylobiose 4.2 % (144)
Galactobiose 26.8 %
(144)
a
After optimization of reaction conditions; +1: Immobilized with CLEAs methodology; n.q.: detected but not quantified; StFaeC: FAE type C from S.
thermophile ATCC 34628; StFaeA: FAE type B from S. thermophile ATCC 34628; FoFae-I/FoFae-II: FAE from F. oxysporum F3; AnFaeA: FAE type A
from A. niger; Ultraflo L/ Depol 740L: multienzymatic preparation from H. insolens; Depol 670L: multienzymatic preparation from T. reesei; TsFaeC: FAE
type C from T. stipitatus; Multifect P 3000: multienzymatic preparation from Bacillus amyloliquefaciens; Ceremix: multienzymatic preparation from Bacillus
spp.; RP-1: multienzymatic preparation from Bacillus subtilis; Depol 670L: multienzymatic preparation from T. reesei; Flavourzyme: multienzymatic
preparation from A. oryzae; Depol 740L: multienzymatic preparation from H. insolens. Systems with conversion <1 % were not included
C. Katsimpouras et al.
5 Role and Applications of Feruloyl Esterases … 111
Table 5.4 Enzymatic synthesis of hydroxycinnamic acid esters in ILs reported in the literature
Hydroxycinnamate Acceptor Enzyme Ionic liquid Yield (h) T (°C) References
system
Sinapic acid Glycerol AnFaeA [C2OHmim] 72.5%a (24) 50 Vafiadi et al.
[PF6] (85%), (2009)
MOPS-NaOH
pH 6.0 (15%)
[C5OHmim] 76.5%a (24)
[PF6] (85%),
MOPS-NaOH
pH 6.0 (15%)
Methyl sinapate AnFaeA+1 [ΒΜΙm][PF6] 13% (0.5) 40 Zeuner et al.
AnFaeA (85%), MOPS (2011)
pH 6.0 (15%)
[C2OHMIm] 21% (0.5)
[BF4] (85%),
MOPS pH 6.0
(15%)
[C5OHmim] 72.5%a (24) 50 Vafiadi et al.
[PF6] (85%), (2009)
MOPS-NaOH
pH 6.0 (15%)
[C2OHmim] Up to 2.5%
[PF6] (94 %), (24)
MOPS-NaOH
pH 6.0 (6%)
[C5OHmim] Up to 7%
[PF6] (94%), (24)
MOPS-NaOH
pH 6.0 (6%)
Sinapic acid Glycerol AndFaeC [ΒΜΙm][PF6] 1.1% (0.5) 40 Zeuner et al.
(85%), MOPS (2011)
pH 6.0 (15%)
Sinapic acid Glycerol Ultraflo L [ΒΜΙm][PF6] 1% (0.5) 40 Zeuner et al.
(85%), MOPS (2011)
pH 6.0 (15%)
Ferulic acid Raffinose Depol [ΒΜΙm][PF6] 2% (144) 35 Couto et al.
740L (96%), (2011)
MES-NaOH
pH 6.0 (3%)
D-xylose 4.2% (144)
D- 2.8% (144)
Arabinose
D- 1% (144)
Galactose
a
After optimization of reaction conditions; +1: Immobilized with CLEAs methodology; AnFaeA: FAE type
A from A. niger; AnFaeC: FAE type C from Aspergillus nidulans; Ultraflo L/ Depol 740L: multienzymatic
preparation from H. insolens. Systems with conversion <1 % were not included
Table 5.5 Enzymatic synthesis of hydroxycinnamic acid esters in organic solvents reported in the literature
112
After optimization of reaction conditions; +2: Immobilized in mesoporous silica MPS-9D; n.q.: detected but not quantified; FAE-PL: FAE from A. niger (purified from
the commercial preparation Pectinase PL “Amano”); Depol 740L: multienzymatic preparation from H. insolens. Systems with conversion <1 % were not included
5 Role and Applications of Feruloyl Esterases … 113
5.7 Conclusions
A clean technology for the sustainable production of biofuels and fine chemicals is
anticipated, utilizing the agricultural and agroindustrial residual biomass. The use of
enzymes is essential for the production and bioconversion of bioactive compounds
from waste residues resulting in a competitive price of biofuels through the
biorefinery concept. FAEs will have a profound role in utilizing waste biomass for
the production of energy and high-added value compounds. Since their discovery in
1990s, novel FAEs have been isolated and characterized showing different varia-
tions in physical characteristics that result in a diverse specificity for the hydrolysis
of model and natural substrates. In addition, many attempts have been made to
elucidate the functional relationships between sequence-diverse enzymes in respect
to their corresponding structure; however, the limited number of crystal structures
available complicates this mission. In this review, FAEs are shown to be involved
not only in the degradation of residual plant biomass, but also as tools for the
synthesis of novel bioactive components for use in health and cosmeceutical
industries. The improvements in molecular biology and process engineering permit
the overproduction of FAEs in different expression hosts, opening the way for their
commercial production that is prerequisite for many industrial applications.
References
Abokitse K, Wu M, Bergeron H, Grosse S, Lau PCK (2010) Thermostable feruloyl esterase for the
bioproduction of ferulic acid from triticale bran. Appl Microbiol Biotechnol 87:195–203
Barbe C, Dubourdieu D (1998) Characterisation and purification of a cinnamate esterase from
Aspergillus niger industrial pectinase preparation. J Sci Food Agric 78:471–478
Barr KA, Hopkins SA, Sreekrishna K (1992) Protocol for efficient secretion of HAS developed
from Pichia pastoris. Pharm Eng 12:48–51
Barron C, Surget A, Rouau X (2007) Relative amounts of tissues in mature wheat (Triticum aestivum
L.) grain and their carbohydrate and phenolic acid composition. J Cereal Sci 45:88–96
Bartolome B, Faulds CB, Kroon PA, Waldron K, Gilbert HJ, Hazlewood G, Williamson G (1997a)
An Aspergillus niger esterase (ferulic acid esterase III) and a recombinant Pseudomonas
fluorescens subsp. cellulosa esterase (XylD) released a 5-5′ dihydrodimer (diferulic acid) from
barley and wheat cell walls. Appl Environ Microbiol 63:208–212
Bartolome B, Faulds CB, Williamson G (1997b) Enzymic release of ferulic acid from barley spent
grain. J Cereal Sci 25:285–288
Bauer S, Vasu P, Persson S, Mort AJ, Somerville CR (2006) Development and application of a
suite of polysaccharide-degrading enzymes for analyzing plant cell walls. Proc Natl Acad
Sci USA 103:11417–11422
Benoit I, Navarro D, Marnet N, Rakotomanomana N, Lesage-Meessen L, Sigoillot JC, Asther M
(2006) Feruloyl esterases as a tool for the release of phenolic compounds from agro-industrial
by-products. Carbohydr Res 314:1820–1827
5 Role and Applications of Feruloyl Esterases … 115
Benoit I, Danchin EGJ, Bleinchodt RJ, de Vries RP (2008) Biotechnological applications and
potential fungal feruloyl esterases based on prevalence, classification and biochemical
diversity. Biotechnol Lett 30:387–396
Blum DL, Kataeva IA, Li X-L, Ljungdahl LG (2000) Feruloyl esterase activity of the Clostridium
thermocellum cellulosome can be attributed to previously unknown domains of XynY and
XynZ. J Bacteriol 182:1346–1351
Borneman WS, Ljungdahl LG, Hartley RD, Akin DE (1992) Purification and partial character-
ization of two feruloyl esterases from the anaerobic fungus Neocallimastix strain MC-2. Appl
Environ Microbiol 58:3762–3766
Borneman WS, Ljungdahl LG, Hartley RD, Akin DE (1993) Feruloyl and p-coumaroyl esterases
from the anaerobic fungus Neocallimastix strain MC-2: properties and functions in plant cell
wall degradation. In: Coughland MP, Hazlewood GP (eds) Hemicellulose and hemicellulases.
Portland Press, London and Chapel Hill, pp 85–102
Braga CMP, da Silva Delabona P, da Silva Lima DJ, Paixão DAA, da Cruz Pradella JG,
Farinas CS (2014) Addition of feruloyl esterase and xylanase produced on-site improves
sugarcane bagasse hydrolysis. Biores Technol 170:316–324
Bunzel M, Ralph J, Funk C, Steinhart H (2003) Isolation and identification of a ferulic acid
dehydrotrimer from saponified maize bran insoluble fiber. Eur Food Res Technol 217:128–133
Bunzel M, Ralph J, Funk C, Steinhart H (2005) Structural elucidation of new ferulic
acid-containing phenolic dimers and trimmers isolated from maize bran. Tetrahedron Lett
46:5845–5850
Bunzel M, Ralph J, Brüning P, Steinhart H (2006) Structural identification of dehydrotriferulic and
dehydrotetraferulic acids isolated from insoluble maize bran fiber. J Agric Food Chem
54:6408–6418
Butt MS, Tahir-Nadeem M, Ahmad Z, Sultan MT (2008) Xylanases and their applications in
baking industry. Food Technol Biotechnol 46:22–31
Caffal KH, Mohnen D (2009) The structure, function, and biosynthesis of plant cell wall pectic
polysaccharides. Carbohydr Res 344:1879–1900
Castanares A, McCrae SI, Wood TM (1992) Purification and properties of a feruloyl/ρ -coumaroyl
esterase from the fungus Penicillium pinophilum. Enzyme Microb Technol 14:875–884
Chen F, Dixon RA (2007) Lignin modification improves fermentable sugar yields for biofuel
production. Nat Biotechnol 25:759–761
Colquhoun IJ, Ralet MC, Thibault JF, Faulds CB, Williamson G (1994) Structure identification of
feruloylated oligosaccharides from sugar-beet pulp by NMR spectroscopy. Carbohydr Res
263:243–256
Couto J, Karboune S, Mathew R (2010) Regioselective synthesis of feruloylated glycosides using
the feruloyl esterase expressed in selected commercial multi-enzymatic preparations as
biocatalysts. Biocatal Biotransform 28:235–244
Couto J, St-Louis R, Karboune S (2011) Optimization of feruloyl esterase-catalyzed synthesis of
feruloylated oligosaccharides by response face methodology. J Mol Catal B Enzym 73:53–62
Couturier M, Haon M, Coutinho PM, Henrissat B, Lesage-Meesen L, Berrin JG (2011) Podospora
anserina hemicellulases potentiate the Trichoderma reesei secretome for saccharification of
lignocellulosic biomass. Appl Env Microbiol 77:237–246
Crepin VF, Faulds CB, Connerton IF (2003a) A non-modular type B feruloyl esterase from
Neurospora crassa exhibits concentration-dependent substrate inhibition. Biochem J 370:
417–427
Crepin VF, Faulds CB, Connerton IF (2003b) Production and characterization of the Talaromyces
stipitatus feruloyl esterase FAEC in Pichia pastoris: identification of the nucleophilic serine.
Protein Expr Purif 29:176–184
Crepin VF, Faulds CB, Connerton IF (2004a) Functional classification of the microbial feruloyl
esterases. Appl Microbiol Biotechnol 63:647–652
Crepin VF, Faulds CB, Connerton IF (2004b) Identification of a type-D feruloyl esterase from
Neurospora crassa. Appl Microbiol Biotechnol 63:567–570
116 C. Katsimpouras et al.
Damásio ARL, Braga CMP, Brenelli LB, Citadini AP, Mandelli F, Cota J, de Almeida RF,
Salvador VH, Paixao DAA, Segato F, Mercadante AZ, de Oliveira Neto M, do Santos WD,
Squina FM (2013) Biomass-to-bio-products application of feruloyl esterase from Aspergillus
clavatus. Appl Microbiol Biotechnol 97:6759–6767
Debeire P, Khoune P, Jeltsch JM, Phalip V (2012) Product patterns of a feruloyl esterase from
Aspergillus nidulans on large feruloyl-arabino-xylo-oligosaccharides from wheat bran.
Bioresour Technol 119:425–428
de Oliveira DM, Finger-Texeira A, Mota TR, Salvador VH, Moreira-Vilar FC, Molinari HBC,
Mitchell RAC, Marchiosi R, Ferrarese-Filho O, dos Santos WD (2014) Ferulic acid:
a key component in grass lignocellulose recalcitrance to hydrolysis. Plant Biotechnol J 13:
1224–1232
de Vries RP, Visser J (1999) Regulation of the feruloyl esterase (fae A) gene from Aspergillus
niger. Appl Environ Microbiol 65:5500–5503
de Vries RP, Michelsen B, Poulsen CH, Kroon PA, van den Heuvel RH, Faulds CB,
Williamson G, van den Hombergh JP, Visser J (1997) The faeA genes from Aspergillus niger
and Aspergillus tubingensis encode ferulic acid esterases involved in degradation of complex
cell wall polysaccharides. Appl Environ Microbiol 63:4638–4644
de Vries RP, van Kuyk PA, Kester HCM, Visser J (2002) The Aspergillus niger faeB gene
encodes a second feruloyl esterase involved in pectin and xylan degradation and is specifically
induced in the presence of aromatic compounds. Biochem J 363:377–386
Dien BS, Ximenes EA, O’Bryan PJ, Moniruzzaman M, Li X-L, Balan V, Dale B, Cotta MA
(2008) Enzyme characterization for hydrolysis of AFEX and liquid hot-water pretreated
distiller’s grains and their conversion to ethanol. Biores Technol 99:5216–5225
Donaghy JA, Bronnenmeier K, Soto-Kelly PF, McKay AM (2000) Purification and character-
ization of an extracellular feruloyl esterase from the thermophilic anaerobe Clostridium
stercorarium. J Appl Microbiol 88:458–466
Endler A, Persson S (2011) Cellulose synthases and synthesis in Arabidopsis. Mol Plant 4:
199–211
Esteban-Torres M, Reverón I, Mancheño JM, de las Rivas B, Muñoz R (2013) Characterization of
a feruloyl esterase from Lactobacillus plantarum. Appl Environ Microbiol 79:5130–5136
Faulds CB (2010) What can feruloyl esterases do for us? Phytochem Rev 9:121–132
Faulds CB, Williamson G (1991) The purification and characterization of
4-hydroxy-3-methoxycinnamic (ferulic) acid esterase from Streptomyces olivochromogenes.
J Gen Microbiol 137:2339–2345
Faulds C, Williamson G (1993) Ferulic acid esterase from Aspergillus niger: purification and
partial characterization of two forms from a commercial source of pectinase. Biotechnol Appl
Biochem 17:349–359
Faulds CB, Williamson G (1999) Review: the role of hydroxycinnamates in the plant cell wall.
J Sci Food Agric 79:393–395
Faulds CB, Mandalari G, Lo Curto RB, Bisignano G, Christakopoulos P, Waldron KW (2006)
Synergy between xylanases from glycoside hydrolase family 10 and family 11 and a feruloyl
esterase in the release of phenolic acids from cereal arabinoxylan. Appl Microbiol Biotechnol
71:622–629
Faulds CB, de Vries RP, Kroon PA, Visser J, Williamson G (1997) Influence of ferulic acid on the
production of feruloyl esterases by Aspergillus niger. FEMS Microbiol Lett 157:239–244
Fazary AE, Ju YH (2008) The large-scale use of feruloyl esterases in industry. Biotechnol Mol
Biol Rev 3:95–110
Fazary AE, Hamad HA, Lee JC, Koskei T, Lee CK, Ju YH (2010) Expression of feruloyl esterase
from Aspergillus awamori in Escherichia coli: characterization and crystal studies of the
recombinant enzyme. Int J Biol Macromol 46:440–444
Ferreira LMA, Wood TM, Williamson G, Faulds CB, Hazlewood G, Gilbert HJ (1993) A modular
esterase from Pseudomonas fluorescens subsp. cellulosa contains a non-catalytic binding
domain. Biochem J 294:349–355
5 Role and Applications of Feruloyl Esterases … 117
Fillingham IJ, Kroon PA, Williamson G, Gilbert HJ, Hazlewood GP (1999) A modular cinnamoyl
ester hydrolase from the anaerobic fungus Piromyces equi acts synergistically with xylanase
and is part of a multiprotein cellulose-binding cellulase-hemicellulase complex. Biochem J 343
Pt 1:215–224
Ford CW, Hartley RD (1990) Cyclodimers of p-coumaric and ferulic acids in the cell walls of
tropical grasses. J Sci Food Agric 50:29–43
Fry SC (1983) Feruloylated pectins from the primary cell wall: their structure and possible
functions. Planta 157:111–123
Fry SC, Willis SC, Paterson AEJ (2000) Intraprotoplasmic and wall-localised formation of
arabinoxylan bound diferulates and larger ferulate coupling-products in maize cell-suspension
cultures. Planta 211:679–692
Funk C, Ralph J, Steinhart H, Bunzel M (2005) Isolation and structural characterization of
8-O-4/8-O4- and 8-8/8-O4-coupled dehydrotriferulic acids from maize bran. Phytochemistry
66:363–371
Garcia-Conesa M-T, Crepin VF, Goldson AJ, Williamson G, Cummings NJ, Connerton IF,
Faulds CB, Kroon PA (2004) The feruloyl esterase system of Talaromyces stipitatus:
production of three discrete feruloyl esterases, including a novel enzyme, TsFaeC, with a broad
substrate specificity. J Biotechnol 108:227–241
Garleb KA, Fahey GC Jr, Lewis SM, Kerley MS, Montgomery L (1988) Chemical composition
and digestibility of fibre fractions of certain by-product feedstuffs fed to ruminants. J Anim Sci
66:2650–2662
Giuliani S, Piana C, Setti L, Hochkoeppler A, Pifferi PG, Williamson PG, Faulds CB (2001)
Synthesis of pentylferulate by a feruloyl esterase from Aspergillus niger using water-in-oil
microemulsions. Biotechnol Lett 23:325–330
Goldstone DC, Villas-Bôas SG, Till M, Kelly WJ, Attwood GT, Arcus VL (2010) Structural and
functional characterization of a promiscuous feruloyl esterase (Est1E) from the rumen
bacterium Butyrivibrio proteoclasticus. Proteins 78:1457–1469
Gong YY, Yin X, Zhang HM, Wu MC, Tang CD, Wang JQ, Pang QF (2013) Cloning, expression
of a feruloyl esterase from Aspergillus usamii E001 and its applicability in generating ferulic
acid from wheat bran. J Ind Microbiol Biotechnol 40:1433–1441
Gorke J, Srienc F, Kazlauskas R (2010) Towards advances ionic liquids. Polar enzyme-friendly
solvents for biocatalysis. Biotechnol Bioproc Eng 15:40–53
Gottschalk LMF, Oliveira RA, da Silva Bon EP (2010) Cellulases, xylanases, β-glucosidase and
ferulic acid esterase produced by Trichoderma and Aspergillus act synergistically in the
hydrolysis of sugarcane bagasse. Biochem Eng J 51:72–78
Gottschalk LMF, de Sousa Paredes R, Teixeira RSS, da Silva ASA, da Silva Bon EP (2013)
Efficient production of lignocellulolytic enzymes xylanase, β-xylosidase, ferulic acid esterase
and β-glucosidase by the mutant strain Aspergillus awamori 2B.361 U2/1. Braz J Microbiol
44:569–576
Grinna LS, Tschopp JF (1989) Size distribution and general structural features of N-linked
oligosaccharides from the methylotrophic yeast, Pichia pastoris. Yeast 5:107–115
Haase-Aschoff P, Linke D, Berger RG (2013) Detection of feruloyl- and cinnamoyl esterases from
basidiomycetes in the presence of interfering laccase. Bioresour Technol 130:231–238
Hassan S, Hugouvieux-Cotte-Pattat N (2011) Identification of two feruloyl esterases in Dickeya
dadantii 3937 and induction of the major feruloyl esterase and of pectate lyases by ferulic acid.
J Bacteriol 193:963–970
Hegde S, Muralikrishna G (2009) Isolation and partial characterization of alkaline feruloyl
esterases from Aspergillus niger CFR 1105 grown on wheat bran. World J Microbiol
Biotechnol 25:1963–1969
Hemery Y, Lullien-Pellerin V, Rouau X, Abecassis J, Samson MF, Åman P, von Reding W,
Spoerndli C, Barron C (2009) Biochemical markers: efficient tools for the assessment of wheat
grain tissue proportions in milling fractions. J Cereal Sci 49:55–64
118 C. Katsimpouras et al.
Hermoso JA, Sanz-Aparicio J, Molina R, Juge N, Gonzalez R, Faulds CB (2004) The crystal
structure of feruloyl esterase A from Aspergillus niger suggests evolutive functional
convergence in feruloyl esterase family. J Mol Biol 338:495–506
Holmquist M (2000) Alpha/beta-hydrolase fold enzymes: structures, functions and mechanisms.
Curr Protein Pept Sci 1:209–235
Horn SJ, Vaaje-Kolstad G, Westereng B, Eijink VGH (2012) Novel enzymes for the degradation
of cellulose. Biotechnol Biofuels 5:45
Howard RL, Abotsi E, Jansen van Rensburg EL, Howard S (2003) Lignocellulose biotechnology:
issues of bioconversion and enzyme production. Afr J Biotechnol 2:602–619
Lai K-K, Stogios PJ, Vu C, Xu X, Cui H, Molloy S, Savchenko A, Yakunin A, Gonzalez CF
(2011) An inserted α/β subdomain shapes the catalytic pocket of Lactobacillus johnsonii
cinnamoyl esterase. PLoS ONE 6(8):e23269
Iiyama K, Lam TBT, Stone BA (1990) Phenolic acid bridges between polysaccharides and lignin
in wheat internodes. Phytochemistry 29:733–737
Ishii T (1997) Structure and functions of feruloylated polysaccharides. Plant Sci 127:111–127
Ishii T, Hiroi T, Thomas JR (1990) Feruloylated xyloglucan and p-coumaroyl arabinoxylan
oligosaccharides from bamboo shoot cell-walls. Phytochemistry 29:1999–2003
Johnson KG, Silva MC, MacKenzie CR, Schneider H, Fontana JD (1989) Microbial degradation
of hemicellulosic materials. Appl Biochem Biotechnol 20–21:245–258
Kanauchi M, Watanabe S, Tsukada T, Atta K, Kakuta T, Koizumi T (2008) Purification and
characteristics of feruloyl esterase from Aspergillus awamori G-2 strain. J Food Sci 73:
458–463
Katsimpouras C, Christakopoulos P, Topakas E (2014) Fermentation and enzymes. In:
Varzakas T, Tzia C (eds) Food engineering handbook. CRC Press, Boca Raton, pp 495–497
Kheder F, Delaunay S, Abo-Chameh G, Paris C, Muniglia L, Girardin M (2009) Production and
biochemical characterization of a type B ferulic acid esterase from Streptomyces ambofaciens.
Can J Microbiol 55:729–738
Khmelnitsky YL, Hilhorst R, Veeger C (1988) Detergentless microemulsions as media for
enzymatic reactions. Eur J Biochem 176:265–271
Kim Y, Hendrickson R, Mosier NS, Ladisch MR, Bals B, Balan V, Dale BE (2008) Enzyme
hydrolysis and ethanol fermentation of liquid hot water and AFEX pretreated distiller’s grains
at high-solids loadings. Biores Technol 99:5206–5215
Kin KL, Lorca GL, Gonzalez CF (2009) Biochemical properties of two cinnamoyl esterases
purified from a Lactobacillus johnsonii strain isolated from stool samples of diabetes-resistant
rats. Appl Environ Microbiol 75:5018–5024
Knoshaug EP, Selig MJ, Baker JO, Decker SR, Himmel ME, Adney WS (2008) Heterologous
expression of two ferulic acid esterases from Penicillium funiculosum. Appl Biochem
Biotechnol 146:79–87
Koseki T, Furuse S, Iwano K, Matsuzawa H (1998) Purification and characterization of a
feruloylesterase from Aspergillus awamori. Biosci Biotechnol Biochem 62:2032–2034
Koseki T, Fushinobu S, Shirakawa AH, Komai M (2009) Occurrence, properties and application
of feruloyl esterases. Appl Microbiol Biotechnol 84:803–810
Koseki T, Takahashi K, Fushinobu S, Iefuji H, Iwano K, Hashizume K, Matsuzawa H (2005)
Mutational analysis of a feruloyl esterase from Aspergillus awamori involved in substrate
discrimination and pH dependence. Biochim Biophys Acta—Gen Subj 1722:200–208
Kroon PA, Williamson G (1996) Release of ferulic acid from sugar-beet pulp by using
arabinanase, arabinofuranosidase and an esterase from Aspergillus niger. Biotechnol Appl
Biochem 23:263–267
Kroon PA, Faulds CB, Brezillon C, Williamson G (1997) Methyl phenylalkanoates as substrates to
probe the active sites of esterases. Eur J Biochem 248:245–251
Kroon PA, Williamson G, Fish NM, Archer DB, Belshaw NJ (2000) A modular esterase from
Penicillium funiculosum which releases ferulic acid from plant cell walls and binds crystalline
cellulose contains a carbohydrate binding module. Eur J Biochem 267:6740–6752
5 Role and Applications of Feruloyl Esterases … 119
Kühnel S, Pouvreau L, Appeldoorn MM, Hinz SWA, Schols HA, Gruppen H (2012) The ferulic
acid esterases of Chrysosporium lucknowense C1: purification, characterization and their
potential application in biorefinery. Enzyme Microb Tech 50:77–85
Lambertz C, Garvey M, Klinger J, Heesel D, Klose H, Fischer R, Commandeur U (2014)
Challenges and advances in the heterologous expression of cellulolytic enzymes: a review.
Biotechnol Biofuels 7:135
Larsen J, Østergaard Haven M, Thirup L (2012) Inbicon makes lignocellulosic ethanol a
commercial reality. Biomass Bioenergy 46:36–45
Lenfant N, Hotelier T, Velluet E, Bourne Y, Marchot P, Chatonnet A (2013) ESTHER, the
database of the alpha/beta-hydrolase fold superfamily of proteins: tools to explore diversity of
functions. Nucleic Acids Res 41:D423–D429
Levasseur A, Gouret P, Lesage-Meessen L, Asther M, Asther M, Record E, Pontarotti P (2006)
Tracking the connection between evolutionary and functional shifts using the fungal
lipase/feruloyl esterase A family. BMC Evol Biol 6:92
Li J, Cai S, Luo Y, Dong X (2011) Three feruloyl esterases in Cellulosilyticum ruminicola H1 act
synergistically to hydrolyze esterified polysaccharides. Appl Environ Microbiol 77:6141–6147
Linke D, Matthes R, Nimtz M, Zorn H, Bunzel M, Berger RG (2013) An esterase from the
basidiomycete Pleurotus sapidus hydrolyzes feruloylated saccharides. Appl Microbiol
Biotechnol 97:7241–7251
Lombard V, Ramulu HG, Drula E, Coutinho PM, Henrissat B (2013) The carbohydrate-active
enzymes database (CAZy) in 2013. Nucl Acids Res 42:D490–D495
MacKenzie CR, Bilous D (1988) Ferulic acid esterase activity from Schizophyllum commune.
Appl Environ Microbiol 54:1170–1173
Mackenzie CR, Bilous D, Schneider H, Johnson KG (1987) Induction of cellulolytic and
xylanolytic enzyme systems in Streptomyces spp. Appl Environ Microbiol 53:2835–2839
Malinovsky FG, Fangel JU, Willats WGT (2014) The role of the cell wall in plant immunity. Front
Plant Sci 5:1–12
Mandalari G, Bisignano G, Lo Curto RB, Waldron KW, Faulds CB (2008) Production of feruloyl
esterases and xylanases by Talaromyces stipitatus and Humicola grisea var. thermoidea on
industrial food processing by-products. Biores Technol 99:5130–5133
Mathew S, Abraham TE (2005) Studies on the production of feruloyl esterase from cereal brans
and sugar cane bagasse by microbial fermentation. Enzyme Microb Technol 36:565–570
McClendon SD, Shin HD, Chen RR (2011) Novel bacterial ferulic acid esterase from Cellvibrio
japonicus and its application in ferulic acid release and xylan hydrolysis. Biotechnol Lett
33:47–54
McCrae SI, Leith KM, Gordon AH, Wood TM (1994) Xylan-degrading enzyme system produced
by the fungus Aspergillus awamori: isolation and characterization of a feruloyl esterase and a
p-coumaroyl esterase. Enzyme Microb Technol 16:826–834
Moukouli M, Topakas E, Christakopoulos P (2008) Cloning, characterization and functional
expression of an alkalitolerant type C feruloyl esterase from Fusarium oxysporum. Appl
Microbiol Biotechnol 79:245–254
Mueller-Harvey I, Hartley RD, Harris J, Curzon EH (1986) Linkage of p-coumaroyl and feruloyl
groups to cell-wall polysaccharides of barley straw. Carbohydr Res 148:71–85
Mukherjee G, Singh RK, Mitra A, Sen SK (2007) Ferulic acid esterase production by
Streptomyces sp. Biores Technol 98:211–213
Nardini M, Dijkstra BW (1999) Alpha/beta hydrolase fold enzymes: the family keeps growing.
Curr Opin Struct Biol 9:732–737
Nguyen D, Zhang X, Jiang ZH, Audet A, Paice MG, Renaud S, Tsang A (2008) Bleaching of kraft
pulp by a commercial lipase: accessory enzymes degrade hexenuronic acids. Enz Microbial
Technol 43:130–136
Nieter A, Haase-Aschoff P, Linke D, Nimtz M, Berger RG (2014) A halotolerant type A feruloyl
esterase from Pleurotus eryngii. Fungal Biol 118:348–357
Nordkvisk E, Salomonsson AC, Amar P (1984) Distribution of insoluble bound phenolic acids in
barley grain. J Sci Food Agric 35:657–661
120 C. Katsimpouras et al.
Sigoillot C, Camareto S, Vidal T, Record E, Asther M, Boada MP, Martinze MJ, Sigoillot JC,
Asther M, Colom JF, Martinez AT (2005) Comparison of different fungal enzymes from
bleaching high-quality paper pulps. J Biotechnol 115:333–343
Smith MM, Hartley RD (1983) Occurrence and nature of ferulic acid substitution of cell-wall
polysaccharides in graminaceous plants. Carbohydr Res 118:65–80
Sultana R, Ravagna A, Mohmmad-Abdul H, Calabrese V, Butterfield DA (2005) Ferulic acid ethyl
ester protects neurons against amyloid beta-peptide (1-42)-induced oxidative stress and
neurotoxicity: relationship to antioxidant activity. J Neurochem 92:749–758
Suzuki K, Hori A, Kawamoto K, Thangudu RR, Ishida T, Igarashi K, Samejima M, Yamada C,
Arakawa T, Wakagi T, Koseki T, Fushinobu S (2014) Crystal structure of a feruloyl esterase
belonging to the tannase family: a disulfide bond near a catalytic triad. Proteins 82:2857–2867
Tabka MG, Herpoël-Gimbert I, Monod F, Asther M, Sigoillot JC (2006) Enzymatic sacchari-
fication of wheat straw for bioethanol production by a combined cellulose xylanase and
feruloyl esterase treatment. Enz Microbial Technol 39:897–902
Tai ES, Hsieh PC, Sheu SC (2014) Effect of polygalacturonase and feruloyl esterase from
Aspergillus tubingensis on demucilage and quality of coffee beans. Process Biochem 49:
1274–1280
Tenkanen M, Schuseil J, Puls J, Poutanen K (1991) Production, purification and characterisation of
an esterase liberating phenolic acids from lignocellulosics. J Biotechnol 18:69–84
Thakur VV, Jain RK, Mathur RM (2012) Studies on xylanase and laccase enzymatic prebleaching
to reduce chlorine-based chemicals during CEH and ECF bleaching. Bioresources 7:
2220–2235
Thörn C, Gustafsson H, Olsson L (2011) Immobilization of feruloyl esterases in mesoporous
materials leads to improved transesterification yield. J Mol Catal B Enzym 72:57–64
Topakas E, Christakopoulos P (2004) Production and partial characterization of alkaline feruloyl
esterases by Fusarium oxysporum during submerged batch cultivation. World J Microb Biot
20:245–250
Topakas E, Stamatis H, Mastihubova M, Biely P, Kekos D, Macris BJ, Christakopoulos P (2003a)
Purification and characterization of a Fusarium oxysporum feruloyl esterase (FoFAE-I)
catalysing transesterification of phenolic acid esters. Enz Microbial Technol 33:729–737
Topakas E, Stamatis H, Biely P, Kekos D, Macris BJ, Christakopoulos P (2003b) Purification and
characterization of a feruloyl esterase from Fusarium oxysporum catalyzing esterification of
phenolic acids in ternary water-organic solvent mixtures. J Biotechnol 102:33–44
Topakas E, Kalogeris E, Kekos D, Macris BJ, Christakopoulos P (2003c) Production and partial
characterization of feruloyl esterase by Sporotrichum thermophile under solid-state fermen-
tation. Process Biochem 38:1539–1543
Topakas E, Stamatis H, Biely P, Christakopoulos P (2004) Purification and characterization of a
type B feruloyl esterase (StFae-A) from the thermophilic fungus Sporotrichum thermophile.
Appl Microbiol Biotechnol 63:686–690
Topakas E, Christakopoulos P, Faulds CB (2005a) Comparison of mesophilic and thermophilic
feruloyl esterases: characterization of their substrate specificity for methyl phenylalkanoates.
J Biotechnol 115:355–366
Topakas E, Vafiadi C, Stamatis H, Christakopoulos P (2005b) Sporotrichum thermophile type C
feruloyl esterase (StFaeC): purification, characterization and its use for phenolic acid (sugar)
ester synthesis. Enz Microbial Technol 36:729–736
Topakas E, Vafiadi C, Christakopoulos P (2007) Microbial production, characterization and
applications of feruloyl esterases. Proc Biochem 42:497–509
Topakas E, Moukouli M, Dimarogona M, Christakopoulos P (2012) Expression, characterization
and structural modelling of a feruloyl esterase from the thermophilic fungus Myceliophthora
thermophila. Appl Microbiol Biotechnol 94:399–411
Tsuchiyama M, Sakamoto T, Fujita T, Murata S, Kawasaki H (2006) Esterification of ferulic acid
with polyols using a ferulic acid esterase from Aspergillus niger. Biochim Biophys Acta
1760:1071–1079
122 C. Katsimpouras et al.
Wu M, Abokitse K, Grosse S, Leisch H, Lau PCK (2012) New feruloyl esterases to access
phenolic acids from grass biomass. Appl Biochem Biotechnol 168:129–143
Yang SQ, Tang L, Yan QJ, Zhou P, Xu HB, Jiang ZQ, Zhang P (2013) Biochemical characteristics
and gene cloning of a novel thermostable feruloyl esterase from Chaetomium sp. J Mol Catal B
Enzym 97:328–336
Yao J, Chen QL, Shen AX, Cao W, Liu YH (2013) A novel feruloyl esterase from a soil
metagenomic library with tannase activity. J Mol Catal B Enzym 95:55–61
Yu P, Maenz DD, McKinnon JJ, Racz VJ, Christensen DA (2002a) Release of ferulic acid from
oat hulls by Aspergillus ferulic acid esterase and Trichoderma xylanase. J Agric Food Chem
50:1625–1630
Yu P, McKinnon JJ, Maenz DD, Racz VJ, Christensen DA (2002b) The interactive effects of
enriched sources of Aspergillus ferulic acid esterase and Trichoderma xylanase on the
quantitative release of hydroxycinnamic acids from oat hulls. Can J Anim Sci 82:251–257
Yu P, McKinnon JJ, Maenz DD, Olkowski AA, Racz VJ, Christensen DA (2003) Enzymic release
of reducing sugars from oat hulls by cellulase, as influenced by Aspergillus ferulic acid esterase
and Trichoderma xylanase. J Agric Food Chem 51:218–223
Zeng Y, Yin X, Wu MC, Yu T, Feng F, Zhu TD, Pang QF (2014) Expression of a novel feruloyl
esterase from Aspergillus oryzae in Pichia pastoris with esterification activity. J Mol Catal B
Enzym 110:140–146
Zeuner B, Stahlberg T, van Buu ON, Kurov-Kruse AJ, Riisager A, Meyer AS (2011) Dependency
of the hydrogen bonding capacity of the solvent anion on the thermal stability of feruloyl
esterases in ionic liquid systems. Green Chem 13:1550–1557
Zhang SB, Zhai HC, Wang L, Yu GH (2013) Expression, purification and characterization of a
feruloyl esterase A from Aspergillus flavus. Prot Expr Pur 92:36–40
Zwane EN, Rose SH, Van Zyl WH, Rumbold K, Viljoen-Bloom M (2014) Overexpression of
Aspergillus tubingensis faeA in protease-deficient Aspergillus niger enables ferulic acid
production from plant material. J Ind Microbiol Biotechnol 41:1027–1034
Chapter 6
Endo-b-1,4-xylanase: An Overview
of Recent Developments
Abstract This chapter presents the main important aspects about the enzyme
endo-beta-1,4-xylanase. It starts with the biological grounds for the use of the
enzyme. Moving on, the advancements in the production process and optimisation
are substantially described, along with the microorganisms and a broad range of
molecular techniques used to obtain best performance enzyme. Finally, it delineates
the most common applications related to xylanase, such as in biobleaching, baking,
feedstock and the enzymatic role in juice and ink industries, among others.
Furthermore, the material brings updated research performed during the last few
years combined with essential established facts along the time.
6.1 Introduction
Since ancient times the mankind has made used of enzymes to produce goods as
wine, beer and bread. The gained knowledge in fields as chemistry, biology,
engineering and biotechnology have enabled more precise and efficient processes to
obtain and apply such biomolecules not only in the existing products but also to
create new ones, exploring fully the potential of a material and expanding its roles.
In this regard, plants serve us in many aspects, with their fundamental biological
and microbiological functions, and have been increasingly exploited to provide
valuable biomaterials, which are employed to boost advancements in daily life with
better and suitable products, while reducing the damage caused to the environment
thereby making our lives more comfortable. Hemicelluloses are the second most
present polysaccharide in plant cell wall and are made up of important polymers as
xylan (Schaedel et al. 2010). Xylanases are enzymes secreted by microorganisms,
mostly fungi and bacteria are used in a broad range of important manufacturing
pathways in the papermaking industry as well as in the food industry and in the
Plant cell wall contains a variable amount of different chemical compounds making
up their structures. The cell wall itself is composed of three layers. The compounds
present in a plant cell wall are cellulose and hemicellulose, pectin and lignocellu-
lose. After cellulose, hemicellulose is the most abundant natural polysaccharide in
plants and accounts to one-third of plant biomass. However, it varies according to
plant species due to environmental conditions (Caffall and Mohnen 2009; Schaedel
et al. 2010; Yin et al. 2010). Hemicellulose contributes not only to biological
processes in nature but also in our daily life, through the employment of its com-
ponents in several industrial sectors from feedstock to paper making and bread
baking (Yin et al. 2010).
Schulze (1891) first introduced the term ‘hemicellulose’ for the fractions isolated
or extracted from plant materials with diluted alkali. It is an amorphous
heteropolymer formed by aldopentoses (xylose and arabinose) and aldohexoses
(glucose, galactose, mannose), forming a short polymer chain presenting lower
molar mass, when compared to cellulose (Jacobsen 2000). It is present in hardwood
(angiosperm) as well as in softwood (gymnosperm). Hardwood hemicelluloses are
xylans, whereas softwood consists of glucomannans. The portion present in
6 Endo-b-1,4-xylanase: An Overview of Recent Developments 127
hardwood is the most important due to its higher amount. Xylan backbone is highly
branched by side chains consisting of xylose, arabinose and galucuranic acids (Saha
2003). According to the type of sugar residue hemicelluloses are framed into four
categories: xylans, secondary cell walls of hardwood and herbaceous plants;
xyloglucans, primary cell walls of higher plants (bound to cellulose); mannans,
secondary cell walls of conifers and Leguminosae (Schaedel et al. 2010).
The degradation of hemicelluloses is mostly carried out by microorganisms that
can be found either free in nature or as part of the digestive tract in higher animals.
Such degradation is accomplished with the aid of enzymatic complexes and leads to
the production of several important industrial compounds.
These xylan degrading enzymes are expressed by a wide variety of fungi (Wase
et al. 1985; Bakir et al. 2001; Li et al. 2006) and bacteria (Mamo et al. 2006), but
also actinomycetes (Ball and McCarthy 1988), and yeast (Hrmova et al. 1984).
Even if there is a great variety of xylanases and their folds, mechanisms of action,
substrate specificities, hydrolytic activities and physicochemical characteristics
being diverse, mostly attentions have been driven to two of the xylanases con-
taining carbohydrolase families from families 10 and 11 and are based on their
amino acid sequence homologies (Henrissat 1991). These enzymes catalyse the
1,4-beta-D-xylosidic linkages in xylan (Collins et al. 2005). As described in the
literature (Biely et al. 1983), the enzymes not only hydrolyse the substrate but also
catalyse various biomolecular reactions, such as transglycosylation and degradation
of oligosaccharides. The authors explained the dependence of enzymatic activity on
the substrate concentration and the preference of xylanases from different sources,
as those examined from A. niger and C. albidus, for cleaving glycosidic bonds.
The thermophilic fungus Thermomyces lanuginosus, first isolated in 1899 by
Tsiklinskaya from a potato, is referred as the most suitable microorganism for the
cellulose-free xylanases production both in shake flask and bioreactor cultivations;
however, due to intraspecies variability the enzymatic expression may be affected.
According to extensive studies on the molecular, physiological and ecological
properties of T. lanuginosus strains have been carried out (Singh et al. 2003).
Furthermore, the cultivation conditions of each strain require optimisation since the
strains are influenced by nutrient and growth conditions used in their cultivation. In
addition, Aspergillus niger, Trichoderma reesei and Streptomyces have been fre-
quently explored for the production of xylanases (Pradeep et al. 2013). With the
advancements of recombinant DNA technology, the gene responsible for the
enzyme started being cloned and expressed in other microorganisms, enabling
researchers to end up with purer products, presenting higher enzymatic activity and
in increased amounts. The enzyme has been cloned in bacteria, mostly Escherichia
coli, and filamentous fungi such as Trichoderma reesei and Fusarium oxysporum
have also been employed for this purpose (He et al. 2009; Moukouli et al. 2011). In
most cases, when it comes to the heterologous expression of xylanase, the host
models used are widely applied yeast Pichia pastoris and in some cases E. coli (He
et al. 2009; Cai et al. 2011; Moukouli et al. 2011).
(1996) revealed aspects of the enzyme conformation and purposed that two glu-
tamate residues would be involved in the catalysis.
De Groot et al. (1998) cloned the gene encoding xylanase from Agaricus bis-
porus and determined its regulation by heterologous screening techniques. Georis
et al. (1999) presented a recombinant xylanase from Streptomyces sp. The authors
studied the cleavage site between the signal peptide and the protein and compared
the similarities and strict identities of the enzyme with other family 11 xylanases.
By means of crystallography the group of Professor Mizuno in Japan (Fujimoto
et al. 2000) depicted the structure of the xylan-binding domain endo-beta-xylanase
family 10 from Streptomyces olivaceoviridis. Turunen et al. (2001) reported the
synergistic effect of a xylanase II from Trichoderma reesei brought about by sta-
bilising mutations to access the information about its thermal stability, catalytic
function and pH-dependent properties. Zhengqiang et al. (2001) isolated two
xylanases genes from the bacterium Thermotoga maritima. The genes were sub-
sequently cloned and expressed in E. coli and the authors found that the optimum
condition for the enzyme activity was at 90 °C and pH of 6.14. A disulphide bridge
into the N-terminal region of Trichoderma reesei endo-1,4-beta-xylanase was
engineered by Fenel et al. (2004), increasing the enzyme thermostability by about
15 °C. Subsequently Fenel et al. (2006) achieved high stability using a mutant
xylanase from Trichoderma reesei, which increased brightness in sulphate pulp
bleaching. Brutus et al. (2004) showed the cloning, purification and expression of
endo-xylanase from Penicillium funiculosum in E. coli and Pichia pastoris as hosts.
Wakiyama et al. (2008), Jun et al. (2009) and Do et al. (2013) also published studies
related to the expression of xylanase in the same host. The group of Professor
Murray contributed to the elucidation of the structure of a endo-xylanase (family
10) from Thermoascus aurantiacus, showing that the glycone subsite of the enzyme
makes extensive direct and indirect interactions with the arabinose side chain and
concluding that xylan side chains are not just accommodated but can actually
constitute significant substrate specificity determinants in the enzyme under study
(Vardakou et al. 2005).
Another study explored the function of three enzymes, namely xylosidase–ara-
binosidase from Thermoanaerobacter ethanolicus and xylanase from Thermomyces
lanuginosus, with the aim to achieve more efficiency in the decomposition of ara-
binoxylan agricultural residue for the bioconversion into fuels or industrial chemi-
cals (Xue et al. 2009). The multifunctional enzyme was expressed in E. coli and
exhibited enhanced enzymatic activity in the degradation of arabinoxylan into its
monomer constituents when compared to the corresponding free enzymes. Vieira
et al. (2009) reported the effect of temperature and binding of substrate on the
dynamics of the family 11 xylanase from Bacillus circulans and its molecular
dynamics simulations in the presence of xylobiose. The study indicates that the
thermophilicity of the 11/G xylanases might be increased by rational design based on
two variables: mutations that lead to changes in the flexibility of the thumb loop and
mutations of residues in the substrate-binding cleft that aim to optimise the xyla-
nase–substrate intermolecular interaction. With that the authors purposed that the
residues may be considered as preferential targets for site-directed mutagenesis
6 Endo-b-1,4-xylanase: An Overview of Recent Developments 131
experiments that modulate the enzyme substrate affinity, thus increasing the catalytic
efficiency of the enzyme. The work Michaux et al. (2010), based on crystallographic
techniques, with an acidophilic xylanase from Scytalidium acidophilum, a fungi
isolated from an uranium mine in acidic waters (pH 2.0), revealed features about the
stability and acidophilic adaptation. The conclusions were that various sequence and
structure modifications may be responsible for the acidophilic characteristic of the
enzyme. They would be the presence of aspartic acid hydrogen bonded to the
acid/base catalyst; the nature of specifically conserved residues in the active site; the
negative potential at the surface and the decreased number of salt bridges and
hydrogen bonds in comparison with highly alkaline enzymes.
A xylanase cloned and expressed in E. coli by Hwang et al. (2010) had its activity
not affected by most salts, such as NaCl, LiCl, KCl, NH(4)Cl, CaCl(2), MgCl(2),
MnCl(2) and CsCl(2) at 1 mM; however, the enzymatic activity was affected by
CuSO(4), ZnSO(4) and FeCl(3). As described, xylobiose was the major product
obtained. The presence of NaCL at 12.5 proved to be not only beneficial but
also the optimum condition to an endo-xylanase from thermophilic bacterium
Thermoanaerobacterium saccharolyticum isolated from a hydrothermal vent (Hung
et al. 2011). Furthermore, the expression of endo-xylanase has been described in
homologous (Anasontzis et al. 2011) and heterologous (Damásio et al. 2011) models
and applied in the production of ethanol and functional foods, respectively. In 2012
Trevizano et al., by means of direct evolution, improved the thermostability of
endo-xylanase. The mutant enzymes generated through error-prone PCR presented a
good performance in extreme pH conditions, retaining their ability to hydrolyse
birchwood and oat spelt xylans, according to the authors. Kim et al. (2012) studied
the intra-molecular hydrophobic interactions of xylanase, applying network analysis
to interpret such events, and noticed that the enzyme half-life was increased by
78-fold, highlighting the advantages of interpreting collective hydrophobic inter-
action patterns. Verma and Satyanarayana (2012) also reported a thermostable gene
from Geobacillus thermoleovorans expressed in E. coli with enzymatic production
of 27-fold higher than the wild strain. Thermostabilisation was once more investi-
gated by Li et al. (2013). Using Dictyoglomus thermophilum as an enzymatic source
of xylanase, the researchers stabilised the enzyme by an engineered N-terminal
disulphide bridge. The stabilisation was then tested against high temperatures, being
achieved at 100–110 °C and the ionic liquid used in the study appeared to affect the
functioning of the enzyme’s active site to a greater extent than the stability of an
already thermostable protein structure. Recently, Mander et al. (2014) produced a
cellulase-free xylanase from Streptomyces sp. using an agricultural residue (wheat
bran) as a growth substrate. The outcome was that the enzyme was not only able to
hydrolyse commercially available pure beechwood xylan to xylose, xylobiose and
xylotriose, but also abundantly available lignocellulosic agricultural residues in
nature such as wheat bran to xylooligosaccharides. Other studies have also
approached the thermal and alkaline stability, as well as enzyme–substrate interac-
tion of hybrid xylanases, as the ones reported by Song et al. (2014), Stephens et al.
(2014) and Chen et al. (2014).
132 A.G. Rodrigues
A specific range of techniques has been used to enable researchers and producers
transforming what nature provides us, making important and always better goods in
daily life. From molecular biology to analytical chemistry and biochemistry engi-
neering, these methods and their improvements assure better results and more
efficient processes to be conducted.
Regarding the recombinant DNA techniques used in the process of obtaining,
characterising and optimising key factors of the enzyme, polyacrylamide gel
electrophoresis (SDS–PAGE) is a widely used technique to determine molecular
mass of proteins, thus aiding the characterisation of such biomolecules (Pribowo
et al. 2012). The technique has been applied through the decades as a key tool to
identify xylanases and is part of the essential technique repertoire (Biely et al. 1983;
Coughlan and Hazlewood 1993; Takahashi et al. 2013). Polymerase chain reaction
(PCR) and gene cloning comprises the paramount employed in the analysis and
improvement of the enzyme production and analysis of its properties (Stephens
et al. 2014).
Among the analytical chemistry techniques, UV/vis spectrophotometry has been
employed to analyse the enzymatic activity of xylanases (Sonia et al. 2005).
Chromatographic methods, namely high performance liquid chromatography
(HPLC), have been used to purify the enzyme as demonstrated in a recent work
(Pribowo et al. 2012). Column chromatography is another method to purify the
enzymes from cultivation (Heck et al. 2006), as well as thin layer chromatography
(TLC) (Chanwicha et al. 2015), used in the analysis of xylanases. Along with the
above-mentioned methods, mass spectrometry finds its use in the xylanases process
as described by Jänis et al. (2007) for the determination of steady-state kinetic
parameters for a xylanase-catalysed hydrolysis of xylooligosaccharides. Puchart
and Biely (2008) carried out the production of endo-beta-1,4-xylanase using
Thermomyces lanuginosus, and the authors studied the compounds formed
employing mass spectrometry, suggesting that the fungus might be of importance
regarding the conversion of xylan into xylooligosaccharides.
6 Endo-b-1,4-xylanase: An Overview of Recent Developments 133
Xylanolytic enzymes from various sources, and most importantly from microor-
ganisms, have been studied along the last decades to understand their physical and
biochemical characteristics, due to their applications in several fields (Polizeli et al.
2005).
Xylanases from natural sources are produced only to relatively low yields;
therefore, the enzyme production cannot meet the demand of our societies. In view
of this, heterologous expression approaches have been implemented to design
xylanases with the desired characteristics for the production at the industrial scale.
The expression of fungal xylanase genes in Saccharomyces cerevisiae, Pichia
pastoris and other microorganisms has been reported in the literature (Karaoglana
et al. 2014). The enzyme properties and yield somehow vary depending on the gene
source, host system and secretion signal sequences, besides the process parameters.
Until the advent of recombinant DNA technology, enzymes were produced by
fermentation of the microorganisms that express the enzymes. Cultivation of fila-
mentous fungi for large-scale protein production is very complex, often ending up
in many interfering enzymes. Purification of target enzymes from a pool of proteins
requires several purification steps thereby increasing their costs (Kormelink et al.
1993). Recombinant DNA technology allows large-scale expression of these
enzymes in both homologous and heterologous protein expression hosts (Puchart
and Biely 2008; Juturu and Wu 2012).
Regarding the methods of obtaining endo-b-1,4-xylanases, solid-state fermen-
tation (SSF) is the most present in the literature (Filho et al. 1993; Pandey 2002;
Sonia et al. 2005; Chanwicha et al. 2015). SSF is the most suitable method for
xylanase production, since it deals with low water activity environment and allows
the use of agro waste substrates like corn and sugarcane bagasse. The method
received more attention after during 1950–1970 with the steroid transformation
using fungal cultures and the production of proteins for cattle feed enrichment. Later
on SSF would gain more importance with the design of bioreactors and the microbial
production of food, feed and enzymes such as lipase, phytase, cellulase, ligninase,
amylase, glutaminase, pectinase and xylanase, besides the production of organic
acids, secondary metabolites, fatty acids and biofuel (Pandey 2002; Bhargav et al.
2008). Compared to submerged fermentation (SmF), it facilitates the recovery of the
final product, requires less energy (no need for vigorous agitation) and has high
volumetric production, being thus eco-friendly and industrially favourable (Filho
et al. 1993; Pandey 2002; Bhargav et al. 2008). The method is applied under
microbial conditions in which cultures are closer to their natural habitat, considering
that fungi grow in nature on solid substrates such as pieces of wood, seeds, stems and
roots, therefore having their activity increased. The process can still be classified on
the grounds of substrate nature, which can be pure or mixed (Bhargav et al. 2008).
Key factors have to be satisfied if appropriate production is intended to be
achieved. Among these are moisture and water contend. This is due to the fact that
water activity of the substrate influences the microbial activity, spore production
134 A.G. Rodrigues
Apart from its use in the pulp and paper industry, xylanases are also used as food
additive, in wheat flour for improving dough handling and quality of baked
products, for the extraction of coffee, plant oils and starch, and in the improvement
of nutritional properties of animal feed, and in combination with pectinase and
cellulase for clarification of fruit juices (Biely 1985; Wong et al. 1988; Maat et al.
1992; Beg et al. 2001).
6.8 Biobleaching
Xylanases and other side-cleaving enzymes have been used in pulp bleaching
primarily to reduce lignin and increase the brightness of the pulp. This process is
necessary due to the presence of residual lignin and its derivatives in the pulping
6 Endo-b-1,4-xylanase: An Overview of Recent Developments 135
process, which causes the resultant pulp to gain a characteristic brown colour. The
use of xylanase in bleaching pulp requires the use of enzymes with special char-
acteristics (Beg et al. 2001).
A key requirement is to be cellulose-free, to avoid damaging the pulp fibres, as
cellulose is the primary product in the paper industry (Srinivasan and Rele 1999;
Subramaniyan and Prema 2000). In pulp bleaching, the enzyme must tolerate high
temperature and pH after alkaline cooking to be economical (Turunen et al. 2001).
The process of lignin removal from chemical pulps to produce bright or completely
white finished pulp is called ‘bleaching’. It is necessary for aesthetic reasons and for
improvement of paper properties, because the left-over residual lignin after sulphite
pulping imparts an undesirable brown colour to the paper. The bleaching of kraft
pulp uses large amounts of chlorine-based chemicals and sodium hydrosulfite.
These bleaching chemicals cause several effluent-based problems in the pulp and
paper industries. Byproducts from using these chemicals are chlorinated organic
substances, some of which are toxic, mutagenic, persistent and bioaccumulate, and
cause numerous harmful disturbances in biological systems (Onysko 1993). In
response to public interests, paper industries are currently changing practices to
minimise the use of chlorine-based chemicals. The available options are oxygen
delignification, extended cooking and substitution of chlorine dioxide for chlorine,
hydrogen peroxide and ozone. But most of these methods involve high capital
investment for process change. Thus, an alternative and cost-effective method, i.e.,
use of enzymes, has provided a very simple and economic way to reduce the use of
chlorine and other bleaching chemicals. Biobleaching involves using microorgan-
isms and enzymes for bleaching pulp. It relies on the ability of some microor-
ganisms to depolymerize lignin directly and on the use of microorganism or
enzymes that attack hemicellulose and hence favour subsequent depolymerization
(Jimenez et al. 1997).
Several criteria are essential for choosing a microorganism to produce xylanases.
To give the desired bleaching effect, the resulting enzyme preparation must be
completely free of any cellulase activity (Srinivasan and Rele 1999; Subramaniyan
and Prema 2000; Beg et al. 2001), since any cellulase activity will have serious
economic implications in terms of cellulose loss, degraded pulp quality and
increased effluent treatment cost.
Use of hemicellulolytic enzymes was the first large-scale application of enzymes
in the pulp and paper industry (Viikari et al. 1986). Limited hydrolysis of hemi-
cellulose in pulps by hemicellulases (mainly xylanases) increased the extractability
of lignin from the kraft pulps and reduced the chlorine required in subsequent
bleaching. The xylanase from T. reesei has been reported to act uniformly on all
accessible surfaces of kraft pulp and to be effective during biobleaching (Saake
et al. 1995; Suurnakki et al. 1996; Bhat 2000).
The enzyme is already on the market and has been successfully applied in the
bleaching process as demonstrated by Shatalov and Pereira (2007), nevertheless
many efforts have been done with the aim to improve its performance.
An xylanase with highly specific activity towards xylan was reported by Li et al.
(2005), with a narrow optimum pH range (7.0–7.5) and decrease in the chlorine
136 A.G. Rodrigues
6.9 Baking
The hydrolysis of xylan by endo-xylanases has already been reported during the last
decades in order to elucidate the structure of such xylooligosaccharides and the
enzymatic activity (Gorbacheva and Rodionova 1977; Meagher et al. 1988;
Kormelink et al. 1993; Anand and Vithayathil 1996; Vardakou et al. 2005; Rantanen
et al. 2007; Yang et al. 2011; Li et al. 2012, 2014).
Recently, the manufacture of xylooligosaccharides (XOS) from lignocellulosic
materials as novel sweeteners and functional foods attracts growing interest.
Oligosaccharides are generally defined as saccharides containing between 2 and 10
sugar moieties (Wang et al. 2009; Damen et al. 2012). As one of the non-digestible
oligosaccharides (NDOs), XOS exhibit many important and interesting functional
138 A.G. Rodrigues
6.11 Feedstock
6.12 Xylitol
Xylitol is the second most abundant polyol derived from xylose (Kirilin et al.
2012). Its application hits a broad spectrum in the actual scenario and more are
promising to come true (Aranda-Barradas et al. 2010; Kaialy et al. 2014). The sugar
alcohol is metabolised in the body by insulin-independent pathways and presents
good sweetness as sucrose, however, being less caloric. For these reasons the
molecule has been applied in the food industry (present in chewing gums, sweets,
soft drinks and ice creams, for instance) as an artificial sweetener in the treatment of
diabetes and erythrocytic glucose-6-phosphate dehydrogenase deficiency (Ping
et al. 2013).
It has also been used as a starting material in organic chemistry reactions, as a
potential anticancer compound, in the biomedical field, as a dental caries inhibitor
and oral biofilm formation (Huang et al. 2011; Li et al. 2015a, b; Cardoso et al.
2014; Ma et al. 2014). The molecule has a strong market position, in the range of
90–340 million dollars a year, with a production of 20,000–40,000 tons per year,
and therefore has been extensively studied regarding its production parameters
(Granström et al. 2007; Aranda-Barradas et al. 2010; Li et al. 2015a, b).
Kirilin et al. (2012) described the aqueous phase reforming (APR) process to
produce hydrogen. The work led to interesting results about the stability of
hydrogen formed and the conversion of xylitol. Apart from that, the researchers also
obtained higher yield and selectivities of xylitol than sorbitol towards hydrogen.
Methods to improve the performance in the food industry have also been attempted
as the microencapsulation of xylitol in gum arabic by complex coacervation method
(Santos et al. 2015). Saleh et al. (2014) explored the obtainment of xylitol through
olive stones. The researchers employed response surface methodology (RSM) to
optimise the hydrolysis conditions, analysing treatment temperature and process
time as key factors. Another study (Zhang et al. 2015) conducted at high temper-
ature, using engineered Kluyveromyces marxianus, resulted in high yield, demon-
strating the potential of thermo-tolerant organisms in the biomass conversion of
xylitol, using glycerol as the best substrate in this case. The simultaneous of ethanol
and xylitol production by pinch analysis was performed by Franceschin et al.
(2011).
The group of professor Miyatake in Japan recently reported synthesis of xylitol
microcapsules to encapsulate phase-change materials (PCM), a material capable of
storing and releasing large amount of energy, thereby improving its durability while
adding the advantages of smaller particles in relation to the bulk material (Makuta
et al. 2015).
6 Endo-b-1,4-xylanase: An Overview of Recent Developments 141
Acknowledgments The author thanks Prof. V.K. Gupta for the invitation to write this chapter.
The previous work in the field was supported by the Coordination for the Improvement of Higher
Level Personnel (CAPES) and São Paulo Research Foundation (FAPESP), Brazil.
References
Anand L, Vithayathil PJ (1996) Xylan-degrading enzymes from the thermophilic fungus Humicola
Zanuginosa (Griffon and Maublanc) Bunce: action pattern of xylanase and p-glucosidase on
xylans, xylooligomers and arabinoxylooligomers. J Ferment Bioeng 81:511–517
Anasontzis GE, Zerva A, Stathopoulou PM, Haralampidis K, Diallinas G, Karagouni AD,
Hatzinikolaou DG (2011) Homologous overexpression of xylanase in Fusarium oxysporum
increases ethanol productivity during consolidated bioprocessing (CBP) of lignocellulosics.
J Biotechnol 152:16–23
Aranda-Barradas JS, Garibay-Orijel C, Badillo-Corona JA, Salgado-Manjarrez E (2010) A
stoichiometric analysis of biological xylitol production. Biochem Eng J 50:1–9
Autio K, Harkonen H, Parkkonen T, Frigard T, Poutanen K, Siika-aho M, Aman P (1996) Effects
of purified endo-beta-xylanase and endo-beta-glucanase on the structural and baking
characteristics of rye doughs. LWT Food Sci Technol 29:18–27
Bailey MJ, Biely P, Poutanen K (1992) Interlaboratory testing of methods for assay of xylanase
activity. J Biotechnol 23:257–270
Bakir U, Yavascaoglu S, Guvenc F, Ersayin A (2001) An endo-beta-1,4-xylanase from Rhizopus
oryzae: production, partial purification and biochemical characterization. Enzyme Microb
Technol 29:328–334
Ball AS, McCarthy AJ (1988) Saccharification of straw by actinomycete enzymes. J Gen Microb
134:2139–2147
Battan B, Sharma J, Dhiman SS, Kuhad RC (2007) Enhanced production of cellulase-free
thermostable xylanase by Bacillus pumilus ASH and its potential application in paper industry.
Enzyme Microb Technol 41:733–739
Beachemin KA, Rode LM, Karren D (1999) Use of enzymes in feed lot finishing diets. Can J
Anim Sci 79:243–246
Beg QK, Kapoor M, Mahajan L, Hoondal GS (2001) Microbial xylanases and their industrial
applications: a review. Appl Microb Biotechnol 56:326–338
Belkacemi K, Hamoudi S (2003) Enzymatic hydrolysis of dissolved corn stalk hemicelluloses:
reaction kinetics and modeling. J Chem Technol Biotechnol 78:802–808
Bhargav S, Panda BP, Ali M, Javed S (2008) Solid-state fermentation: an overview. Chem
Biochem Eng Q 22:49–70
Bhat MK (2000) Cellulases and related enzymes in biotechnology. Biotechnol Adv 18:355–383
Bibi Z, Shahid F, Qader SAU, Aman A (2015) Agar–agar entrapment increases the stability of
endo-beta-1,4-xylanase for repeated biodegradation of xylan. Int J Biol Macromol 75:121–127
Biely P (1985) Microbial xylanolytic systems. Trends Biotechnol 3:286–290
6 Endo-b-1,4-xylanase: An Overview of Recent Developments 143
Esteghlalian AR, Kazaoka MM, Lowery BA, Varvak A, Hancock B, Woodward T, Turner JO,
Blum DL, Weiner D, Hazlewood GP (2008) Prebleaching of softwood and hardwood pulps by
a high performance xylanase belonging to a novel clade of glycosyl hydrolase family 11.
Enzyme Microb Technol 42:395–403
Fenel F, Leisola M, Jänis J, Turunen O (2004) A de novo designed N-terminal disulphide bridge
stabilizes the Trichoderma reesei endo-1,4-beta-xylanase II. J Biotechnol 108:137–143
Fenel F, Zitting A-J, Kantelinen A (2006) Increased alkali stability in Trichoderma reesei
endo-1,4-beta-xylanase II by site directed mutagenesis. J Biotechnol 121:102–107
Filho EXF, Puls J, Coughlan MP (1993) Physicochemical and catalytic properties of a
low-molecular-weight endo-l,4-beta-D-xylanase from Myrothecium verrucaria. Enzyme
Microb Technol 15:535–540
Franceschin G, Sudiro M, Ingram T, Smirnova I, Brunner G, Bertucco A (2011) Conversion of rye
straw into fuel and xylitol: a technical and economical assessment based on experimental data.
Chem Eng Res Des 89:631–640
Fujimoto Z, Kuno A, Kaneko S, Kobayashi SYH, Kusakabe I, Mizuno H (2000) Crystal structure
of Streptomyces olivaceoviridis E-86 beta-xylanase containing xylan-binding domain. J Mol
Biol 300:575–585
Georis J, Giannotta F, Lamotte-Brasseur J, Devreese B, van Beeumen J, Granier B, Frere J-M
(1999) Sequence, overproduction and purification of the family 11 endob-1,4-xylanase
encoded by the xyl1 gene of Streptomyces sp. S38. Gene 237:123–133
Gomes DJ, Gomes J, Steiner W (1994) Factors influencing the induction of endo-xylanase by
Thermoascus aurantiacus. J Biotechnol 33:87–94
Gonçalves TA, Damásio ARL, Segato F, Alvarez TM, Bragatto J, Brenelli LB, Citadini APS,
Murakami MT, Ruller R, Leme AFP, Prade RA, Squina FM (2012) Functional characterization
and synergic action of fungal xylanase and arabinofuranosidase for production of
xylooligosaccharides. Bioresour Technol 119:293–299
Gorbacheva IV, Rodionova NA (1977) Studies on xylan-degrading enzymes II. Action pattern of
endo-1,4-beta-xylanase from aspergillus niger str. 14 on xylan and xylooligosaccharides.
Biochim Biophys Acta 484:94–102
Granström TB, Izumori K, Leisola M (2007) A rare sugar xylitol. Part II: biotechnological
production and future applications of xylitol. Appl Microb Biotechnol 74:273–276
Guerfali M, Gargouri A, Belghith H (2011) Catalytic properties of Talaromyces thermophilus-
l-arabinofuranosidase and its synergistic action with immobilized endo-beta-1,4-xylanase.
J Mol Catal B Enzym 68:192–199
Haapala R, Parkkinen E, Suominen P, Linko S (1996) Production of endo-1,4-beta-glucanase and
xylanase with nylon-web immobilized and free Trichoderma reesei. Enzyme Microb Technol
18:495–501
Harris AD, Ramalingam C (2010) Xylanases and its application in food industry: a review. J Exp
Sci 1:01–11
He J, Yu B, Zhang K, Ding X, Chen D (2009) Expression of endo-1,4-beta-xylanase from
Trichoderma reesei in Pichia pastoris and functional characterization of the produced enzyme.
BMC Biotechnol 9:56
Heck JX, Soares LHB, Hertz PF, Ayub MAZ (2006) Purification and properties of a xylanase
produced by Bacillus circulans BL53 on solid-state cultivation. Biochem Eng J 32:179–184
Henrissat B (1991) A classification of glycosyl hydrolases based on amino acid sequence
similarities. Biochem J 280:309–316
Hrmova M, Biely P, Vrsanska M, Petrakova E (1984) Induction of cellulose- and xylan-degrading
enzyme complex in the yeast Trichosporon cutaneum. Arch Microb 138:371–376
Huang CF, Jiang YF, Guo GL, Hwang WS (2011) 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 102:3322–3329
Hung K-S, Liu S-M, Tzou W-S, Lin F-P, Pan C-L, Fang T-Y, Sund K-H, Tang S-J (2011)
Characterization of a novel GH10 thermostable, halophilic xylanase from the marine bacterium
Thermoanaerobacterium saccharolyticum NTOU1. Process Biochem 46:1257–1263
6 Endo-b-1,4-xylanase: An Overview of Recent Developments 145
Huy ND, LeNguyen C, Seo J-W, Kim D-H, Park S-M (2015) Putative endoglucanase PcGH5 from
Phanerochaete chrysosporium is a b-xylosidase that cleaves xylans in synergistic action with
endo-xylanase. J Biosci Bioeng 119:416–420
Hwang T, Lim HK, Song HY, Cho SJ, Chang J-S, Park N-J (2010) Cloning and characterization of
a xylanase, KRICT PX1 from the strain Paenibacillus sp. HPL-001. Biotechnol Adv
28:594–601
Irfan M, Nadeem M, Syed Q (2014) One-factor-at-a-time (OFAT) optimization of xylanase
production from Trichoderma viride-IR05 in solid-state fermentation. J Radiat Res Appl Sci
7:317–326
Jacobsen SE, Wyman CE (2000) Cellulose and hemicellulose hydrolysis models for application to
current and novel pretreatment processes. Appl Biochem Biotechnol 84–86:81–96
Jänis J, Pulkkinen P, Rouvinen J, Vainiotalo P (2007) Determination of steady-state kinetic
parameters for a xylanase-catalyzed hydrolysis of neutral underivatized xylooligosaccharides
by mass spectrometry. Anal Biochem 365:165–173
Jimenez L, Martinez C, Perez I, Lopez F (1997) Biobleaching procedures for pulp and agricultural
residues using Phanerochaete chroysosporium and enzymes. Process Biochem 4:297–304
Jun H, Bing Y, Keying Z, Daiwen C (2009) Functional characterization of a recombinant
thermostable xylanase from Pichia pastoris: a hybrid enzyme being suitable for xylooligosac-
charides production. Biochem Eng J 48:87–92
Juturu V, Wu JC (2012) Microbial xylanases: engineering, production and industrial applications.
Biotechnol Adv 30:1219–1227
Kaialy W, Maniruzzaman M, Shojaee S, Nokhodchi A (2014) Antisolvent precipitation of novel
xylitol-additive crystals to engineer tablets with improved pharmaceutical performance. Int J
Pharm 477:282–293
Kapoor M, Nair LM, Kuhad RC (2008) Cost-effective xylanase production from free and
immobilized Bacillus pumilus strain MK001 and its application in saccharification of Prosopis
juliflora. Biochem Eng J 38:88–97
Karaoglana M, Yildiza H, Inan M (2014) Screening of signal sequences for extracellular
production of Aspergillus niger xylanase in Pichia pastoris. Biochem Eng J 92:16–21
Kim T, Jeong JC, Yoo YJ (2012) Hydrophobic interaction network analysis for thermostabiliza-
tion of a mesophilic xylanase. J Biotechnol 161:49–59
Kim HM, Lee KH, Kim KH, Lee D-S, Nguyen QA, Bae H-J (2014) Efficient function and
characterization of GH10 xylanase (Xyl10 g) from Gloeophyllum trabeum in lignocellulose
degradation. J Biotechnol 172:38–45
Kirilin AV, Tokarev AV, Kustov LM, Salmi T, Mikkola J-P, Yu DM (2012) Aqueous phase
reforming of xylitol and sorbitol: comparison and influence of substrate structure. Appl Catal A
Gen 435–436:172–180
Knob A, Carmona EC (2008) Xylanase production by Penicillium sclerotiorum and its
characterization. World Appl Sci J 4:277–283
Ko C-H, Tsai C-H, Tu J, Yang B-Y, Hsieh D-L, Jane W-N, Shih T-L (2011) Identification of
Paenibacillus sp. 2S-6 and application of its xylanase on biobleaching. Int Biodeterior
Biodegrad 65:334–339
Koch E (1886) Pharm. Z. Russland, 26, 657. Ber. them. Ges., Ref., 20, 145. 1887
Kormelink FJM, Hoffmann RA, Gruppen H, Voragen AGJ, Kamerling JP, Vliegenthart FG (1993)
Characterisation by H NMR spectroscopy of oligosaccharides derived from alkali-extractable
wheat-flour arabinoxylan by digestion with endo-(1 + 4)-P-D-xylanase III from Aspergillus
awamori. Carbohydr Res 249:369–382
Krengel U, Dijkstra BW (1996) Three-dimensional structure of endo-1,4-beta-xylanase I from
Aspergillus niger: molecular basis for its low pH optimum. J Mol Biol 263:70–78
Kulkarni N, Shendye A, Rao M (1999) Molecular and biotechnological aspects of xylanases.
FEMS Microb Rev 23:411–456
Kumar KS, Manimaran A, Permaul K, Singh S (2009) Production of b-xylanase by a
Thermomyces lanuginosus MC 134 mutant on corn cobs and its application in biobleaching of
bagasse pulp. J Biosci Bioeng 107:494–498
146 A.G. Rodrigues
Li XT, Jiang ZQ, Li LT, Yang SQ, Feng WY, Fan JY, Kusakabe I (2005) Characterization of a
cellulase-free, neutral xylanase from Thermomyces lanuginosus CBS 288.54 and its
biobleaching effect on wheat straw pulp. Bioresour Technol 96:1370–1379
Li L, Tian H, Cheng Y, Jiang Z, Yang S (2006) Purification and characterization of a thermostable
cellulase-free xylanase from the newly isolated Paecilomyces themophila. Enzyme Microb
Technol 38:780–787
Li X, She Y, Sun B, Song H, Zhu Y, Lv Y, Song H (2010) Purification and characterization of a
cellulase-free, thermostable xylanase from Streptomyces rameus L2001 and its biobleaching
effect on wheat straw pulp. Biochem Eng J 52:71–78
Li X, Li E, Zhu Y, Teng C, Sun B, Song H, Yang R (2012) A typical endo-xylanase from
Streptomyces rameus L2001 and its unique characteristics in xylooligosaccharide production.
Carbohydr Res 359:30–36
Li H, Kankaanpää A, Xiong H, Hummel M, Sixtac H, Ojamo H, Turunen O (2013)
Thermostabilization of extremophilic Dictyoglomus thermophilum GH11 xylanase by an
N-terminal disulfide bridge and the effect ofionic liquid [emim]OAc on the enzymatic
performance. Enzyme Microb Technol 53:414–419
Li J, Zhou P, Liu H, Xiong C, Lin J, Xiao W, Gong Y, Liu Z (2014) Synergism of cellulase,
xylanase, and pectinase on hydrolyzing sugarcane bagasse resulting from different pretreatment
technologies. Bioresour Technol 155:258–265
Li H, Voutilainen S, Ojamo H, Turunen O (2015a) Stability and activity of Dictyoglomus
thermophilum GH11 xylanaseand its disulphide mutant at high pressure and temperature.
Enzyme Microb Technol 70:66–71
Li Z, Guo X, Feng X, Li C (2015b) An environment friendly and efficient process for xylitol
bioconversion from enzymatic corncob hydrolysate by adapted Candida tropicalis. Chem
Eng J 263:249–256
Lin X-Q, Han S-Y, Zhang N, Hu H, Zheng S-P, Ye Y-R, Lin Y (2013) Bleach boosting effect of
xylanase A from Bacillus halodurans C-125 in ECF bleaching of wheat straw pulp. Enzyme
Microb Technol 52:91–98
Liu J-R, Yu B, Liu F-H, Cheng K-J, Zhao, X (2005) Expression of rumen microbial fibrolytic
enzyme genes in probiotic Lactobacillus reuteri. Appl Environ Microbiol 71:6769–6775
Ma P, Li T, Wu W, Shi D, Duan F, Bai H, Dong W, Chen M (2014) Novel poly(xylitol sebacate)/
hydroxyapatite bio-nanocomposites via one-step synthesis. Polym Degrad Stab 110:50–55
Maalej-Achouri I, Guerfali M, Romdhane IB-B, Gargouri A, Belghith H (2012) The effect of
Talaromyces thermophilus cellulase-free xylanase and commercial laccase on lignocellulosic
components during the bleaching of kraft pulp. Int Biodeterior Biodegrad 75:43–48
Maat J, Roza M, Verbakel J, Stam H, da Silra MJS, Egmond MR, Hagemans MLD, van
Garcom RFM, Hessing JGM, van Derhondel C, van Rotterdam C (1992) Xylanases and their
application in baking. In: Visser J, Beldman G, van Someren MAK, Voragen AGJ (eds) Xylan
and xylanases. Elsevier, Amsterdam, pp 349–360
Makuta T, Kadoya K, Izumi H, Miyatake M (2015) Synthesis of cyanoacrylate-covered xylitol
microcapsules for thermal storage. Chem Eng J 273:192–196
Mamo G, Hatti-Kaul R, Mattiasson B (2006) A thermostable alkaline active endo-beta-
1-4-xylanase from Bacillus halodurans S7: Purification and characterization. Enzyme Microb
Technol 39:1492–1498
Mander P, Choi YH, Pradeep GC, Choi YS, Hong JH, Cho SS, Yoo JC (2014) Biochemical
characterization of xylanase produced from Streptomyces sp. CS624 using an agro residue
substrate. Process Biochem 49:451–456
McCleary BV, McKie VA, Draga A, Rooney E, Mangan D, Larkin J (2015) Hydrolysis of wheat
flour arabinoxylan, acid-debranched wheat flour arabinoxylan and arabino-xylo-
oligosaccharides by b-xylanase, a-L-arabinofuranosidase and b-xylosidase. Carbohydr Res
407:79–96
Meagher MM, Tao BY, Chow JM, Reilly PJ (1988) Kinetics and subsite mapping of A
D-xylobiose- and D-XmsE producing Aspergiflus niger endo-(1+4)-/W-xylanase. Carbodydr
Res 173:273–283
6 Endo-b-1,4-xylanase: An Overview of Recent Developments 147
Michaux C, Pouyez J, Mayard A, Vandurm P, Housen I, Wouters J (2010) Structural insights into
the acidophilic pH adaptation of a novel endo-1,4-beta-xylanase from Scytalidium aci-
dophilum. Biochimie 92:1407–1415
Moukouli M, Topakas E, Christakopoulos P (2011) Cloning and optimized expression of a GH-11
xylanase from Fusarium oxysporum in Pichia pastoris. New Biotechnol 28:369–374
Nakamura S, Wakabayashi K, Nakai R, Aono R, Horikoshi K (1993) Purification and some
properties of an alkaline xylanase from alkaliphilic Bacillus sp. Strain 41 M-1. Appl Environ
Microb 59:2311–2316
Onysko KA (1993) Biological bleaching of chemical pulps: a review. Biotechnol Adv 11:179–198
Pandey A (2002) Solid-state fermentation. Biochem Eng J 13:81–84
Panwar D, Srivastava PK, Kapoor M (2013) Production, extraction and characterization of alkaline
xylanase from Bacillus sp. PKD-9 with potential for poultry feed. Biocatal Agric Biotechnol
3:118–125
Pastor FIJ, Gallardo Ó,Sanz-Aparicio J, Díaz P (2007) Xylanases: molecular properties and
applications. In: Polaina J, MacCabe AP (eds) Industrial enzymes. Springer, Dordrecht, The
Netherlands, pp 65–82
Pedersena MB, Dalsgaarda S, Arenta S, Lorentsena R, Knudsen KEB, Yua S, Lærke HN (2015)
Xylanase and protease increase solubilization of non-starch polysaccharides and nutrient
release of corn- and wheat distillers dried grains with solubles. Biochem Eng J 98:99–106
Phakachoed N, Lounglawan P, Suksombat W (2012) Effects of xylanase supplementation on
ruminal digestibility in fistula tednon-lactating dairy cows fed rice straw. Livest Sci
149:104–108
Ping Y, Ling H-Z, Song G, Ge J-P (2013) Xylitol production from non-detoxified corncob
hemicellulose acid hydrolysate by Candida tropicalis. Biochem Eng J 75:86–91
Polizeli MLTM, Rizzatti ACS, Monti R, Terenzi HF, Jorge JA, Amorim DS (2005) Xylanases
from fungi: properties and industrial applications. Appl Microb Biotechnol 67:577–591
Poutanen K (1997) Enzymes: an important tool in the improvement of the quality of cereal foods.
Trends Food Sci Technol 8:300–306
Prade RA (1996) Xylanases: from biology to biotechnology. Biotechnol Genet Eng Rev
13:101–131
Pradeep GC, Choi YH, Choi YS, Seong CN, Cho SS, Lee HJ, Yoo JC (2013) A novel
thermostable cellulase free xylanase stable in broad range of pH from Streptomyces sp. CS428.
Process Biochem 48:1188–1196
Prasad DY (1993) Enzymatic deinking of laser and xerographic office wastes. Appita J
46:289–292
Pribowo A, Arantes J, Saddler VN (2012) The adsorption and enzyme activity profiles of specific
Trichoderma reesei cellulase/xylanase components when hydrolyzing steam pretreated corn
stover. Enzyme Microb Technol 50:195–203
Puchart V, Biely P (2008) Simultaneous production of endo-beta-1,4-xylanase and branched
xylooligosaccharides by Thermomyces lanuginosus. J Biotechnol 137:34–43
Purkarthofer H, Steiner W (1995) Induction of endo-P-xylanase in the fungus Thermomyces
lanuginosus. Enzyme Microb Technol 17:114–118
Qi Si J, Drost-Lustenberger C (2002) Enzymes for bread, pasta and noodle products. In:
Whitehurst RJ, Law BA (eds) Enzymes in food technology. Sheffield Academic Press,
Sheffield, pp 19–56
Qian C, Liu N, Yan X, Wang Q, Zhou Z, Wang Q (2015) Engineering a high-performance,
metagenomic-derived novelxylanase with improved soluble protein yield and thermostability.
Enzyme Microb Technol 70:35–41
Rantanen H, Virkki L, Tuomainen P, Kabel M, Schols H, Tenkanen M (2007) Preparation of
arabinoxylobiose from rye xylan using family 10 Aspergillus aculeatus endo-
1,4-beta-D-xylanase. Carbohydr Polym 68:350–359
Saake B, Clark T, Puls J (1995) Investigations on the reaction mechanism of xylanases and
mannanases on sprucewood chemical pulps. Holzforschung 49:60–68
148 A.G. Rodrigues
Várnai A, Huikko L, Pere J, Siika-aho M, Viikari L (2011) Synergistic action of xylanase and
mannanase improves the total hydrolysis of softwood. Bioresour Technol 102:9096–9104
Verma D, Satyanarayana T (2012) Cloning, expression and applicability of thermo-alkali-stable
xylanase of Geobacillus thermoleovorans in generating xylooligosaccharides from
agro-residues. Bioresour Technol 107:333–338
Vieira DS, Degrève L, Ward RJ (2009) Characterization of temperature dependent and
substrate-binding cleft movements in Bacillus circulans family 11 xylanase: a molecular
dynamics investigation. Biochim Biophys Acta 1790:1301–1306
Viikari L, Ranva M, Kantelinen A, Sandquist J, Linko M (1986) Bleaching with enzymes. In:
Third international conference in biotechnology in pulp and paper industry, Stockholm,
pp 67–69
Wakiyama M, Tanaka H, Yoshihara K, Hayashi S, Ohta K (2008) Purification and properties of
family-10 endo-1,4-b-xylanase from Penicillium citrinum and structural organization of
encoding gene. J Biosci Bioeng 105:367–374
Wang J, Sun BG, Cao YP, Tian Y, Wang CT (2009) Enzymatic preparation of wheat bran
xylooligosaccharides and their stability during pasteurization and autoclave sterilization at low
pH. Carbohydr Polym 77:816–821
Wase DAJ, Raymahasay S, Wang CW (1985) Production of/ -D-glucosidase, endo- 1,4-/ -
D-glucanase and D-xylanase from straw by Aspergillus fumigatus IMI 255091. Enzyme
Microb Technol 7:225–229
Wong KKY, Tan LUL, Saddler JN (1988) Multiplicity of b-1,4-xylanase in microorganisms:
functions and applications. Microbiol Rev 52:305–317
Xue Y, Peng J, Wang R, Song X (2009) Construction of the trifunctional enzyme associating the
Thermoanaerobacter ethanolicus xylosidase-arabinosidase with the Thermomyces lanuginosus
xylanase for degradation of arabinoxylan. Enzyme Microb Technol 45:22–27
Yang H, Wang K, Song X, Xu F (2011) Production of xylooligosaccharides by xylanase from
Pichia stipitis based on xylan preparation from triploid Populas tomentosa. Bioresour Technol
102:7171–7176
Yang Q, Gao Y, Huang Y, Xu Q, Luo X-M, Liu J-L, Feng J-X (2015) Identification of three
important amino acid residues of xylanaseAfxynA from Aspergillus fumigatus for enzyme
activity and formation of xylobiose as the major product. Process Biochem 50:571–581
Yin Y, Chen H, Hahn MG, Mohnen D, Xu Y (2010) Evolution and function of the plant cell wall
synthesis-related glycosyltransferase family 81. Plant Physiol 153:1729–1746
Zhang S, Zhang K, Chen X, Chu X, Sun F, Dong Z (2010) Five mutations in N-terminus confer
thermostability on mesophilic xylanase. Biochem Biophys Res Commun 395:200–206
Zhang J, Zhang B, Wanga D, Gao X, Hong J (2015) Improving xylitol production at elevated
temperature with engineered Kluyveromyces marxianus through over-expressing transporters.
Bioresour Technol 175:642–645
Zhao L-C, Wang Y, Lin J-F, Guo L-Q (2012) Adsorption and kinetic behavior of recombinant
multifunctional xylanase in hydrolysis of pineapple stem and bagasse and their hemicellulose
for Xylo-oligosaccharide production. Bioresour Technol 110:343–348
Zheng H, Liu Y, Liu X, Han Y, Wang J, Lu F (2012) Overexpression of a Paenibacillus
campinasensis xylanase in Bacillus megaterium and its applications to biobleaching of cotton
stalk pulp and saccharification of recycled paper sludge. Bioresour Technol 125:182–187
Zhengqiang J, Kobayash A, Ahsan MM, Lite L, Kitaoka M, Hayashi K (2001) Characterization of
a thermostable family 10 endo-xylanase (XynB) from Thermotoga maritima that cleaves
p-nitrophenyl-P-D-xyloside. J Biosci D Bioeng 92:423–428
Chapter 7
Microbial Xylanases: Sources, Types,
and Their Applications
7.1 Introduction
Table 7.2 The hemicellulolytic enzymes and their substrates (Shallom and Shoham 2003)
Enzymes Substrates
Endo-b-1,4-xylanase b-1,4-Xylan
Exo-b-1,4-xylosidase b-1,4-Xylooligomers
a-L-Arabinofuranosidase a-Arabinofuranosyl, (1–3) xylooligomers, a-1,5-arabinan
Endo-a-1,5-arabinanase a-1,5-Arabinan
a-Glucuronidase 4-O-methyl-a-glucuronic acid, (1–2) xylooligomers
Endo-b-1,4-mannanase b-1,4-Mannan
Exo-b-1,4-mannosidase b-1,4-Mannooligomers, mannobiose
b-Galactosidase a-1,6-Galactopyranose, mannooligomers
b-Glucosidase b-1,4-Glucopyranose, mannopyranose
Endo-galactanase b-1,4-Galactan
In most terrestrial plant species, secondary cell wall thickening takes place by the
deposition of xylan and they are found in-between cellulosic fiber sheets and lignin
with complex binding relationship through hydrogen bonding. Xylan thus forms a
think wall over cellulose and enables the protection from degradation by different
cellulose degrading enzymes. Xylan constitutes 15–30 % of the plant biomass in
hardwood and 7–12 % of the plant biomass in softwood. Since they are the major
hemicellulose group in most of the plant species, they have been considered as one
among the renewable source of energy in the form of biomass. Structure of xylan is
complex in nature as they are composed of b-1,4 linked-D-xylopyranose units with
a-L-arabinofuranose and 4-O-methyl-a-D-glucuronopyranosyl acid side chains.
Figure 7.1 shows the major enzymes involved and their site of action during the
enzymatic hydrolysis is xylan (Polizeli et al. 2005).
154 H.A. El Enshasy et al.
Hardwood xylan (e.g., birchwood xylan) consists of more than 70 b-1,4 linked-
D-xylopyranose units linked by b-1,4 glycosidic bonds and 4-O-methyl-a-D-glu-
curonopyranosyl acid side chains which are found attached at second carbon of
every tenth xylopyranose units. Acetylation rate is higher in these groups and it is
quite frequent at second and third carbon atom on xylopyranose ring. In softwood
xylan, acetylation rate is zero but instead of an acetyl group they possess a-L-
arabinofuranose linked by a-1-3 glycosidic bond at the third carbon atom on
xylopyranose ring. Based on the side chains of the xylan backbone, they are
classified into different groups such as linear homoxylan, arabinoxylan, glu-
curonoxylan, and glucuronoarabinoxylan. These four groups are heterogenic in
terms of their degree and nature of branching. In many plants, xylans are found
partially acetylated, which protect them from complete degradation by xylolytic
enzymes. This could be the reason behind the fact that the complete degradation of
xylan is accomplished only by synergetic interaction of acetyl xylan esterase and
endo-xylanase enzymes.
7 Microbial Xylanases: Sources, Types, and Their Applications 155
enzymes are able to release glucuronic acid. However, the enzyme activity is only
limited on short xylooligomers due to the partial hindrance by acetyl group present in
xylan backbone (Puls et al. 1991). Acetylxylan esterase: (EC 3.1.1.6) it acts by
removing the O-acetyl groups from second and third carbon on b-D-xylopyranosyl
ring of acetyl xylan. This enzyme plays important role in xylan hydrolysis by
removing O-acetyl side chains, since O-acetyl groups in acetyl xylan interfere the
activity of endo-1,4-b-xylanase and b-D-xylosidases on xylan backbone. Xylan
extraction from hardwoods is usually mediated by alkali treatment, resulting in acetyls
hydrolysis, and probably this could be the reason of the late discovery of this enzyme
(Shao and Wiegel 1995; Blum et al. 1999). Ferulic acid esterase: Ferulic acid esterases
(EC 3.1.1.-) act by hydrolyzing the bond between arabinose side chain and ferulic acid
group attached (Crepin et al. 2004). The cooperative functioning among the
above-mentioned xylolytic enzymes enhance the complete biodegradation of xylan. It
was also observed that acetylxylan esterase activity on xylan results in deacetylation
of xylan backbone and these deacetylated xylans are more easily hydrolyzed by
endo-xylanase. Similarly, there was a significant increase in hydrolysis efficiency
when arabinoxylan was pretreated by a-arabinofuranosidase. This was due to the
removal of arabinan side chains which act as hindrance to endo-xylanase activity (de
Vries et al. 1999).
bacteria and fungi have the capacity for hydrolyzing xylans by synthesizing a
variety of xylolytic enzymes. Early reports reveal that many of these microbial
species are plant pathogens as they play an important role in degradation and
invasion of plant tissues. It also shows that xylanases can elicit defense mechanisms
in plants by a collective functioning with other cellulolytic enzyme groups.
Many of the fungal species that are pathogenic to plants produce plant cell wall
polysaccharide degrading enzymes. Endo-1,4-b-xylanase is one among the major
group of such enzymes and they result in partial degradation of cell wall structures
to the region of penetration (Subramaniyan 2000). A hypothetical model of fungal
invasion on plant tissue is shown in Fig. 7.2.
A number of fungal species have been used to produce xylanase since they are
the major producer of xylanases in nature. Table 7.3 shows the major fungal species
used to produce xylanases. Phanerochaete chrysosporium, a potent plant pathogen,
reported to produce xylanase activity of 15–20 U mL−1 along with considerable
amount of cellulose activity (Copa-Patino et al. 1993). Other study also showed that
the thermophilic fungus Thermomyces lanuginosus has the capacity of high xyla-
nase production up to 3576 U mL−1 (Singh et al. 2000). Many fungal species have
been reported as xylanase producers; however, species with high xylanase activity
and negligible cellulase activity are found to be very rare in nature. It was also
observed that the pH optima of many of these fungal xylanases are between pH 5
and 6 even though they are stable at pH between 3 and 8. On the other hand,
bacterial xylanases showed slightly higher pH optima that makes them suitable for
many of the industrial applications. Thermal tolerance of the majority of fugal
xylanase reported so far has been found below 50 °C and in most cases this
Fig. 7.2 Fungal invasion on plant tissue by xylolytic degradation (Prade 1996)
158 H.A. El Enshasy et al.
particular property of fungal xylanase makes them less favorable for application in
harsh industrial conditions. Reduced xylanase yield in fermenter studies is another
major problem associated with fungi xylanase. This was due to poor oxygen
transfer, shear force during fermentation, and the enzyme production process in
fungi is highly regulated by growth morphology (El Enshasy et al. 1999, 2006; El
Enshasy 2007). Generally, the fungal species has filamentous growth pattern, the
fungal growth in fermenters is restricted due to the shear stress, and eventually this
results in poor xylanase production (Dean et al. 1991).
Since 1980s, several bacterial strains which are capable of producing alkaline and
highly thermophilic xylanase enzymes have been reported. Among these wide
spectrum of bacteria, Bacilluss species are the most predominant producers of
endo-1,4-b-xylanase enzyme with negligible amount of cellulose activity under
optimized growth conditions. Other author showed a high yield of xylanase up to
506 IU mL−1 by Bacillus SSP-34 (Subramaniyan and Prema 2002). In general,
bacterial endo-1,4-b-xylanase is suitable for industrial application since they have a
wide pH optima and improved thermostability when compared to fungal xylanases.
Ratto et al. (1992) reported Bacillus circulans strain with xylanase activity of
400 IU mL−1. They observed the optimal enzyme activity was at pH 7 and retained
40 % of activity at high pH up to 9.2. Many other researchers reported
cellulose-free xylanase enzyme from Bacillus stearothermophilus strain T6 (Khasin
et al. 1993; Lundgren et al. 1994). A relatively high xylanase activity was reported
in Bacillus sp. strain NCL when it was grown in zeolite-induced medium
(Balakrishnan et al. 2000). Thermophile bacteria such as Rhodothermus marinus
produced approximately 1.8–4.03 IU mL−1 of thermostable xylanase along with
7 Microbial Xylanases: Sources, Types, and Their Applications 159
and Streptomyces sp. QG-11-3 (Beg et al. 2001). However, several exceptional cases
have also been reported such as in Cellulomonas flavigena; xylan acts as a poor
inducer of xylanase gene (Avalos et al. 1996). In some rare cases, for example, in
yeast strain Trichosporon cutaneum, xylanase biosynthesis is induced by certain
positional isomers. In many bacterial species, the xylanase induction is also possible
with various sugars such as D-xylose, D-maltose, D-glucose, and D-arabinose. On the
other hand, in many fungal species, the natural lignocelluloses such as corn cobs,
rice straw, sugarcane bagasse, and wheat bran were found to be capable of triggering
the xylanase induction (Gupta et al. 2000). Other study conducted by Kumar and
Satyanarayana (2011) reported on the wheat bran-mediated xylanase induction in
Bacillus halodurans TSEV1 strain. However, the hypothetical mechanism of direct
induction is questionable when the transportation across the cell wall can be blocked
by larger molecules. Based on the study of Gomes et al. (1994), there has been a
universally supported concept based on intracellular b-xylosidases. They explains
that the larger xylooligomers formed during xylan hydrolysis are directly transported
into the cell matrix and it was further degraded into xylose residues by b-xylosidase
which are present in the cytoplasm.
Hydrolysis of xylooligomers by hydrolytic transporter proteins during their
transportation through cell membrane is another possible explanation for above
phenomenon. There are also rare cases in which the high xylose concentration in
the media inhibits the xylanase gene expression (Strauss et al. 1995). Exclusion of
inducer transport across the cell membrane is another possible reason for poor gene
expression. In similar cases such as in E. coli it has been observed that the presence
of glucose in the media prevents the lactose transportation through the cell mem-
brane which is inducer for lac operon (Borralho et al. 2002). It was also observed
that the xylanase biosynthesis is mediated by complex metabolic pathway in which
the inducer level and the level of repressor molecule to that particular inducer vary
with the organism and their growth medium. For example, the xylanase production
by Streptomycetes sp. was increased when cellulose substrates are used in the
growth media, whereas in Cellulomonas favigena, sugarcane bagasse was found to
be the best inducer of xylanase enzyme (Alejandro et al. 2007). Addition of xylose
in fermentation media increased xylanase production to a significant amount in
Bacillus sp. (Gupta and Kar 2009). However, other studies showed that the pres-
ence of readily utilizable sugars such as glucose, xylose, and galactose can suppress
xylanase biosynthesis in other strains such as in Streptomyces sp. (Bajaj and Singh
2010) and in T. reesei (Mach-Aigner et al. 2010).
Xylanases are considered as one of the major hydrolyzing enzyme groups as they
mediate the xylan degradation to simpler utilizable units. Activity of these enzymes
depends on the substrate specificity as well as substrate complexity. Wong et al.
(1988) classified xylanases into two major groups based on the end product of
7 Microbial Xylanases: Sources, Types, and Their Applications 161
hydrolysis reaction. They were debranching enzymes which liberate arabinose and
non-debranching enzymes which do not liberate arabinose residues from a,1,3,L-
arabinofuranosyl. A few years later, another classification system was introduced
based on the physiological properties of xylanase enzymes such as isoelectric point
(PI) and molecular weight (MW). Therefore, xylanases were classified into two
major groups: end xylanase enzymes with molecular weight lower than 30 kDa and
basic pH and those with molecular weight higher than 30 kDa and acidic pH.
However, few years later, this classification system was found incorrect since it
matches for only a few xylanase enzymes.
Henrissat and Bairoch (1993) proposed a broad classification system based on
the similarities in amino acid sequence and catalytic module. The glycoside
hydrolases are further classified into different families. This classification system
has got wide acceptance since it gives information regarding the structural prop-
erties, the catalytic mechanisms of the enzyme, and their evolutionary relationship
within the group. Xylanases were grouped under glycoside hydrolases (GH) 5, 8,
10, and 11 families. However, the enzymes with multi-domain that exhibit
detectable amount of xylanase activity are grouped into GH7, 16, 43, and 62
families (Cantarel et al. 2009). Table 7.5 comprises major xylolytic enzymes, their
classification into CE, GH families, and their current crystallographic structure.
Some of these GH families are further classified into super families that represent a
more distant common evolutionary ancestor. Major GH family xylanase and their
general structures are shown in Fig. 7.3 (Table 7.6).
Table 7.5 Xylolytic enzymes, their classification into CE, GH families, and the their current
crystallographic structure status
Enzyme Enzyme family Structure status
EC GH
Endo-b-1,4-xylanase 3.2.1.8 5, 8, 10, 11 and GH5-Cryst. (1BQC)
43 GH8-Cryst. (1IS9)
GH10-1FH7,
1CLX, 1XYZ
GH11-1H4G, 1BCX,
1IGO
GH43-(1GYD)
Exo-b-1,4-xylosidase/b-xylosidases 3.2.1.37 3, 39, 43, 52 GH3-Cryst. (1EX1)
and 54 GH39-Cryst.
GH43-(1GYD)
a-L-Arabinofuranosidase 3.2.1.55 3, 43 and 51 GH3-(1EX1)
GH43-(1GYD)
GH51-Cryst
Endo-a-1,5-arabinanase 3.2.1.99 43 1GYD
a-Glucuronidase 3.2.1.139 67 1K9D, 1GQI
Acetyl xylan esterase 3.1.1.72 1 (1JJF)
5 1QOZ, 1G66
Ferulic acid esterases 3.1.1.73 1 1JJF, 1GKK
Source Shallom and Shoham 2003
162 H.A. El Enshasy et al.
Fig. 7.3 Structures of xylanase major GH family xylanase (Collins et al. 2005a, b)
Table 7.6 CBMs of known structure linked to xylanases (Berrin and Juge 2008)
CBMs family Source of xylanase GH family PDB code
CBM-2 Cellulomonas fimi 10A, 11A 1EXG, 2XBD, 1HEH
CBM-4 Rhodothermus marinus 10A 1K45
CBM-6 Clostridium thermocellum 11A 1UXX, 1NAE, 1UY4
CBM-9 Thermotoga maritima 10A 1I8A
CBM-10 Cellvibrio japonicus 10A 1QLD
CBM-13 Streptomyces olivaceovidris 10A 1XYF
CBM-15 Cellvibrio japonicus 10C 1GNY
CBM-22 Clostridium thermocellum 10B 1DYO
CBM-36 Paenibacillus polymyxa 43A 1UX7
Most of the glycoside hydrolases which are classified under GH10 are
endo-b-1,4-xylanases. Apart from these enzyme, G10 group also consists of a small
number of endo-b-1,3-xylanases (EC 3.2.1.32), which cleave b-1,3-glycosidic
linkages randomly in b-1,3-xylans backbone. A typical GH10 xylanases has a wide
spectrum of substrate specificity as they are able to hydrolyze various forms of xylans
in nature. These enzymes are capable to attack not only the linear forms of xylan but
also the highly branched heteroxylans and xylosomes. Kolenova et al. (2006)
7 Microbial Xylanases: Sources, Types, and Their Applications 163
analyzed the hydrolysis product from glucuronoxylan and confirmed that GH10
xylanases can attack xylan at its non-reducing end if there are two or more unbran-
ched xylose residues which are present on backbone. It was also found that GH10
xylanases can even hydrolyse the highly branched arabinoxylan into small fragments.
Plant xylanases which have been identified so far are grouped under GH10 family. In
recent years many studies were carried out to characterize these plant xylanases and it
was observed that these enzymes have limited substrate specificity (Chithra and
Muralikrishna 2008). The study of Van Campenhout et al. (2007) was focused on the
characterization of GH10 xylanase isolated from barley and they reported that the
enzyme can release small xylose units from various substrates. GH10 xylanases, in
addition to their xylolytic activity, are also active against glucose-derived substrates
like aryl-cello-oligosaccharides (Charnock et al. 1997; Andrews et al. 2000).
Three-dimensional structure of xylan was first described by Atkins (1992).
Subsequently, more studies were carried out using NMR spectrometry and X-ray
crystallography techniques to obtain more carbohydrate structure details. It was
found that under crystallized condition, the xylan backbone exhibits a threefold
left-handed confirmation and the geometry showed that there is no effect on b-1,4
glycosidic bonds by the side chains of those attached. Electron diffraction pattern
study confirmed the structure and the hexagonal morphology. The hydrogen at the
fifth carbon atom on xylose ring has binding properties either intra- or interchain.
The study also suggested that the structural confirmation of D-xylose ring indicated
the di-equatorial binding property of both 1–4 and 1–3 glycosidic linkages.
The X-ray crystallographic image of endo-b-1,4-xylanases enzyme isolated from
Thermoascus aurantiacus has been used to explain the general structure of GH
family 10 xylanase (Fig. 7.4). It was found that the 32.5 kDa long polypeptide
chain with a b/a-fold TIM barrel structure consists of eight major b strands which
are arranged side by side and parallel, forming a cylinder in the center followed by
eight major a helixes. In addition, around six short helixes are also found attached
with the polypeptide chain. The catalytic domains of family 10 xylanases are found
in cylindrical shape and the overall side view resembles a ‘Salad bowl’ (Fig. 7.5).
In this configuration, the catalytic sites are seen relatively closer to the carboxyl
terminus end of xylan backbone. Since the molecular weight of these xylanase is
higher, they always exhibit low degree of oligosaccharide polymerization. Both the
disulfide bond and salt bridges present on the xylan structure improve its ther-
mostability. In the overall structure, the top phase of the molecule which is b-barrel
side has larger than the bottom face, the a-b turns. This was due to the elaborate
architecture of b–a loops (Natesh et al. 1999).
GH11 xylanase unlike the other xylanase families, they only consist of
endo-b-1,4-xylanases that are exclusive enzymes cleaves b-1,4-xylosidic bonds
between xylose monomers. These enzymes are often sub-classified based on their
164 H.A. El Enshasy et al.
PI values into acidic and alkaline xylanases (Joshi et al. 1997). Many studies
demonstrated the correlation between the pH optima and the amino acid residues
which are present adjacent to the catalytic site of these enzymes. In many cases, it
was observed that the acidic xylanase (pH < 5) consists of asparagine, whereas in
alkaline xylanase has arginine (Krengel and Dijkstra 1996; Fushinobu et al. 1998).
This was also confirmed by other study which reported that arginine introduction to
bacterial xylanase shifts the pH optimum to acidic range (Pokhrel et al. 2013).
Heteroxylan, xylobiose, and xylotriose are the major subunits formed during GH11
xylanase-mediated xylan hydrolysis. Further hydrolytic activities on these subunits
were found negligible; however, hydrolysis products such as xylotetrose,
xylopentose, and xylohexose are further hydrolyzed by GH11 xylanases
(Cervera-Tison et al. 2009). These enzymes cleave the xylose backbone at
unsubstituted regions which are quite away from the branched xylose present at
non-reducing end. It was also found that GH11 xylanases require minimum of three
unsubstituted consecutive xylose residues for the primary binding and initiation of
hydrolysis (Biely et al. 1997; Katapodis and Christakopoulos 2008).
Ko et al. (1992) reported the first structural description of a GH11 xylanase that
was isolated from Bacillus pumilus. But his analysis was not very precise and he
failed to deposit the 3D structure in protein database. To date, more than 100 GH11
xylanase 3D structures from different fungal and microbial species have been
solved and made available in CAZy database. Xylanases isolated from a ther-
mophilic fungi Thermomyces lanuginosus are widely used to explain the general
structure of GH11 xylanases. The structure shows a globular protein composed of
two b-sheets (b-a, b-b) with a molecular weight of 25 kDa. The outer b-sheet (b-a)
is composed of five antiparallel b strands with polar and uncharged amino acids
such as threonine and serine. The inner b-b sheet is made up of nine antiparallel b
strands of which the outer sides are developed to catalytic sites and the inner sides
are found attached with b-a forming a hydrophilic core (Torronen et al. 1994). In
general, most of the GH11 xylanases are made of only a single catalytic domain and
found in b-jelly roll structure. It also resembles the shape of partially closed right
hand in which two b-sheets represent the ‘fingers,’ the loops between b-strands of
b-b such as B7 and B8 represent the ‘thumb,’ and the twisted inner sides of b-b and
a-helix represent the ‘palm’ of the hand (Gruber et al. 1998).
While comparing these two major GH families, the GH11 family consists only
of ‘true xylanases’ displaying substrate specificity toward D-xylose containing xylan
substrates. Their catalytic versatility is lower when compared to GH10 xylanases
and the products from its action such as xylobiose and xylotriose usually required
further hydrolysis by other xylanase enzymes. The GH10 xylanases can mediate
hydrolysis until the release of xylose as end product. Another interesting feature of
GH10 xylanases is their ability to tolerate glucose-derived to a certain extend in
addition to their xylanolytic activity that exhibits in addition to their xylanolytic
activity (Collins et al. 2005a). It is generally accepted that GH10 can cleave dec-
orated regions on AX backbone and its activity is less hampered by a-L-arabino-
furanosyl and acetyl or 4-O-methyl-D-glucuronate side chains present on xylan
166 H.A. El Enshasy et al.
backbone, whereas GH11 xylanases are very specific in their action as they cleave
only at the unsubstituted regions (Biely et al. 1997). This property is also reflected
in the shape of their active binding sites, where GH10 xylanase has a shallow
groove active site which has less affinity toward the unsubstituted regions and
GH11 xylanases possess a cleft-shaped active site which has higher affinity toward
unsubstituted consecutive xylose.
position of catalytic residues on the enzymes (Collins et al. 2005b). Many of the
endo-1,4-b-xylanase exhibit a double-displacement mechanism in their operation
on specific substrate molecules (Fig. 7.6). In most cases, the first acid/base catalytic
residue initiate the hydrolysis by protonating the xylopyranosyl linkages between
the xylan monomers and this process is termed as glycosylation. The second cat-
alytic residue which acts as a nucleophile attacks the same linkage and results in the
formation of enzyme–glycosyl intermediate while passing through an
oxocarbenium-ion-like transition state. In consecutive steps which are also termed
as deglycosylation, the first catalytic residue exhibits basic properties by activating
the incoming water molecule and abstracting a proton from it. The activated water
molecule readily attacks the anomeric carbon of the enzyme–glycosyl complex
formed in the previous step and release the product with an a-configuration at the
anomeric carbon (Sidhu et al. 1999; Chiku et al. 2008).
Among the above-mentioned families, endo-xylanases GH10 and GH11 are the
most studied. Endo-xylanase enzymes with relatively higher molecular weight are
grouped into family F/10 and they are found with cellulose-binding domain and a
catalytic domain that are usually connected by linker peptide. Enzyme belonging to
this family also found to have a (b/a)-fold TIM barrel structure (Biely et al. 1997),
whereas the family G/11 xylanase is generally of low molecular weight and is further
168 H.A. El Enshasy et al.
classified into two major groups based on their PI value. Similarly, b-xylosidases,
another major xylolytic enzyme is grouped under GH classification system into
various families such as 3, 39, 43, 52, and 54. Studies showed that b-xylosidases
from the families 3, 39, 52, and 54 use a retaining mechanism to hydrolyze
xylooligomers, while those from family 43 mediate the hydrolysis by inverting the
anomeric configuration. Advanced bioinformatics tools such as Basic Local
Alignment Search (BLAST) and pair-wise alignments of the protein sequences have
been used to identify the xylanase enzymes and compared with closely related
enzymes within the families. The sequences of family 10 xylanase are generally used
to identify mutually exclusive enzymes using BLAST search. Further confirmations
are also made using X-ray crystallographic studies. A continuous update to all these
GH family classifications is provided by the carbohydrate-active enzyme database
(CAZy) since there is a direct relationship between the protein folding and the
sequence. This system is universally accepted as it can provide structural features of
enzymes that help to predict their sole substrate specificity; it helps to identify and
reveal the evolutionary relationships between the enzyme groups, and also act as a
convenient tool to derive mechanistic information.
The X-ray crystallographic studies on the xylanase structure reveal that most of
them have a modular structure composed of a catalytic site and one or more
carbohydrate-binding modules (CBMs) that are interconnected by flexible linkages
(Kulkarni et al. 1999; Subramaniyan and Prema 2002; Collins et al. 2005b).
A CBM is defined as a specially arranged amino acid sequence with high affinity to
carbohydrate molecules within a carbohydrate-active hydrolytic enzyme. These
unique sites are designed to bind with specific polysaccharides on plant tissue and
mediate the structural damage by enzymatic hydrolysis (Bolam et al. 2001;
Boraston et al. 2002). Similar to the catalytic modules of glycoside hydrolases,
CBMs are classified into 64 families according to their amino acid sequence sim-
ilarity and this classification is available in CAZy database (Tomme et al. 1994).
A study of Boraston and his group reported a detailed overview about the structure
and functions of CBMs (Boraston et al. 2000). However, a very few members of
GH11 xylanases carry CMB. Among these only a few of them have been experi-
mentally demonstrated, but others are hypothetically explained based on their
sequential similarities with well-studied ones. It have been also reported that certain
CBMs from Xyl-11 are specific for cellulose and this characteristic feature helps the
xylanase to localize the substrates indirectly even when they are in close association
with cellulose. Major CBMs of known structure which are linked to xylanases are
presented in Table 7.7. Xylanases isolated from fungi such as Penicillium funicu-
losum XynB (Brutus et al. 2004) and Neocallimastix patriciarum XynS20 (Liu
et al. 2008) are found to have CMB1 modules. It was also reported that the
7 Microbial Xylanases: Sources, Types, and Their Applications 169
Almost all the plant species have developed a self-defense mechanism as response to
the pathogenic microbial attack by hydrolases enzymes. Xylanase inhibition by the
proteinaceous inhibitors is one among those defense response widely seen in both
soft- and hardwood plants. A numerous studies have been carried out to understand
this defense mechanism since 1990s and it was found that Triticum aestivum
xylanase inhibitor (TAXI) and xylanase inhibitor protein (XIP) are the two major
groups of proteins responsible for Xyl-11 in specific (Debyser et al. 1999; Rouau
et al. 2006). Apart from the above two groups, the third group of inhibitor known as
thaumatin-like xylanase inhibitor (TLXI) also been isolated from wheat. It was also
found that none of these xylanase inhibitors are effective against plant origin xyla-
nase enzymes (Dornez et al. 2010a, b). Occurrence of these inhibitors seems to be a
significant hindrance in xylanase-mediated industrial process such as bread making
and brewing as they inhibit effective hydrolysis of substrates (Sorensen and Sibbesen
2006). Therefore, it has been an important topic of research to understand the mode
of action of these inhibitors. Sensitivity of GH11 xylanases to proteinaceous inhi-
bitors (XIP-1, TAXI-1, and TLXI) are as given in Table 7.7 (Fig. 7.8).
Fig. 7.8 a Structure of the inhibition complex between P. funiculosum Xyl-11 (in green) and
XIP-I inhibitor (in blue). XIP-I inhibiting loop is in red, catalytic residues are in yellow.
b Structure of the inhibition complex between B. subtilis Xyl 11 (in green) and TAXI-IA inhibitor
(in blue). His374 residue is in red, catalytic residues are in yellow (Paës et al. 2012)
7 Microbial Xylanases: Sources, Types, and Their Applications 171
These are proteins with pI values of 8.0, which occur in two different molecular
forms known as A and B (Gebruers et al. 2008). Form A is seen as full-length
proteins with molecular mass near to 40 kDa, whereas the form B consists of two
polypeptide chains (a longer chain and a shorter chain of size 28–30 and 10–
12 kDa, respectively) that bound together by disulfide bonds (Croes et al. 2009).
These inhibitors are seen largely in wheat grains during the early stages of grain
development and specifically inhibit GH11 xylanase disrespect to their origin
(Tables 7.8, 7.9). Further analysis of originally identified TAXI inhibitor proteins
revealed that it is a mixture of two different proteins of different pI values. They are
named as TAXI-I and TAXI-II and have different inhibition activities on GH11
endo-xylanase isolated from A. niger (Gebruers et al. 2004). Other research carried
out by Raedschelders et al. (2004) identified and isolated the genes coding for these
two inhibitor proteins from ray and barley grains and based on this study Fierens
et al. (2004) developed a Pichia pastoris-based expression system for the
large-scale production. By 2005, two more new TAXI class named TAXI-III and
TAXI-IV were also identified (Igawa et al. 2005). TAXI-type inhibitors are found
to be very specific in their inhibitory characteristics. The available research data
shows that they inhibit only GH11 endo-xylanases of bacterial and fungal origin.
However, the subclass TAXI-II was turned to be an exception, as they are unable to
inhibit GH11 endo-xylanases such as XynBc1 and ExlA isolated from B. cinerea
and A. niger, respectively (Brutus et al. 2005).
Table 7.8 Occurrence of TAXI-type inhibitors in different cereals (Raedschelders et al. 2004)
Cereal Inhibitor proteins
TAXI-I TAXI-II Inhibitor level (ppm)
Wheat + + ca. 38
Durum wheat + + ca. 11
Rye + − ca. 21–37
Barley + − ca. 4–5
Maize, rice, oats, buckwheat − − −
The advances in molecular biology and genetic engineering opened up several new
applications of recombinant DNA technology. The new opportunity for the con-
struction of GM microbial strains with selected enzyme machinery was one of the
7 Microbial Xylanases: Sources, Types, and Their Applications 173
is usually carried out based on the expression host and a wide variety vectors have
been introduced in last few decades. Even though the expression rate is higher in
recombinant strain, the enzyme activity is usually less than in the native producer
strain. Several reasons have been identified for this activity loss and the intercellular
accumulation of the enzyme due to the lack of post-translational modification is
suggested to be the key reason (Schlacher et al. 1996).
It has been more than three decades since filamentous fungi are used in the pro-
duction of xylanase enzyme. Fungal cells are the widely used biofactories in
large-scale production of xylanases and have been considered as more potent
xylanase producers as they are capable of secreting much higher amounts of xyla-
nolytic enzymes in fermentation media than bacteria or yeast (Bergquist et al. 2002;
Fang et al. 2008). Large numbers of fungal strains are widely used for xylanases
production such as many species from Penicillium, Aspergillus, Fusarium,
Trichoderma, and Disporotrichum. However, in many of these fungi, the difficulty
in isolating pure form of xylanase was the major problem since they generally
exhibit a low amount of cellulase activity. Recombinant DNA technology has been
widely applied to overcome this drawback and production of xylanase expression
systems which are free of cellulase activity which improved its acceptability in
industrial applications. Apart from the bacterial expression hosts, filamentous fungi
are also attractive hosts for xylanase expression. It is because of their ability to
secrete the proteins into the fermentation media in large quantities. Aspergillus and
Trichoderma species are the widely used expression hosts in homolog expression
system. Kitamoto et al. (1999) successfully overexpressed XynF1 gene in A. oryzae
under a strong TEF1 promoter in glucose-based submerged fermentation. XynF1,
XynF3, and XynG2 are another important xylanase genes expressed efficiently in
A. oryzae under stronger promoter called P-No 8142 in similar fermentation con-
ditions. Aspergillus niger is another attractive host for recombinant xylanases
production due to its ability to produce and secrete protein in high capacity up to
30 g L−1 in the fermentation media (Hessing et al. 1994). Xylanase coding gene
from T. reesei was successfully expressed in A. niger cD15 stain under
G6P-dehydrogenase promoter and glaA terminator which is isolated from
Aspergillus awamori (Rose and van Zyl 2002). This recombinant xylanase retained
its 75 % activity even after 3 h of incubation at 50 °C. The optimal condition of the
enzyme activity was observed at temperature between 50 and 60 °C and pH 5–6.
Levasseur et al. (2005) also reported on the potential use of Aspergillus niger as host
to express xynB under promoter sequence isolated from Aspergillus nidulans. They
showed that, A. nidulans can be used as efficient host for xylanase expression and
production. Kimura et al. (1998) expressed xynG1 in A. nidulans under its own
promoter showed that the xylanase expression was induced in the presence of
7 Microbial Xylanases: Sources, Types, and Their Applications 175
1992). It has been also observed that a significant increase in xylanase expression in
yeast transformed with the xynA gene without intron (26.2 U mL−1) when compared
to the expression from the gene with intron (16.7 U mL−1) (Li and Ljungdahl 1996).
It is also worthy to note that Chavez et al. (2004) made the first successful attempt to
clone and express the xynA gene which is isolated from P. purpurogenum in S.
cerevisiae. Isolated xynA gene that consists of eight introns regions was spliced
correctly and was integrated on yeast chromosome. The Xln A gene was then
expressed under transcriptional control (XlnR and CreA) where xylose or xylan acts
as inducer and glucose acts as repressor (van Peij et al. 1997). The methanotropic
yeast Pichia pastoris is another widely used xylanase expression host which has
many advantages over S. cerevisiae such as improved secretion efficiency, ease to
attain high cell density culture in inexpensive media, and relative ease in scaled up
fermentation process (Cereghino and Cregg 2000). Moreover, the presence of strong
and regulatory alcohol oxidase (AOX I and AOX 11) promoters which are involved in
methanol metabolism promotes the overexpression of xylanase gene upon the
addition of methanol in the fermentation media (Tsai and Huang 2008). Table 7.11
summarizes some examples of successful cloning of fungal xylanase in yeast.
pUC19 plasmid vector, the xylanase expression was found to be reduced. However,
when the same gene is expressed in B. subtilis using plasmid pBA7, 14-fold
increase in expression was observed. Therefore, Bacillus became more attractive
hosts compared to E. coli and thus used widely in heterogeneous protein produc-
tion. Bacillus sp. used in industrial fermentation are non-pathogenic in nature and in
bacillus-based expression, the proteins are usually secreted out of the cell.
Moreover, they are gram-positive bacteria and therefore, they do not produce
endotoxins (lipopolysaccharides). These features could be of great advantage by
reducing some purification steps in terms of breaking the cells and extraction from
large number of intracellular proteins.
For the past several decades, E. coli has been recognized as the ideal platform for
expression of recombinant proteins. Factors such as rapid growth on inexpensive
media, non-pathogenicity, easy to make gene modification without effecting the cell
growth, easy methods required for isolation and purification of expressed proteins
made E. coli the most successful expression host for industrial use. In spite of these
advantages, due to the lake of specific post-translational modification such as
glycosylation and disulfide bond formation, many heterologous proteins were not
successfully expressed in E. coli. Xylanase was one of those which require
N-glycosylation during post-translation stage, whereas E. coli only can perform
O-glycosylation. However, many exceptional cases such as successful expression
of a glycosylated b-xylosidase gene isolated from thermophilic fungi
P. thermophile have been reported (Teng et al. 2011). It was also observed that the
endo-1,4-b-xylanase expression levels are significantly higher in gram-positive
hosts such as Lactobacillus species and B. subtilis compared to E. coli as they can
perform N-glycosylation.
In general, the recombinant xylanases expressed in E. coli are accumulated in the
cytoplasm or in periplasmic place in the absence of a proper secretory channel on cell
wall (Schlacher et al. 1996). However, in many cases, xylanase activity has also been
detected outside the cell (Karlsson et al. 1997; Ebanks et al. 2000). Xylanase
expression rate in E. coli mainly depends on the efficiency of transcription of
xylanase gene which is cloned under promoter sequence. Even though there are
several cloning vectors and host strains have been introduced, eukaryotic gene
expression in E. coli is unable to carryout due to the absence of a functional promoter
sequence. Basaran et al. (2001) isolated b-xylanase gene from Pichia stipites and
successfully cloned and expressed in E. coli. The recombinant enzyme activity was
low (4 U mg−1) when compared to the activity original enzyme activity (30 U mg−1)
in wild stain. Although a variety of cloning vectors have been used, pET expression
vector systems were reported to be the most efficient for protein expression in E. coli.
Xylanase genes isolated from different sources are cloned under the control of a
strong T7 promoter which is isolated from bacteriophage genome. This promoter
178 H.A. El Enshasy et al.
remains silent until T7 RNA polymerase is produced by the host cell genome.
Table 7.12 shows the major cloning vectors used in E. coli-based expression system.
Ogasawara et al. (2006) amplified xynIII gene from T. reesei genomic DNA
using a RT-PCR and was cloned and expressed in E. coli JM109 using pAG9-3
plasmid vector. When the xylanase gene expression was induced by IPTG, they
could produce 26 mU mL−1 of xylanase activity in cellular extract, which was very
low. However, the XynIII activity increased by about 500-fold (13.2 U mL−1)
compared to the activity in soluble supernatant, when the xylanase inclusion bodies
were refolded in 8 mM urea solution. Zhou et al. (2008a) reported the maximum
xylanase activity up to 49.4 U mg−1 when they cloned and expressed xynII cDNA
isolated from A. usamii in E. coli BL21-Codon plus (DE3) RIL strain using
pET-28a (+) expression vector. The xylanase protein expressed is attached to
6X-His-tag which ease protein purification by affinity chromatography. Jun et al.
(2008) also used the same pET vector to clone and express a b-xylanase gene, xyn2,
isolated from T. reesei in E. coli and reported the enhanced xylanase activity of 650
U mg−1. In some cases, the heterologous cloning and expression results in the
7 Microbial Xylanases: Sources, Types, and Their Applications 179
Table 7.14 Significant achievements made in xylanase protein engineering (Verma and
Satyanarayana 2012)
Xylanase source Approaches/alterations Significant inferences References
Bacillus sp. NG-27 Deletion of Phe4 and Complete loss in activity, Bhardwaj
Trp6 while Trp6 and Tyr343 et al.
affected folding (2010)
B. circulans Amino acid Slight increase in half-life Joo et al.
modification (N52Y) and Tm as compared to (2011)
wild type
B. circulans Asn35Asp Improvement in catalytic Li and
activity Wang
(2011)
N. patriciarum D57N Enhancement in You et al.
thermostability by 10– (2010)
15 °C
G. stearothermophilus Substitution with Significant improvement Sun et al.
N-terminal of T. fusca in thermostability (2007)
T. reesei Improvement of Half-life of the mutant Turunen
C-terminal packaging xylanase was increased et al.
by 63 min (2001)
T. lanuginosus Introduction of Improvement in Gruber
DSM5726 disulfide bond thermostability et al.
between Cys100 and (1998)
Cys154
B. stearothermophilus Introduction of Enhancement in Jeong
No. 236 between Ser and Cys) thermostability by 5 °C et al.
100 and (Asn to Cys) (2007)
150
A. niger BCC144505 Substitution of Enhancement in the Sriprang
arginine in place of thermostability et al.
Ser/Thr (2006)
Thermobifida fusca Error prone PCR to Improved stability at Wang and
obtain a mutant having alkaline pH, 4.5-fold Xia (2008)
T21A, G25P, V87P, increase in Km and
I91T, and G217L 12-fold enhancement in
Kcat/Km
Thermomyces Error prone PCR Alkali-stable mutant Stephens
lanuginosus A54T et al.
(2009)
Although many microbial species such as bacteria, fungi, and actinomycetes have
been reported as xylanase producers, a few number of strains are acceptable by the
industrial society. To meet the current and forecasting market demands, xylanase
production need to be increased by many folds through efficient production
182 H.A. El Enshasy et al.
Table 7.15 Major commercial xylanases produced by SmF techniques (Polizeli et al. 2005)
Commercial Distributors Microorganism Application
name
Allzym PT Alltech Aspergillus Animal feed
niger improvement
Bio-Feed Plus Novo Nordisk Humicola Animal feed
insolens
EcopulpX-200 Primalco Trichoderma Cellulose pulp bleaching
reesei
Ecosane Biotec Trichoderma Animal feed
reesei
Grindazym GP, Danisco ingredients A. niger Bird and pig feed
GV
Irgazyme 40 Nalco-Genencor, Ciba, T. Paper industry and
Geigy longibrachiatum animal feed
Solvay Solvay Enzymes T. reesei Starch and bread making
pentonase
Xylanase Seikagaku Trichoderma sp. Carbohydrate structural
studies
Xylanase GS35 Iogen T. reesei Pulp bleaching
In SSF, biomass and xylanase productions depend on several factors such as the
type of microorganism, inoculums size, temperature, oxygen diffusion, the period
of cultivation, the substrate used and its pretreatment, size, water activity, relative
humidity, pH, and many other factors. On the other hand, for SmF, the temperature,
pH, agitation, and dissolved O2 are the most critical physical parameters affecting
xylanase production in SmF. Most of the xylanase-producing microorganisms
reported so far are either mesophilic or thermophilic (Coughlan and Hazlewood
7 Microbial Xylanases: Sources, Types, and Their Applications 183
1993). It was observed that these microbes show highest xylanase activity and
optimum growth at temperature between 40 and 70 °C. However, those extremo-
philic such as Thermotoga sp., Dictyoglomus, and Geobacillus exhibit even higher
optimum temperature range of up to 80 °C (Ko et al. 2011; Sharma et al. 2007;
Qureshy et al. 2002). Xylanases of fungal origin usually are found less thermophilic
in nature. Even though some exceptional cases have been reported, the optimum
temperature for most fungal xylanases is below 55 °C. This has been one of the
major drawbacks of fungal xylanase in industrial application. On the other hand,
bacterial xylanases exhibit higher thermostability. The pH of the fermentation
media is another major factor effecting the xylanase production as well as the
enzyme stability. Fungal xylanases usually are highly alkaliphilic as they can be
active at wide ranges of pH, usually between 5.0 and 10.0, whereas the bacterial
xylanases have quite narrow pH spectrum (Dhillon et al. 2000). Aeration and
agitation rates during fermentation are the other crucial factors in case of xylanase
production process. Studies show that xylanase production by Bacillus sp. is
enhanced at an agitation rate of 200–250 rpm and 30 % and above dissolved O2
saturation (Anuradha et al. 2007; Sanghi et al. 2009). In cretin fungal species which
exhibit mycelia growth pattern, higher agitation rate results in poor growth as well
as less xylanase production due to the high sheer force. Studies made by Archana
and Satyanarayana (1998) and Sharma et al. (2007) deal with the optimization of
physical parameters effecting xylanase production.
In addition, carbon and nitrogen sources and their concentration in the fer-
mentation media are found to be not only effecting the xylanase production but also
the localization of enzymes. The fungal xylanase producers are able to utilize wide
verity lignocellulosic-derived substrates such as corn cob, corn straw, wheat bran,
rice straw, and sugarcane bagasse. For production in SSF system, these substrates
are used as they are cheaper and easily available raw materials (Chauhan et al.
2007). However, in bacterial cells in SmF, simpler and readily utilizable nutrients
are used. Organic nitrogen sources such as tryptone, peptone, yeast extract, and
soybean meal are widely used as they enhance the xylanase production as well as
biomass. Kapoor et al. (2008) observed a significant increase in xylanase produc-
tion which was observed when the Bacillus sp. SSP-34 was provided with yeast
extract in combination with peptone. In similar study, xylanase production by
Bacillus sam-3 was increased up to 25-folds higher when the inorganic nitrogen
source was replaced by soybean meal. Use of corn steep powder also significantly
enhanced xylanase production by T. reesei (Lappalainen et al. 2000). Apart from
these major nutrients, the incorporation of micronutrients such as trace elements,
vitamins, and amino acids improved the xylanase production to great extend in
cases of thermo-anaerobe-mediated xylanase fermentation. In cases of recombinant
strains, the xylanase expression greatly depends on the gene inducers present in the
fermentation media. IPTG is the most widely used chemical inducer for ‘lac’-based
expression system (Mamo et al. 2007). However, in large-scale production, IPTG is
not applicable due to its high cost.
In cases of recombinant xylanase, the expression of xylanase genes usually
depends on several aspects such as biosynthesis mechanism, gene regulation, vector
184 H.A. El Enshasy et al.
construction, expression host used, expression media, and other growth parameters
during the fermentation. For commercial applications, xylanases have to be ideally
produced quickly and in large quantities. The conventional production strategies
have been followed are found to be less effective to meet the growing need in the
present xylanase market. Therefore, many researches are going on for developing
recombinant strains for xylanase production. However, a very limited number of
these recombinant strains are studied further for the scaling up of xylanase pro-
duction in pilot scale. Hence, more investigations on this area are required and it is
very significant to have more efficient and economically viable process to reduce
the cost of xylanase production.
Media screening and medium composition optimization are usually the preliminary
steps involved in bioprocess optimization which are always carried out in shake
flask level to select the most suitable medium formulation. Both physical and
chemical parameters effecting the cell growth as well as xylanases production are
monitored and optimized at this stage with a number of experiments. Basar et al.
(2010) observed that the recombinant xylanase production in E. coli is a
growth-associated process. They observed a rapid growth of E. coli, reaching the
stationary phase after 16 h and the maximal xylanase production up to 324.72
U mL−1 was obtained in defined medium. They also reported the effects of different
concentrations of glucose in growth media on cell growth and xylanase production.
Farliahati et al. (2009) reported the xylanase expression in recombinant E. coli
DH5a in different media compositions (defined, semi-defined, and complex).
Optimization study of pH and agitation ranges also carried out in shake flask
cultivation mode. From these studies, they observed the highest xylanase produc-
tion (up to 2122 U mL−1) in defined media at initial pH 7 and agitation speed of
200 rpm. A suitable cultivation strategy for the production of two truncated ther-
mostable recombinant xylanases (XynlAN and XynlANC), isolated from
Rhodothermus marinas, was developed by Karlsson et al. (1998). They found that
the fed-batch cultivations of E. coli strain BL21 (DE3) with a controlled expo-
nential glucose feed led to high specific production of the recombinant proteins and
addition of complex nutrients such as Tryptone Soya Broth (TSB) to the media
exhibits positive impact on cell growth rate and cell productivity. Single addition of
either lactose or isopropyl-thio-go-galactoside (IPTG) was used for induction. In
lactose-induced cells, the production of recombinant xylanase was delayed for
approximately 30 min compared to those induced with IPTG, but the specific
product levels were comparable at 3 h after induction. Huang et al. (2006) cloned a
xylanase gene from B. subtilis into E. coli and found that the xylanase distributions
in extracellular, intracellular, and periplasmic fractions were in ratio 22.4, 28.0, and
49.6 %, respectively, with an optimal enzyme production at pH 6.0 and 50 °C. The
7 Microbial Xylanases: Sources, Types, and Their Applications 185
from A. usmani was cloned in pET-28a expression vector and when it is expressed
in E. coli BL21, the maximum xylanase activity was found to be 49.4 U mg−1. The
gene expression was induced by IPTG and the xylanase was found expressed with
6-X histidine tag that facilitates an easy purification of expressed protein (Zhou
et al. 2008b). A very successful cultivation strategy for the production of two
truncated thermostable recombinant xylanases (XynIAN and XynIANC), isolated
from Rhodothermus marinas, was developed by Karlsson et al. (1999). They found
that the fed-batch cultivations of E. coli strain BL21 (DE3) with a controlled
exponential glucose feed led to high specific production of the recombinant proteins
with additional complex medium. Highest xylanase production in defined medium
was achieved by Farliahati et al. (2009). In this study, xylanase production reached
up to 2122.5 U mL−1 at pH 7.4 and when (NH4)2HPO4 was used as the main
nitrogen source. They also reported that the xylanase production was a
growth-associated process as the enzyme secretion was greatly dependent on bio-
mass concentration for strain E. coli DH5a.
Biswas et al. (2010) used mutant IITD3A of Melanocarpus albomyces for the
production of high titer cellulose-free xylanase in 14-L stirred tank bioreactor where
an optimized (RSM based) production media and process parameters are used. The
authors reported the xylanase activity reached up to 415 U mL−1 after 24 h by pH
recycling strategy (between 7.8 and 8.2) during the production phase. They also
reported that the xylanase production is significantly influenced by aeration and
agitation. A 5.2-fold increase (overall volumetric productivity of 22,000 U L−1 h−1)
in activity was observed when the cells exhibited pellet form at an agitation speed
of 600 RPM.
Kumar et al. (2009) optimized the xylanase production in lab scale stirred tank
bioreactor using mycelial fungal strain T. lanuginosus MC 134. Fermentation media
contains corn cob as a sole source of carbon which was inoculated with T.
lanuginosus MC 134 and observed the production in terms of xylanase activity over
a period of 9 days at various process parameters. A highest xylanase activity, 2009
U mL−1, was obtained after 6 days of fermentation. However, a significant decrease
(40 %) in productivity was observed in bioreactors when compared to shake flask
culture. Similarly, xylanase production by several microbial strains using corn cob
was reported and from these studies it was clear that the production largely depends
on rheological factors in bioreactors and the type of extraction methods used.
Ruanglek et al. (2007) successfully cloned and expressed XylB (564 bp) encoding
endo-1,4-b-xylanase obtained from A. niger BCC14405 in Pichia pastoris under
AOX1 promoter. A maximum of 7352 U mg−1 of xylanase-specific activity was
obtained in 2 L lab scale bioreactor with BMGY/BSM medium under exponential
feeding strategy. During the batch phase which was aimed to achieve HCDC, the
temperature and pH were maintained at 30 °C and 5.0, respectively. Agitation was
maintained in the range of 600–1000 rpm. Fed-batch phase was started after 16–
20 h by feeding 67 % (v/v) glycerol plus 1 % (v/v) trace mineral solution in
exponential manner. Table 7.16 shows examples for xylanase production in lab and
pilot-scale submerged fermentation by different fungal strains.
188 H.A. El Enshasy et al.
Table 7.16 A comparison of xylanase production in lab and pilot-scale submerged fermentation
using fungal strains
Name of the Substrate Mode of Xylanase Productivity References
organism used operation activity (IU L−1 h−1)
(IU L−1)
Aspergillus niger Xylan (from Batch (20 L) 290,000 3820 Yuan et al.
corncob) (2005)
Aspergillus oryzae Spent sulfite Batch (15 L) 199,000 13,100 Chipeta
liquor et al.
(2008)
Thermomyces Oat husk Batch (2 L) 210,000 4380 Xiong
lanuginosus DSM hydrolysate et al.
10635 (2004)
Thermomyces Coarse Batch (5 L) 2,010,000 13,950 Kumar
lanuginosus MC corncob et al.
134 (2009)
Trichoderma Oat husk Fed-batch (5 1,350,000 14,060 Xiong
reesei hydrolysate L) et al.
(2005)
Melanocarpus Soluble Batch with 550,000 22,000 Biswas
albomyces wheat straw pH cycling et al.
extract (14 L) (2010)
• Presence of disulfide bridges which may strengthen the whole protein frame-
work would give more stability to the enzyme.
• Presence of ion and aromatic pairs and their interactions may favor to more
stable framework.
• Strategic position of water molecules on the framework, since the readily
available water bindings are required for the catalytic activity even at higher
temperature.
• Complex folding nature of enzyme. The higher the folding complexity, the
higher the stability.
• Occurrence of oligomerization and the presence of certain aromatic residues
result in the hydrophobic patches which allows stronger monomeric interactions
enhancing the enzyme stability.
Certain GH11 endo-1,4-b-xylanases display disulfide bridges at position
between b-strand and a-helix. However, these enzymes do not exhibit higher
thermostability, whereas endo-1,4-b-xylanase produced by Dictyoglomus ther-
mophilum exhibits thermostability even though that do not have any disulfide
bridges on their structure. Therefore, the presence of disulfide bridges does not
necessarily account for the thermophilic characters of the enzymes (Table 7.17).
Industrial production of xylanases has been grown significantly during last few
decades based on their increased application. Xylan can be converted into many
useful products such as xylitol and furfurol through enzymatic hydrolysis
(Subramaniyan and Prema 2002). The hydrolysis could be also carried out through
acid treatment process which is relatively easy, fast, and cheap. However, the
application of this method in industries is limited due to the formation of toxic end
194 H.A. El Enshasy et al.
substances that may have cross reactions with product of interest (Beg et al. 2001).
Therefore, enzymatic hydrolysis method is widely applied even though it is a slow
process but more efficient and eco-friendly. During industrial application, the
xylolytic enzymes are not commonly applied in pure form but in combination with
other hydrolytic enzymes to achieve better results. The industrial applications of
xylanases cover different areas and the list of applications increased by time and
thus increases the market demand of xylanases accordingly.
Xylanases are widely used in animal feed manufacturing. They are usually applied
in combination of other hydrolytic enzymes such as pectinase, protease, cellulose,
lipase, glucanase, alginase, and phytase. These enzymes are used to reduce the
viscosity of raw materials used in animal feeds. The complex arabinoxylan that is
present on the cell wall of most grains may reduce the effective digestion and
absorption of nutrients in poultry. Studies on this area have reported that the
xylanase-treated feed has better digestibility and nutrient absorption rate when
compared to raw feeds (Bedford and Classen 1992). In animal feed industries, the
feed conversion rate (FCR) is defined as the ratio of feed consumed to the body
weight gained by the animal. A lower FCR means improved feed digestibility and
nutrient absorption. Xylanase enzymes are widely incorporated in poultry feed to
reduce the FCR. It have been reported that the addition of endo-xylanase is isolated
from Acidothermus cellulolyticus in poultry feed and observed a significant
decrease in FCR, indicating an improved digestibility of feed (Clarkson et al. 2001).
Similar results were observed when endo-xylanase is isolated from Neocallimastix
patriciarum added to canola seed after oil extraction and applied for cattle feeding.
The digestibility and stability of nutritive value of silage for cattle are found to be
enhanced when it was treated with xylanase along with probiotic lactobacillus. The
xylose formed by the de-polymerization of hemicellulose was fermented by the
probiotic lactobacillus and thus enhanced the feed nutritive value. It was also
reported that xylanase supplementation in the animal diet improved the animal
growth through the better nutrition absorbability and enhancement of immune
response as well as the gut micro-flora (Gao et al. 2008). In other study, Vandeplas
et al. (2010) reported that when T. viride xylanase is incorporated in poultry feed, it
increased nutrient digestion by 60 % and resulted in a significant increase in
chicken weight. Pretreatment of forage crops with xylanase leads to better digestion
in cattle and these enzymes are often found in synergy with other cellulolytic
enzymes. The assimilation of ruminant feed was also improved many folds by
xylanases addition (Yang and Xie 2010).
7 Microbial Xylanases: Sources, Types, and Their Applications 195
Food and beverages industries use also xylanases in wide range of applications.
Xylanases are capable of hydrolyzing the hemicellulose in wheat that enhances the
water absorption of dough at leavening level in bread making process. This could
possibly enhance the fermentation rate and as a result the softness of the bread and
volume was increased (Rouau 1993). Xylanase cocktails are often used to reduce
the stickiness of the dough thus enhance the crumb (Goesaert et al. 2005; Butt et al.
2008). Other study showed also that under specific process condition, incorporation
of xylanase in the dough increased the shelf life of bread and also improved the
volume up to 24 % (Verjans et al. 2010). However, enzyme selection is critical in
these conditions since the cereals used for bread making often exhibit the xylanase
inhibition by TAXI, XIP, or TLXI (Gebruers et al. 2004). Pasta processing is
another area where xylanases are widely used as solubilizing agent of durum wheat
semolina without changing its rheological properties (Ingelbrecht et al. 2001; Brijs
et al. 2004). It was also reported that xylanase enzymes can improve the separation
of wheat flour into starch and gluten (Frederix et al. 2004). However, in certain
cases, action of Xyl-11 does not favor the separation due to high viscosity caused
by the un-extractable arabino xylanase hydrolysis. In such conditions, synergetic
dosages of Xyl-10 and Xyl-11 are used as it can mediate-improved agglomeration
of gluten protein even though the wheat quality is very poor (Van Der Borght et al.
2005). Table 7.18 shows some examples of commonly used xylanase preparations
for the baking industry. Xylanases are widely used also in juice, beer, and wine
making industries. The enzymatic breakdown of fruits and vegetables hemicellulose
can increase the clarity and stability of juice and wine. Xylanases can be used along
with several other enzymes such as pectinase and amylase to improve the quality of
beverage products in terms of aroma, taste, clarity, and stability (Wong et al. 1988).
It was reported that up to 27 % decrease in insoluble fibers was achieved in juice
clarification by addition of xylanase enzyme isolated from Sclerotinia sclerotiorum
(Olfa et al. 2007). In beer production, partial hydrolysis of arabinoxylan found in
the barley cell wall gives a middy viscous appearance to the fermented beer. In such
Table 7.18 Examples of commonly used xylanase preparations for the baking industry (Collins
et al. 2005a, b)
Supplier Brand Enzyme source
Beldem-Puratos Bel’ase B210 Bacillus subtilis
Bel’ase F25 Aspergillus niger
Genencor Intl. Multifectw Trichoderma reesei
Danisco Grindamyle H Aspergillus niger
DSM Bakezymew HS, BXP Aspergillus niger
Novozymes Pentopan Mono Thermomyces lanuginosus
AB enzymes Veronw 191 Aspergillus niger
Veronw Special Bacillus subtilis
196 H.A. El Enshasy et al.
cases, various xylanase enzyme groups are used to increase the filtration perfor-
mance of beer wort by hydrolysis of viscous AXs fractions. Xylanase inhibitors
present in the malt are the major hinderance to xylanase action. The xylanase
activity is retained with the help of other synergetic enzymes used in the process
(Debyser et al. 1997).
Endo-1,4-b-xylanase enzymes play significant role in paper and pulp industry since
they facilitate an eco-friendly bio-bleaching of wood pulp by reducing the use of
chemical bleaching agents such as chlorine and chlorine dioxide. For the past few
decades, kraft pulp bleaching industry is recognized as the biggest area of xylanase
application. Several studies reported that the removal of lignin content can be
achieved by xylanase enzymes far better than by other lignin-degrading enzymes
(Lundgren et al. 1994). Use of chemical bleaching agents in conventional pulp
bleaching methods cause serious environmental pollution issues and this drove the
researcher’s attention on xylanase enzymes (Subramaniyam and Prema 2002).
Table 7.19 shows major endo-1,4-b-xylanase supplier in pulp bleaching industry.
Application of endo-1,4-b-xylanase enzymes in pulp bleaching can reduce using
chlorine as well as the total energy requirement for this process. Kulkarni et al.
(1999) observed that xylanase pretreatment of kraft pulp lowers the usage of
chemicals by 10–20 %. The concept of TCF (total chlorine free) bleaching tech-
nologies was also recently introduced to make the process more eco-friendly.
However, cellulase-free endo-1,4-b-xylanase is required in this process since the
cellulose enzyme may degrade the cellulose which is the main target of the industry
Table 7.19 Major xylanase supplier in pulp bleaching industry (Beg et al. 2001)
Supplier Product trade name
Alko Rajamaki, Finland Ecopulp
Sandoz, Charlotte, N.C. and Basle, Switzerland Cartazyme
Clarient, UK Cartazyme HS, HT and SR 10
Genercor, Finland; Ciba Giegy, Switzerland Irgazyme 40–4X/Albazyme 40–4X
Novo Nordisk, Denmark Pulpzyme HA, HB and HC
Biocon India, Bangalore Bleachzyme F
Rohn Enzyme 0Y; Primalco, Finland Ecopulp X-100, X-200, and X-200/4
Solvay Interox, USA Optipulp L-8000
Thomas Swan, UK Ecozyme
Iogen, Canada GS-35, HS70
7 Microbial Xylanases: Sources, Types, and Their Applications 197
(Bajpai and Bajpai 2001). Relatively lower molecular weight of xylanase enzymes
helps them to penetrate more deep into the lignocelluloses substrates and facilitate
better hydrolytic actions (Beaugrand et al. 2005; Beg et al. 2001). Higher ther-
mostability and alkaline stability are other key features that make xylanase more
favorable for bio-bleaching industry (Buchert et al. 1995). The mechanisms of pulp
bleaching by xylanase have been studied and many researchers believe that xyla-
nase hydrolysis takes place at lignin xylan interaction and thus results in a loosely
packed cellulosic framework (Shatalov and Pereira 2007). However, it was recently
reported the potential use of Aspergillus oryzae MDU-4-based fungal xylanase for
effective deinking of newspaper pulp where a significant improvement in optical
properties such as brightness up to 57.9 % ISO with decreased (211 ppm) residual
ink concentration (Chutani and Sharma 2015).
Apart from the traditional applications in pulp beaching industry, xylanases have
been used in several other areas as well. The broad spectrum of xylanase application
includes the saccharification of xylan to xylose, plant cell protoplasting, coffee
198 H.A. El Enshasy et al.
References
Abou Hachem M, Karlsson EN, Bartonek-Roxa E, Raghothama S, Simpson PJ, Gilbert HJ,
Williamson MP, Holst O (2000) Carbohydrate-binding modules from a thermostable
Rhodothermus marinus xylanase: cloning, expression and binding studies. Biochem J
345:53–60
Alejandro SH, Jesús VE, del María CMH, María EHL (2007) Purifcation and characterization of
two sugarcane bagasse-absorbable thermophilic xylanases from the mesophilic Cellulomonas
Xavigena. Microbiol Biotechnol 34:331–338
Akila G, Chandra TS (2002) A novel cold-tolerant Clostridium strain PXYL1 isolated from a
psychrophilic cattle manure digester that secretes thermolabile xylanase and cellulase. FEMS
Microbiol Lett 219:63–67
Andrade SV, Polizeli MLTM, Terenzi HF, Jorge JA (2004) Effect of carbon source on the
biochemical properties of b-xylosidases produced by Aspergillus versicolor. Process Biochem
39:931–938
Andre-Leroux G, Berrin JG, Georis J, Arnaut V, Juge V (2008) Structure-based mutagenesis of
Penicillium griseofulvum xylanase using computational design. Proteins 72(4):1298–1307
Andrews SR, Charnock SJ, Lakey JH, Davies GJ, Claeyssens M, Nerinckx V, Underwood M,
Sinnott ML, Warren RA, Gilbert HJ (2000) Substrate specificity in glycoside hydrolase family
10. Tyrosine 87 and leucine 314 play a pivotal role in discriminating between glucose and
xylose binding in the proximal active site of Pseudomonas cellulosa xylanase 10A. J Biol
Chem 275(30):23027–23033
Anuradha P, Vijayalakshmi K, Prasanna ND, Sridevi K (2007) Production and properties of
alkaline xylanases from Bacillus sp. isolated from sugarcane fields. Curr Sci 92(9):1283–1286
Arase A, Yomo T, Urabe I, Hata Y, Katsube Y, Okada H (1993) Stabilization of xylanase by
random mutagenesis. FEBS Lett 316(2):123–127
Archana A, Satyanarayana T (1998) Cellulase-free xylanase production by thermophilic Bacillus
licheniformis A99. Indian J Microbiol 38:135–139
Atkins EDT (1992) Three-dimensional structure, interactions and properties of xylans. Xylanas
Xylanases 7:39–50
Avalos OP, Noyola TP, Plaza IM, Torre M (1996) Induction of xylanase and b-xylosidase in
Cellulomonas flavigena growing on different carbon sources. Appl Microbiol Biotchnol
46:405–409
Bai Y, Wang J, Zhang Z, Yang P, Shi P, Luo H, Meng K, Huang H, Yao B (2010) A new xylanase
from thermoacidophilic Alicyclobacillus sp. A4 with broad-range pH activity and pH stability.
J Ind Microbiol Biotechnol 37(2):187–194
Bajpai P, Bajpai PK (2001) Development of a process for the production of dissolving kraft pulp
using xylanase enzyme. Appita J 54(4):381–384
Bajaj BK, Singh NP (2010) Production of xylanase from an alkali tolerant Streptomyces sp. 7b
under solid-state fermentation, its purification, and characterization. Appl Biochem Biotechnol
162(6):1804–1818
Balakrishnan H, Srinivasan MC, Rele MV, Chaudhari K, Chandwadkar AJ (2000) Effect of
synthetic zeolites on xylanase production from an alkalophilic Bacillus sp. Curr Sci 79:95–97
Basar B, Shamzi MM, Rosfarizan M, Puspaningsih NNT, Ariff AB (2010) Enhanced production
of thermophilic xylanase by recombinant Escherichia coli DH5a through optimization of
medium and dissolved oxygen level. Int J Agric Biotechnol 12:321–328
Basaran P, Hang YD, Basaran N, Worobo RW (2001) Cloning and heterologous expression of
xylanase from Pichia stipitis in Escherichia coli. J Appl Microbiol 90(2):248–255
Battan B, Dhiman SS, Ahlawat S, Mahajan R, Sharma J (2012) Application of thermostable
xylanase of Bacillus pumilus in textile processing. Indian J Microbiol 52(2):222–229
Beaugrand J, Paës G, Reis D, Takahashi M, Debeire P, O’Donohue MJ, Chabbert B (2005)
Probing the cell wall heterogeneity of micro-dissected wheat caryopsis using both active and
inactive forms of a GH11 xylanase. Planta 222:246–257
200 H.A. El Enshasy et al.
Bedford MR, Classen HL (1992) Reduction of intestinal viscosity through manipulation of dietary
rye and pentosanase concentration is effected through changes in the carbohydrate composition
of the intestinal aqueous phase and results in improved growth rate and food conversion
efficiency of broiler chicks. J Nutr 122:560–569
Beg QK, Bhushan B, Kappor M, Hoondal G (2000) Production and characterization of
thermostable xylanase and pectinase from Streptomyces sp. QG-11-3. J Ind Microbiol
Biotechnol 24:396–402
Beg QK, Kapoor M, Mahajan L, Hoondal GS (2001) Microbial xylanases and their industrial
applications: a review. Appl Microbiol Biotechnol 56(3–4):326–338
Bergquist P, Te’o V, Gibbs M, Cziferszky A, de Faria FP, Azevedo M, Nevalainen H (2002)
Expression of xylanase enzymes from thermophilic microorganisms in fungal hosts.
Extremophiles 6(3):177–184
Berrin JG, Juge N (2008) Factors affecting xylanase functionality in the degradation of
arabinoxylans. Biotechnol Lett 30(7):1139–1150
Berrin JG, Williamson G, Puigserver A, Chaix JC, McLauchlan WR, Juge N (2000) High-level
production of recombinant fungal endo-beta-1,4-xylanase in the methylotrophic yeast Pichia
pastoris. Protein Expr Purif 19(1):179–187
Bhalerao J, Patki AH, Bhave V, Khurana I, Deobagkar DN (1990) Molecular-cloning and
expression of a xylanase gene from Cellulomonas sp. into Escherichia coli. Appl Microbiol
Biotechnol 34(1):71–76
Bhardwaj A, Leelavathi S, Mazumdar-Leighton S, Ghosh A, Ramakumar S, Reddy VS (2010) The
critical role of N- and C-terminal contact in protein stability and folding of a family 10
xylanase under extreme conditions. PLoS ONE 5(6):e11347
Biely P, Vrsanska M, Kluepfel M, Tenkanen D (1997) Endo-beta-1,4-xylanase families:
differences in catalytic properties. J Biotechnol 57(1–3):151–166
Biely P, Leathers TD, Cziszarova M, Vrsanska M, Cotta MA (2008) Endo-beta-1,4-xylanase
inhibitors in leaves and roots of germinated maize. J Cereal Sci 48(1):27–32
Biswas R, Sahai V, Mishra S, Bisaria VS (2010) Bioprocess strategies for enhanced production of
xylanase by Melanocarpus albomyces IITD3A on agro-residual extract. J Biosci Bioeng
110:702–708
Blum DL, Li XL, Chen H, Ljungdahl LG (1999) Characterization of an acetyl xylan esterase from
the anaerobic fungus Orpinomyces sp. strain PC-2. Appl Environ Microbiol 65(9):3990–3995
Bocchini DA, Alves-Prado HF, Baida LC, Roberto IC, Gomes E, Da-Silva R (2002) Optimization
of xylanase production by Bacillus circulans D1 in submerged fermentation using response
surface methodology. Process Biochem 38:727–731
Bolam DN, Xie HF, White P, Simpson PJ, Hancock SM, Williamson MP, Gilbert HJ (2001)
Evidence for synergy between family 2b carbohydrate binding modules in Cellulomonas fimi
xylanase 11A. Biochemistry 40(8):2468–2477
Boraston AB, Tomme P, Amandoron EA, Kilburn DG (2000) A novel mechanism of xylan
binding by a lectin-like module from Streptomyces lividans xylanase 10A. Biochem J 350(Pt
3):933–941
Boraston AB, McLean BW, Chen V, Li A, Warren RA, Kilburn DG (2002) Co-operative binding
of triplicate carbohydrate-binding modules from a thermophilic xylanase. Mol Microbiol 43
(1):187–194
Borralho T, Chang Y, Jain P, Lalani M, Parghi K (2002) Lactose induction of the lac operon in
Escherichia coli B23 and its effect on the o-nitrophenyl b-galactoside assay. J Exp Microbiol
Immunol 2:117–123
Bridgeman TG, Jones JM, Shield I, Williams PT (2008) Torrefaction of reed canary grass, wheat
straw and willow to enhance solid fuel qualities and combustion properties. Fuel 87
(6):844–856
Brijs K, Ingelbrecht JA, Courtin CM, Schlichting L, Marchylo BA, Delcour JA (2004) Combined
effects of endoxylanases and reduced water levels in pasta production. Cereal Chem
81:361–368
7 Microbial Xylanases: Sources, Types, and Their Applications 201
Brutus A, Villard C, Durand A, Tahir T, Furniss C, Puigserver A, Juge N, Giardina T (2004) The
inhibition specificity of recombinant Penicillium funiculosum xylanase B towards wheat
proteinaceous inhibitors. Biochim Biophys Acta 1701(1–2):121–128
Brutus A, Reca IB, Herga S, Mattei B, Puigserver A, Chaix JC, Juge V, Bellincampi D, Giardina T
(2005) A family 11 xylanase from the pathogen Botrytis cinerea is inhibited by plant
endoxylanase inhibitors XIP-I and TAXI-I. Biochem Biophys Res Commun 337(1):160–166
Buchert J, Teleman A, Harjunpa V, Tenkanen M, Viikari L, Vuorinen T (1995) Effect of cooking
and bleaching on the structure of xylan in conventional pine kraft pulp. Tappi J 78
(11):125–130
Butt MS, Nadeem MT, Ahmad Z, Sultan MT (2008) Xylanases and their applications in baking
industry. Food Technol Biotechnol 46(1):22–31
Cannio R, Di Prizito N, Rossi M, Morana A (2004) A xylan-degrading strain of Sulfolobus
solfataricus: isolation and characterization of the xylanase activity. Extremophiles 8
(2):117–124
Cantarel BL, Coutinho PM, Rancurel C, Bernard T, Lombard V, Henrissat B (2009) The
carbohydrate-active enzymes database (CAZy): an expert resource for glycogenomics. Nucleic
Acids Res 37:D233–D238
Cao Y, Qiao J, Li Y, Lu W (2007) De novo synthesis, constitutive expression of Aspergillus
sulphureus beta-xylanase gene in Pichia pastoris and partial enzymic characterization. Appl
Microbiol Biotechnol 76(3):579–585
Carmona EC, Brochetto-braga MR, Pizzirani-kleiner AA, Jorge JA (1998) Purification and
biochemical characterization of an endoxylanase from Aspergillus versicolor. FEMS Microbiol
Lett 166:311–315
Carmona EC, Fialho MB, Buchgnani EB, Coelho GD, Brocheto-Braga MR, Jorge JA (2005)
Production, purification and characterization of a minor form of xylanase from Aspergillus
versicolor. Process Biochem 40(1):359–364
Cereghino JL, Cregg JM (2000) Heterologous protein expression in the methylotrophic yeast
Pichia pastoris. FEMS Microbiol Rev 24(1):45–66
Cervera-Tison MC, Andre-Leroux G, Lafond M, Georis J, Juge N, Berrin J-G (2009) Molecular
determinants of substrate and inhibitor specificities of the Penicillium griseofulvum family 11
xylanases. Biochim Biophys Acta 1794:438–445
Charnock SJ, Lakey JH, Virden R, Hughes N, Sinnott ML, Hazlewood GP, Pickersgill R,
Gilbert HJ (1997) Key residues in subsite F play a critical role in the activity of Pseudomonas
fluorescens subspecies cellulosa xylanase A against xylooligosaccharides but not against
highly polymeric substrates such as xylan. J Biol Chem 272(5):2942–2951
Chauhan K, Trivedi U, Patel KC (2007) Statistical screening of medium components by Plackett–
Burman design for lactic acid production by Lactobacillus sp. KCP01 using date juice.
Bioresour Technol 98:98–103
Chauthaiwale VM, Deshpande VV (1992) Molecular cloning and expression of the xylanase gene
from Chainia in Escherichia coli. FEMS Microbiol Lett 99(2–3):265–270
Chavez R, Schachter K, Navarro C, Peirano A, Bull P, Eyzaguirre J (2004) The acetyl xylan
esterase II gene from Penicillium purpurogenum is differentially expressed in several carbon
sources, and tightly regulated by pH. Biol Res 37(1):107–113
Chutani P and Sharma KK (2015) Biochemical evaluation of xylanases from various filamentous
fungi and their application for the deinking of ozone treated newspaper pulp. Carbohydrate
Polymers 127:54–63
Cheng YF, Yang CH, Liu WH (2005) Cloning and expression of Thermobifida xylanase gene in
the methylotrophic yeast Pichia pastoris. Enz Microb Technol 37(5):541–546
Chiku K, Uzawa J, Seki H, Amachi S, Fujii T, Shinoyama H (2008) Characterization of a novel
polyphenol-specific oligoxyloside transfer reaction by a family 11 xylanase from Bacillus
sp. KT12. Biosci Biotechnol Biochem 72(9):2285–2293
Chipeta ZA, Du-Preez JC, Christopher L (2008) Effect of cultivation pH and agitation rate on
growth and xylanase production by Aspergillus oryzae in spent sulphite liquor. J Ind Microbiol
Biotechnol 35:587–594
202 H.A. El Enshasy et al.
Dean JFD, Gross KC, Anderson JD (1991) Ethylene biosynthesis-inducing xylanase. 3. Product
characterization. Plant Physiol 96(2):571–576
Debyser W, Peumans WJ, Van Damme EJM, Delcour JA (1999) Triticum aestivum xylanase
inhibitor (TAXI), a new class of enzyme inhibitor affecting breadmaking performance. J Cereal
Sci 30(1):39–43
Debyser W, Derdelinckx G, Delcour JA (1997) Arabinoxylan solubilization and inhibition of the
barley malt xylanolytic system by wheat during mashing with wheat whole meal adjunct:
evidence for a new class of enzyme inhibitors in wheat. Am Soc Brew Chem 55:153–156
Degefu Y, Lohtander K, Paulin L (2004) Expression patterns and phylogenetic analysis of two
xylanase genes (htxyl1 and htxyl2) from Helminthosporium turcicum, the cause of northern
leaf blight of maize. Biochimie 86(2):83–90
Dey D, Hinge J, Shendye A, Rao M (1992) Purification and properties of extracellular
endoxylanases from alkalophilic thermophilic Bacillus sp. Can J Microbiol 38:436–442
Dheeran P, Nandhagopal N, Kumar S, Jaiswal YK, Adhikari DK (2012) A novel thermostable
xylanase of Paenibacillus macerans IIPSP3 isolated from the termite gut. J Ind Microbiol
Biotechnol 39(6):851–860
Dhillon A, Gupta JK, Khanna S (2000) Enhanced production, purification and characterisation of a
novel cellulase-poor thermostable, alkalitolerant xylanase from Bacillus circulans AB 16.
Process Biochem 35(8):849–856
Dhiman SS, Sharma J, Battan B (2008) Industrial applications and future prospects of microbial
xylanases: a review. BioResources 3(4):1377–1402
Dijkerman R, Ledeboer J, Op den Camp HJM, Prins RA, van der Drift C (1997) The anaerobic
fungus Neocallimastix sp. strain L2: growth and production of (Hemi) cellulolytic enzymes on
a range of carbohydrate substrates. Curr Microbiol 34(2):91–96
Dornez E, Croes E, Gebruers K, De Coninck B, Cammue BPA, Delcour JA, Courtin CM (2010a)
Accumulated evidence substantiates a role for three classes of wheat xylanase inhibitors in
plant defense. Crit Rev Plant Sci 29(4):244–264
Dornez E, Croes E, Gebruers K, Carpentier S, Swennen R, Laukens K, Witters E, Urban M,
Delcour JA, Courtin CM (2010b) 2-D DIGE reveals changes in wheat xylanase inhibitor
protein families due to Fusarium graminearum DeltaTri5 infection and grain development.
Proteomics 10(12):2303–2319
Durand A, Hughes R, Roussel A, Flatman R, Henrissat B, Juge N (2005) Emergence of a
subfamily of xylanase inhibitors within glycoside hydrolase family 18. FEBS J 272
(7):1745–1755
Ebanks R, Dupont M, Shareck F, Morosoli R, Kluepfel D, Dupont C (2000) Development of an
Escherichia coli expression system and thermostability screening assay for libraries of mutant
xylanase. J Ind Microbiol Biotechnol 25(6):310–314
El Enshasy H, Hellmuth K, Rinas U (1999) Fungal morphology in submerged cultures and its
relation to glucose oxidase excretion by recombinant Aspergillus niger. Appl Biochem
Biotechnol 81:1–11
El Enshasy H, Kleine J, Rinas U (2006) Agitation effects on morphology and protein productive
fractions of filamentous and pelleted growth forms of recombinant Aspergillus niger. Process
Biochem 41:2103–2112
El Enshasy H (2007) Filamentous fungal culture—process characteristics, products, and
applications. In: Yang ST (ed) Bioprocessing for value-added products from renewable
resources. Elsevier Press, Amsterdam. ISBN-10: 0-444-52114-3
Emami K, Hack E (2001) Characterisation of a xylanase gene from Cochliobolus sativus and its
expression. Mycol Res 105:352–359
Esteban R, Villanueva JR, Villa TG (1982) b-D-xylanases of Bacillus circulans WL-12. Can J
Microbiol 28:733–739
Fang HY, Chang SM, Lan CH, Fang TJ (2008) Purification and characterization of a xylanase
from Aspergillus carneus M34 and its potential use in photoprotectant preparation. Process
Biochem 43(1):49–55
204 H.A. El Enshasy et al.
Georis J, Giannotta F, De Buyl E, Granier B, Frère J (2000) Purification and properties of three
endo-b-1,4-xylanases produced by Streptomyces sp. strain S38 which differ in their ability to
enhance the bleaching of kraft pulps. Enzyme Microb Technol 26:178–186
Gomes DJ, Gomes J, Steiner W (1994) Factors influencing the induction of endo-xylanase by
Thermoascus aurantiacus. J Biotechnol 33(1):87–94
Gomez LD, Steele-King CG, McQueen-Mason SJ (2008) Sustainable liquid biofuels from
biomass: the writing’s on the walls. New Phytol 178(3):473–485
Gruber K, Klintschar G, Hayn M, Schlacher V, Steiner W, Kratky C (1998) Thermophilic
xylanase from Thermomyces lanuginosus: high-resolution X-ray structure and modeling
studies. Biochemistry 37(39):13475–13485
Guo B, Chen XL, Sun CY, Zhou BC, Zhang YZ (2009) Gene cloning, expression and
characterization of a new cold-active and salt-tolerant endo-beta-1,4-xylanase from marine
Glaciecola mesophila KMM 241. Appl Microbiol Biotechnol 84(6):1107–1115
Gupta S, Bhushan B, Hoondal GS (2000) Isolation, purification and characterization of xylanase
from Staphylococcus sp. SG-13 and its application in biobleaching of kraft pulp. J Appl
Microbiol 88:325–334
Gupta U, Kar R (2009) Xylanase production by a thermos-tolerant Bacillus species under
solid-state and submerged fermentation. Braz Arch Biol Technol 52:1363–1371
Hakulinen N, Tenkanen M, Rouvinen J (1998) Crystallization and preliminary X-ray diffraction
studies of the catalytic core of acetyl xylan esterase from Trichoderma reesei. Acta
Crystallogr D Biol Crystallogr 54(Pt 3):430–443
Henrissat B, Bairoch A (1993) New families in the classification of glycosyl hydrolases based on
amino acid sequence similarities. Biochem J 293(3):781–788
Hessing JGM, van Rotterdam C, Verbakel JMA, Roza M, Maat J, van Gorcom RFM, van den
Hondel CAMJJ (1994) Isolation and characterization of a 1,4-b-endoxylanase gene of A.
awamori. Curr Genet 26:228–232
Himmel ME, Ding SY, Johnson DK, Adney WS, Nimlos MR, Brady JW, Foust TD (2007)
Biomass recalcitrance: engineering plants and enzymes for biofuels production. Science 315
(5813):804–807
Honda Y and Kitaoka M (2004) A family 8 glycoside hydrolase from Bacillus halodurans C-125
(BH2105) is a reducing end xylose-releasing exo-oligoxylanase. J Biol Chem 279:55097–55103
Hu YJ, Smith DC, Cheng KJ, Foresberg CW (1991) Cloning of a xylanase gene from Fibrobacter
succinogenes 135 and its expression in Escherichia coli. Can J Microbiol 37(7):554–561
Huang J, Wang G, Xiao L (2006) Cloning, sequencing and expression of the xylanase gene from a
Bacillus subtilis strain B10 in Escherichia coli, Bioresour Technol 97:802–808
Humphry DR, GeorgeA Black GW, Cummings SP (2001) Flavobacterium frigidarium sp. an
aerobic, psychrophilic, xylanolytic and laminarinolytic bacterium from Antarctica. Int J Syst
Evol Microbiol 51:1235–1243
Hung KS, Liu SM, Fang TY, Tzou WS, Lin FP, Sun KH, Tang SJ (2011) Characterization of a
salt-tolerant xylanase from Thermoanaerobacterium saccharolyticum NTOU1. Biotechnol Lett
33(7):1441–1447
Igawa T, Tokai T, Kudo T, Yamaguchi I, Kimura M (2005) A wheat xylanase inhibitor gene,
Xip-I, but not Taxi-I, is significantly induced by biotic and abiotic signals that trigger plant
defense. Biosci Biotechnol Biochem 69(5):1058–1063
Ingelbrecht JA, Moers K, Abecassis J, Rouau X, Delcour JA (2001) Influence of arabinoxylans
and endoxylanases on pasta processing and quality. Production of high-quality pasta with
increased levels of soluble fiber. Cereal Chem 78:721–729
Ito K, Ikemasu T, Ishikawa T (1992) Cloning and sequencing of the xynA gene encoding xylanase
A of Aspergillus kawachii. Biosci Biotechnol Biochem 56(6):906–912
Jeong MY, Kim S, Yun CW, Choi YJ, Cho SG (2007) Engineering a de novo internal disulfide
bridge to improve the thermal stability of xylanase from Bacillus stearothermophilus No. 236.
J Biotechnol 127(2):300–309
206 H.A. El Enshasy et al.
Joo JC, Pack SP, Kim YH, Yoo YJ (2011) Thermostabilization of Bacillus circulans xylanase:
computational optimization of unstable residues based on thermal fluctuation analysis.
J Biotechnol 151(1):56–65
Joshi MD, Hedberg A, McIntosh LP (1997) Complete measurement of the pKa values of the
carboxyl and imidazole groups in Bacillus circulans xylanase. Protein Sci 6(12):2667–2670
Jun H, Bing Y, Zhang K, Ding X, Daiwen C (2008) Expression of a Trichoderma reesei
B-xylanase gene in Escherichia coli and activity of the enzyme on fiber-bound substrates.
Protein Expr Purif 67:1–6
Kapoor M, Nair LM, Kuhad RC (2008) Cost-effective xylanase production from free and
immobilized Bacillus pumilus strain MK001 and its application in saccharification of Prosopis
juliflora. Biochem Eng J 38:88–97
Karlsson EN, BartonekRoxa E, Holst O (1997) Cloning and sequence of a thermostable
multidomain xylanase from the bacterium Rhodothermus marinus. Biophys Acta Gene Struct
Expr 1353(2):118–124
Karlsson EN, Bartonek Roxa E, Holst O (1998) Evidence for substrate binding of a recombinant
thermostable xylanase originating from Rhodothermus marinus. FEMS Microbiol Lett 168
(1):1–7
Karlsson EN, Holst O, Tocaj A (1999) Efficient production of truncated thermostable xylanases
from Rhodothermus marinus in Escherichia coli fed-batch cultures. J Biosci Bioeng
87:598–606
Katapodis P, Christakopoulos P (2008) Enzymatic production of feruloyl xylo-oligosaccharides
from corn cobs by a family 10 xylanase from Thermoascus aurantiacus. LWT Food Sci
Technol 41(7):1239–1243
Khandeparkar R, Bhosle NB (2006) Purification and characterization of thermoalkalophilic
xylanase isolated from the Enterobacter sp. MTCC 5112. Res Microbiol 157(4):315–325
Khasin A, Alchanati I, Shoham Y (1993) Purification and characterization of a thermostable
xylanase from Bacillus stearothermophilus T-6. Appl Environ Microbiol 59(6):1725–1730
Kimura T, Kitamoto N, Kito Y, Karita S, Sakka K, Ohmiya K (1998) Molecular cloning of
xylanase gene xynG1 from Aspergillus oryzae KBN 616, a shoyu koji mold, and analysis of its
expression. J Ferm Bioeng 85(1):10–16
Kimura T, Ito J, Kawano A, Makino T, Kondo H, Karita S, Sakka K, Ohmiya K (2000)
Purification, characterization, and molecular cloning of acidophilic xylanase from Penicillium
sp. 40. Biosci Biotechnol Biochem 64(6):1230–1237
Kitamoto N, Yoshino S, Ohmiya K, Tsukagoshi N (1999) Purification and characterization of the
over expressed Aspergillus oryzae xylanase, XynF1. Biosci Biotechnol Biochem 63
(10):1791–1794
Ko EP, Akatsuka H, Moriyama H, Shinmyo A, Hata Y, Katsube Y, Urabe I, Okada H (1992)
Site-directed mutagenesis at aspartate and glutamate residues of xylanase from Bacillus
pumilus. Biochem J 288(Pt 1):117–121
Ko CH, Tsai CH, Tu J, Yang BY, Hsieh DL, Jane WN, Shih TL (2011) Identification of
Paenibacillus sp. 2S-6 and application of its xylanase on biobleaching. Int Biodeterior
Biodegrad 65:334–339
Kohli U, Nigam P, Singh D, Chaudhary K (2001) Thermostable, alkalophilic and cellulase free
xylanase production by Thermoactinomyces thalophilus subgroup C. Enzyme Microb Technol
28(7–8):606–610
Kolenova K, Vrsanska M, Biely P (2006) Mode of action of endo-beta-1,4-xylanases of families
10 and 11 on acidic xylooligosaccharides. J Biotechnol 121:338–345
Kormelink FJ, Gruppen H, Vietor RJ, Voragen AG (1993) Mode of action of the xylan-degrading
enzymes from Aspergillus awamori on alkali-extractable cereal arabinoxylans. Carbohydr Res
249(2):355–367
Krengel U, Dijkstra BW (1996) Three-dimensional structure of Endo-1,4-beta-xylanase I from
Aspergillus niger: molecular basis for its low pH optimum. J Mol Biol 263(1):70–78
Kudo T, Ohkoshi A, Horikoshi K (1985) Molecular cloning and expression of a xylanase gene of
alkalophilic Aeromonas sp. no. 212 in Escherichia coli. J Gen Microbiol 131(10):2825–2830
7 Microbial Xylanases: Sources, Types, and Their Applications 207
Kui H, Luo H, Shi P, Bai Y, Yuan T, Wang Y, Yang P, Dong S, Yao B (2010) Gene cloning,
expression, and characterization of a thermostable xylanase from Nesterenkonia xinjiangensis
CCTCC AA001025. Appl Biochem Biotechnol 162(4):953–965
Kulkarni N, Lakshmikumaran M, Rao M (1999) Xylanase II from an alkaliphilic thermophilic
Bacillus with a distinctly different structure from other xylanases: evolutionary relationship to
alkaliphilic xylanases. Biochem Biophys Res Commun 263(3):640–645
Kumar KS, Manimaran A, Permaul K, Singh S (2009) Production of b-xylanase by a
Thermomyces lanuginosus MC 134 mutant on corn cobs and its application in biobleaching of
bagasse pulp. J Biosci Bioeng 107(5):494–498
Kumar PR, Eswaramoorthy S, Vithayathil PJ, Viswamitra MA (2000) The tertiary structure at 1.59
Å resolution and the proposed amino acid sequence of a family-11 xylanase from the
thermophilic fungus Paecilomyces varioti Bainier. J Mol Biol 295(3):581–593
Kumar V, Satyanarayana T (2011) Applicability of thermo-alkali-stable and cellulase-free
xylanase from a novel thermo-halo-alkaliphilic Bacillus haloduransin producing xylooligosac-
charides. Biotechnol Lett 33:2279–2285
La Grange DC, Claeyssens M, Pretorius IS, Van Zyl WH (2000) Coexpression of the Bacillus
pumilus beta-xylosidase (xynB) gene with the Trichoderma reesei beta xylanase 2 (xyn2) gene
in the yeast Saccharomyces cerevisiae. Appl Microbiol Biotechnol 54(2):195–200
Lagaert S, Van Campenhout S, Pollet A, Bourgois TM, Delcour JA, Courtin CM, Volckaert G
(2007) Recombinant expression and characterization of a reducing-end xylose-releasing
exo-oligoxylanase from Bifidobacterium adolescentis. Appl Environ Microbiol 73:5374–5377
Lama L, Calandrelli V, Gambacorta A, Nicolaus B (2004) Purification and characterization of
thermostable xylanase and beta-xylosidase by the thermophilic bacterium Bacillus ther-
mantarcticus. Res Microbiol 155(4):283–289
Lamed R, Setter E, Bayer EA (1983) Characterization of a cellulose-binding, cellulase-containing
complex in Clostridium thermocellum. J Bacteriol 156(2):828–836
Lappalainen A, Siika-aho M, Kalkkinen N, Fagerstrom R, Tenkanen M (2000) Endoxylanase II
from Trichoderma reesei has several isoforms with different isoelectric points. Biotechnol Appl
Biochem 31:61e68
Larson SB, Day J, Barba de la Rosa AP, Keen NT, McPherson A (2003) First crystallographic
structure of a xylanase from glycoside hydrolase family 5: implications for catalysis.
Biochemistry 42(28):8411–8422
Lee J (1997) Biological conversion of lignocellulosic biomass to ethanol. J Biotechnol 56(1):1–24
Lee JM, Hu Y, Zhu H, Cheng KJ, Krell PJ, Forsberg CW (1993) Cloning of a xylanase gene from
the ruminal fungus Neocallimastix patriciarum 27 and its expression in Escherichia coli. Can J
Microbiol 39(1):134–139
Lee TH, Lim PO, Lee YE (2007) Cloning, characterization, and expression of xylanase A gene
from Paenibacillus sp. DG-22 in Escherichia coli. J Microbiol Biotechnol 17(1):29–36
Levasseur A, Asther M, Record E (2005) Overproduction and characterization of xylanase B from
Aspergillus niger. Can J Microbiol 51(2):177–183
Li XL, Ljungdahl LG (1996) Expression of Aureobasidium pullulans xynA in, and secretion of the
xylanase from Saccharomyces cerevisiae. Appl Environ Microbiol 62(1):209–213
Li JH, Wang LS (2011) Why substituting the asparagine at position 35 in Bacillus circulans
xylanase with an aspartic acid remarkably improves the enzymatic catalytic activity? A
quantum chemistry-based calculation study. Polym Degrad Stab 96(5):1009–1014
Li C, Knierim B, Manisseri C, Arora R, Scheller HV, Auer M, Vogel KP, Simmons BA, Singh S
(2010) Comparison of dilute acid and ionic liquid pretreatment of switchgrass: biomass
recalcitrance, delignification and enzymatic saccharification. Bioresour Technol 101(13):4900–
4906
Li XL, Skory CD, Ximenes EA, Jordan DB, Dien BS, Hughes SR, Cotta MA (2007) Expression of
an AT-rich xylanase gene from the anaerobic fungus Orpinomyces sp. strain PC-2 in and
secretion of the heterologous enzyme by Hypocrea jecorina. Appl Microbiol Biotechnol
74:1264–1275
208 H.A. El Enshasy et al.
Liu JR, Duan CH, Zhao X, Tzen JT, Cheng KJ, Pai CK (2008) Cloning of a rumen fungal xylanase
gene and purification of the recombinant enzyme via artificial oil bodies. Appl Microbiol
Biotechnol 79(2):225–233
Liu MQ, Weng XY, Sun JY (2006) Expression of recombinant Aspergillus niger xylanase A in
Pichia pastoris and its action on xylan. Protein Expr Purif 48(2):292–299
Lo YC, Lu WC, Chen CY, Chen WM, Chang JS (2010) Characterization and high-level
production of xylanase from an indigenous cellulolytic bacterium Acinetobacter junii F6-02
from southern Taiwan soil. Biochem Eng J 53(1):77–84
Lundgren KR, Bergkvist L, Hogman S, Joves H, Eriksson G, Bartfai T, Vanderlaan J,
Rosenberg E, Shoham Y (1994) TCF mill trial on softwood pulp with korsnas thermostable
and alkaline stable xylanase T6. Fems Microbiology Reviews 13(2–3):365–368
Luthi E, Love DR, McAnulty J, Wallace C, Caughey PA, Saul D, Bergquist PL (1990) Cloning,
sequence analysis, and expression of genes encoding xylan-degrading enzymes from the
thermophile Caldocellum saccharolyticum. Appl Environ Microbiol 56(4):1017–1024
Lynd LR, Cushman JH, Nichols RJ, Wyman CE (1991) Fuel ethanol from cellulosic biomass.
Science 251(4999):1318–1323
Mach-Aigner A, Pucher ME, Mach RL (2010) D-Xylose as a repressor or inducer of xylanase
expression in Hypocrea jecorina (Trichoderma reesei). Appl Environ Microbiol 76:1770–1776
Mamo G, Hatti-Kaul R, Mattiasson B (2006) A thermostable alkaline active endo-b-1-4-xylanase
from Bacillus halodurans S7: Purification and characterization. Enzyme Microb Technol
39:1492–1498
Mamo G, Hatti-Kaul R, Mattiasson B (2007) Fusion of carbohydrate binding modules from
Thermotoga neapolitana with a family 10 xylanase from Bacillus halodurans S7.
Extremophiles 11:169–177
Mannarelli BM, Evans S, Lee D (1990) Cloning, sequencing, and expression of a xylanase gene
from the anaerobic ruminal bacterium Butyrivibrio fibrisolvens. J Bacteriol 172(8):4247–4254
Margeot A, Hahn-Hagerdal B, Edlund M, Slade R, Monot F (2009) New improvements for
lignocellulosic ethanol. Curr Opin Biotechnol 20:372–380
Marrone L, McAllister KA, Clarke AJ (2000) Characterization of function and activity of domains
A, B and C of xylanase C from Fibrobacter succinogenes S85. Protein Eng 13(8):593–601
McCarthy AA, Morris DD, Bergquist PL, Baker EN (2000) Structure of XynB, a highly
thermostable beta 1,4-xylanase from Dictyoglomus thermophilum Rt46B.1, at 1.8 Å resolution.
Acta Crystallogr D56:1367–1375
Mirande C, Mosoni P, Bera-Maillet C, Bernalier-Donadille A, Forano E (2010) Characterization of
Xyn10A, a highly active xylanase from the human gut bacterium Bacteroides xylanisolvens
XB1A. Appl Microbiol Biotechnol 87(6):2097–2105
Moreau A, Durand S, Morosoli R (1992) Secretion of a Cryptococcus albidus xylanase in
Saccharomyces cerevisiae. Gene 116(1):109–113
Morris DD, Gibbs MD, Bergquist PL (1996) Cloning of a family G xylanase gene (XYNB) from
the extremely thermophilic bacterium Dictyoglomus thermophilum and action of the gene
product on kraft pulp. Am Chem Soc 211:234
Muilu J, Torronen A, Perakyla M, Rouvinen J (1998) Functional conformational changes of
endo-1,4-xylanase II from Trichoderma reesei: a molecular dynamics study. Proteins
31:434–444
Mullai P, Fathima NSA, Rene ER (2010) Statistical analysis of main and interaction effects to
optimize xylanase production under submerged cultivation conditions. J Agric Sci 2
(1):144–153
Nakamura S, Wakabayashi K, Nakai R, Aono R, Horikoshi K (1993) Production of alkaline
xylanase by a newly isolated alkaliphilic Bacillus sp strain 41 m-1. World J Microbiol
Biotechnol 9(2):221–224
Narang S, Satyanarayana T (2001) Thermostable alpha-amylase production by an extreme
thermophile Bacillus thermooleovorans. Lett Appl Microbiol 32:31–35
7 Microbial Xylanases: Sources, Types, and Their Applications 209
Nascimento RP, Coelho RRR, Marques S, Alves L, Girio FM, Bon EPS, Amaral-Collaco MT
(2002) Production and partial characterisation of xylanase from Streptomyces sp strain AMT-3
isolated from Brazilian cerrado soil. Enzyme Microb Technol 31(4):549–555
Natesh R, Bhanumoorthy P, Vithayathil PJ, Sekar K, Ramakumar S, Viswamitra MA (1999)
Crystal structure at 1.8 angstrom resolution and proposed amino acid sequence of a
thermostable xylanase from Thermoascus aurantiacus. J Mol Biol 288(5):999–1012
Ogasawara W, Shida Y, Furukawa T, Shimada R, Nakagawa S, Kawamura M, Yagyu T,
Kosuge A, Xu J, Nogawa M, Okada H, Morikawa Y (2006) Cloning, functional expression and
promoter analysis of xylanase III gene from Trichoderma reesei. Appl Microbiol Biotechnol
72(5):995–1003
Ohta K, Moriyama S, Tanaka H, Shige T, Akimoto H (2001) Purification and characterization of
an acidophilic xylanase from Aureobasidium pullulans var. melanigenum and sequence
analysis of the encoding gene. J Biosci Bioeng 92(3):262–270
Okeke BC, Obi SKC (1994) Lignocellulose and sugar compositions of some agro-waste materials.
Bioresour Technol 47(3):283–284
Olfa E, Mondher M, Issam S, Ferid L, Nejib NM (2007) Induction, properties and application of
xylanase activity from Sclerotinia sclerotiorum S2 fungus. J Food Chem 31:96–107
Paës G, Berrin J-G, Beaugrand J (2012) GH11 xylanases: structure/function/properties relation-
ships and applications. Biotechnol Adv 30:564–592
Paës G, O’Donohue MJJ (2006) Engineering increased thermostability in the thermostable GH-11
xylanase from Thermobacillus xylanilyticus. J Biotechnol 125:338–350
Payan F, Leone P, Porciero S, Furniss C, Tahir T, Williamson G, Durand A, Manzanares P,
Gilbert HJ, Juge N, Roussel A (2004) The dual nature of the wheat xylanase protein inhibitor
XIP-I: structural basis for the inhibition of family 10 and family 11 xylanases. J Biol Chem 279
(34):36029–36037
Payan F, Flatman R, Porciero S, Williamson G, Juge N, Roussel A (2003) Structural analysis of
xylanase inhibitor protein I (XIP-I), a proteinaceous xylanase inhibitor from wheat (Triticum
aestivum var soisson). Biochem J 372(Pt 2):399–405
Petrescu I, Lamotte-Brasseur J, Chessa JP, Ntarima P, Claeyssens M, Devreese B, Marino G,
Gerday C (2000) Xylanase from the psychrophilic yeast Cryptococcus adeliae. Extremophiles
4(3):137–144
Pokhrel S, Joo JC, Yoo YJ (2013) Shifting the optimum pH of Bacillus circuulans xylanase
towards acidic side by introducing arginine. Biotechnol Bioprocess Eng 18:35–42
Polizeli MLTM, Rizzatti ACS, Monti R, Terenzi HF, Jorge JA, Amorim DS (2005) Xylanases
from fungi: properties and industrial applications. Appl Microbiol Biotechnol 67:577–591
Pollet A, Schoepe J, Dornez E, Strelkov SV, Delcour JA, Courtin CM (2010) Functional analysis
of glycosidehydrolase family 8 xylanases shows narrow but distinct substrate specificities and
biotechnological potential. Appl Microbiol Biotechnol 87:2125–2135
Prade RA (1996) Xylanases: from biology to biotechnology. Biotechnol Genet Eng Rev 13
(1):101–132
Puchkaev AV, Koo LS, Ortiz de Montellano PR (2003) Aromatic stacking as a determinant of the
thermal stability of CYP119 from Sulfolobus solfataricus. Arch Biochem Biophys 409:52–58
Puls J, Tenkanen M, Korte HE, Poutanen K (1991) Products of hydrolysis of beechwood
acetyl-4-O-methylglucuronoxylan by a xylanase and an acetyl xylan esterase. Enzyme Microb
Technol 13(6):483–486
Qureshy AF, Khan LA, Khanna S (2002) Cloning, regulation and purification of cellulase-free
xylanase from Bacillus circulans Teri-42. Ind J Microbiol 42:35–41
Raedschelders G, Debefve C, Goesaert H, Delcour JA, Volckaert G, Van Campenhout S (2004)
Molecular identification and chromosomal localization of genes encoding Triticum aestivum
xylanase inhibitor I-like proteins in cereals. Theor Appl Genet 109(1):112–121
Ratto M, Poutanen K, Viikari L (1992) Production of xylanolytic enzymes by an alkalitolernt
Bacillus circulans stain. Appl Microbiol Biotechnol 37:470–473
210 H.A. El Enshasy et al.
Rizzatti ACS, Jorge JA, Terenzi HF, Rechia CGV, Polizeli MLTM (2001) Purification and
properties of a thermostable extracellular b-xylosidase produced by a thermotolerant
Aspergillus phoenicis. J Ind Microbiol Biotechnol 26:156–160
Rose SH, van Zyl WH (2002) Constitutive expression of the Trichoderma reesei beta-1,4-xylanase
gene (xyn2) and the beta-1,4-endoglucanase gene (egl) in Aspergillus niger in molasses and
defined glucose media. Appl Microbiol Biotechnol 58(4):461–468
Rouau X (1993) Investigations into the effects of an enzyme preparation for baking on what flour
dough pentosans. J Cereal Sic 18:145–157
Rouau X, Daviet S, Tahir T, Cherel B, Saulnier L (2006) Effect of the proteinaceous wheat
xylanase inhibitor XIP-I on the performances of an Aspergillus niger xylanase in bread
making. J Sci Food Agric 86(11):1604–1609
Ruanglek V, Sriprang R, Ratanaphan N, Tirawongsaroj P, Chantasigh D, Tanapongpipat S,
Pootanakit K, Eurwilaichitr L (2007) Cloning, expression, characterization, and high
cell-density production of recombinant endo-1,4-beta-xylanase from Aspergillus niger in
Pichia pastoris. Enzyme Microb Technol 41(1–2):19–25
Saha BC (2003) Hemicellulose bioconversion. J Ind Microbiol Biotechnol 30(5):279–291
Sakka K, Maeda Y, Hakamada Y, Takahashi N, Shimada K (1991) Purification and some
properties of xylanase from Clostridium stercorarium strain Hx-1. Agric Biol Chem 55
(1):247–248
Salles BC, Cunha RB, Fontes W, Sousa MV, Filho EXF (2000) Purification and characterization
of a new xylanase from Acrophialophora nainiana. J Biotechnol 81(2–3):199–204
Salles BC, Te’o VS, Gibbs MD, Bergquist PL, Filho EX, Ximenes EA, Nevalainen KM (2007)
Identification of two novel xylanase-encoding genes (xyn5 and xyn6) from Acrophialophora
nainiana and heterologous expression of xyn6 in Trichoderma reesei. Biotechnol Lett 29
(8):1195–1201
Sanghi A, Garg N, Kuhar K, Kuhad RC, Gupta VK (2009) Enhanced production of cellulase-free
xylanase by alkalophilic Bacillus subtilis ASH and its application in biobleaching of kraft
pulp. BioResources 4:1109–1129
Sandhu JS, Kennedy JF (1984) Molecular cloning of Bacillus polymyxa (1–4)-P-D-xylanase gene
in Escherichia coli. Enzyme Microb Technol 6:271–274
Sapre MP, Jha H, Patil MB (2005) Purification and characterization of a thermoalkalophilic
xylanase from Bacillus sp. World J Microbiol Biotechnol 21(5):649–654
Schlacher A, Holzmann K, Hayn M, Steiner W, Schwab H (1996) Cloning and characterization of
the gene for the thermostable xylanase XynA from Thermomyces lanuginosus. J Biotechnol 49
(1–3):211–218
Schulze E (1891) Amphibia Europaea, annual report and treatises society of natural scientists in
Magdeburg 1890:163–178
Shallom D, Shoham Y (2003) Microbial hemicellulases. Curr Opin Microbiol 6(3):219–228
Shao W, Wiegel J (1995) Purification and characterization of two thermostable acetyl xylan
esterases from Thermoanaerobacterium sp. strain JW/SL-YS485. Appl Environ Microbiol 61
(2):729–733
Sharma A, Adahikari S, Styanarayana T (2007) Alkali-thermostable and cellulase-free xylanase
production by an extreme thermophile Geobacillus thermoleovorans. World J Microbiol
Biotechnol 23:483–490
Shatalov AA, Pereira AA (2007) Xylanase pre-treatment of giant reed organosolv pulps: direct
bleaching effect and bleach boosting. Ind Crops Prod 25:248–256
Shibuya H, Kaneko S, Hayanshi K (2000) Enhancement of thermostability and hydrolytic activity
of xylananse by random gene shuffling. Biochem J 349:651–656
Shrinivas D, Savitha G, Raviranjan K, Naik GR (2010) A highly thermostable alkaline
cellulase-free xylanase from thermoalkalophilic Bacillus sp. JB 99 suitable for paper and pulp
industry: purification and characterization. Appl Biochem Biotechnol 162(7):2049–2057
Sidhu G, Withers SG, Nguyen NT, McIntosh LP, Ziser L, Brayer GD (1999) Sugar ring distortion
in the glycosyl-enzyme intermediate of a family G/11 xylanase. Biochemistry 38
(17):5346–5354
7 Microbial Xylanases: Sources, Types, and Their Applications 211
Silva CHC, Puls J, de Sousa MV, Ferreira EX (1999) Purification and characterization of a low
molecular weight xylanase from solid-state cultures of Aspergillus fumigatus Fresenius. Rev
Microbiol 30(2):114–119
Simpson HD, Haufler UR, Daniel RM (1991) An extremely thermostable xylanase from the
thermophilic bacterium Thermotoga. Biochem J 277(2):413–417
Simpson PJ, Bolam DN, Cooper A, Ciruela A, Hazlewood GP, Gilbert HJ, Williamson MP (1999)
A family IIb xylan-binding domain has a similar secondary structure to a homologous family
IIa cellulose-binding domain but different ligand specificity. Structure 7(1999):853–864
Singh S, Pillay B, Dilsook V, Prior BA (2000) Production and properties of hemicellulases by a
Thermomyces lanuginosus strain. J Appl Microbiol 88:975–982
Sorensen JF, Sibbesen O (2006) Mapping of residues involved in the interaction between the
Bacillus subtilis xylanase A and proteinaceous wheat xylanase inhibitors. Protein Eng Des Sel
19(5):205–210
Sriprang R, Asano K, Gobsuk J, Tanapongpipat S, Champreda V, Eurwilaichitr L (2006)
Improvement of thermostability of fungal xylanase by using site-directed mutagenesis.
J Biotechnol 126(4):454–462
Srivastava R, Srivastava AK (1993) Characterization of a bacterial xylanase resistant to repression
by glucose and xylose. Biotechnol Lett 15(8):847–852
Sriyapai T, Somyoonsap P, Matsui K, Kawai F, Chansiri K (2011) Cloning of a thermostable
xylanase from Actinomadura sp. S14 and its expression in Escherichia coli and Pichia
pastoris. J Biosci Bioeng 111(5):528–536
St John FJ, Godwin DK, Preston JF, Pozharski E, Hurlbert JC (2009) Crystallization and
crystallographic analysis of Bacillus subtilis xylanase C. Acta Crystallogr Sect F Struct Biol
Crystallogr Commun 65(Pt 5):499–503
Strauss J, Mach RL, Zeilinger S, Hartler G, Stoffler G, Wolschek M, Kubicek CP (1995) Cre1, the
carbon catabolite repressor protein from Trichoderma reesei. FEBS Lett 376:103–107
Stephens DE, Singh S, Permaul K (2009) Error-prone PCR of a fungal xylanase for improvement
of its alkaline and thermal stability. FEMS Microbiol Lett 293(1):42–47
Subramaniyan S (2000) Studies on the production of bacterial xylanases. Ph.D. Thesis, Cochin
University of Science and Technology, India
Subramaniyan S, Prema P (2002) Biotechnology of microbial xylanases: enzymology, molecular
biology and application. Crit Rev Biotechnol 22:33–46
Sun JY, Liu MQ, Weng XY, Qian LC, Gu SH (2007) Expression of recombinant
Thermomonospora fusca xylanase A in Pichia pastoris and xylooligosaccharides released
from xylans by it. Food Chem 104(3):1055–1064
Sunna A, Prowe SG, Stoffregen T, Antranikian G (1997) Characterization of the xylanases from
the new isolated thermophilic xylan-degrading Bacillus thermoleovorans strain K-3d and
Bacillus flavothermus strain LB3A. FEMS Microbiol Lett 148(2):209–216
Suzuki T, Ibata K, Hatsu M, Takamizawa K, Kawai K (1997) Cloning and expression of a 58-kDa
xylanase VI gene (xynD) of Aeromonas caviae ME-1 in Escherichia coli which is not
categorized as a family F or family G xylanase. J Ferm Bioeng 84(1):86–89
Takahashi-Ando N, Inaba M, Ohsato S, Igawa T, Usami R, Kimura M (2007) Identification of
multiple highly similar XIP-type xylanase inhibitor genes in hexaploid wheat. Biochem
Biophys Res Commun 360(4):880–884
Tanaka H, Okuno T, Moriyama S, Muguruma M, Ohta K (2004) Acidophilic xylanase from
Aureobasidium pullulans: efficient expression and secretion in Pichia pastoris and mutational
analysis. J Biosci Bioeng 98(5):338–343
Taneja K, Gupta S, Kuhad RC (2002) Properties and application of a partially purified alkaline
xylanase from an alkalophilic fungus Aspergillus nidulans KK-99. Bioresour Technol 85
(1):39–42
Teng C, Jia H, Yan Q, Zhou P, Jiang Z (2011) High-level expression of extracellular secretion of a
b-Xylosidase gene from Paecilomyces thermophila in Escherichia coli. Bioresour Technol
102:1822–1830
212 H.A. El Enshasy et al.
Tomme P, Gilkes NR, Miller RC, Warren AJ, Kilburn DG (1994) An internal cellulose-binding
domain mediates adsorption of an engineered bifunctional xylanase/cellulase. Protein Eng 7
(1):117–123
Torronen A, Harkki A, Rouvinen J (1994) 3-Dimensional structure of endo-1,4-beta-xylanase-II
from Trichoderma reesei-2 conformational states in the active-site. EMBO J 13
(11):2493–2501
Torronen A and Rouvinen J (1997) Structural and functional properties of low molecular weight
endo-1,4-beta-xylanases. J Biotechnol 57:137–149
Trudel J, Grenier J, Potvin C, Asselin A (1998) Several thaumatin-like proteins bind to
beta-1,3-glucans. Plant Physiology 118:1431–1438
Tsai CT, Huang CT (2008) Overexpression of the Neocallimastix frontalis xylanase gene in the
methylotrophic yeasts Pichia pastoris and Pichia methanolica. Enzyme Microb Technol 42
(6):459–465
Turunen O, Etuaho K, Fenel F, Vehmaanpera J, Wu XY, Rouvinen J, Leisola M (2001) A
combination of weakly stabilizing mutations with a disulfide bridge in the alpha-helix region of
Trichoderma reesei endo-1,4-beta-xylanase II increases the thermal stability through
synergism. J Biotechnol 88(1):37–46
Valenzuela SV, Diaz P, Javier Pastor FI (2010) Recombinant expression of an alkali stable GH10
xylanase from Paenibacillus barcinonensis. J Agric Food Chem 58(8):4814–4818
Van Campenhout S, Pollet A, Bourgois TM, Rombouts S, Beaugrand J, Gebruers K, De Backer E,
Courtin CM, Delcour JA, Volckaert G (2007) Unprocessed barley aleurone
endo-beta-1,4-xylanase X-I is an active enzyme. Biochem Biophys Res Commun 356
(3):799–804
Van Der Borght A, Goesaert H, Veraverbeke WS, Delcour JA (2005) Fractionation of wheat and
wheat flour into starch and gluten: overview of the main processes and the factors involved.
J Cereal Sci 41:221–237
Van Peij NN, Brinkmann J, Vrsanska M, Visser J, de Graaff LH (1997) Beta-xylosidase activity,
encoded by xlnD, is essential for complete hydrolysis of xylan by Aspergillus niger but not for
induction of the xylanolytic enzyme spectrum. Eur J Biochem 245(1):164–173
Van Petegem F, Collins T, Meuwis MA, Gerday C, Feller G, Van Beeumen J (2003) The structure
of a cold-adapted family 8 xylanase at 1.3 Å resolution. Structural adaptations to cold and
investigation of the active site. J Biol Chem 278(9):7531–7539
Vandeplas S, Dauphin RD, Thonart P, Thewis A, Beckers Y (2010) Effect of the bacterial or
fungal origin of exogenous xylanases supplemented to a wheat-based diet on performance of
broiler chickens and nutrient digestibility of the diet. Can J Anim Sci 90:221–228
Vardakou M, Flint J, Christakopoulos P, Lewis RJ, Gilbert HJ, Murray JW (2005) A family 10
Thermoascus aurantiacus xylanase utilizes arabinose decorations of xylan as significant
substrate specificity determinants. J Mol Biol 352:1060–1067
Verjans P, Dornez E, Delcour JA, Courtin CM (2010) Selectivity for water-unextractable
arabinoxylan and inhibition sensitivity govern the strong bread improving potential of an
acidophilic GH11 Aureobasidium pullulans xylanase. Food Chem 123:331–337
Verma D, Satyanarayana T (2012) Cloning, expression and applicability of thermo-alkali-stable
xylanase of Geobacillus thermoleovorans in generating xylooligosaccharides from
agro-residues. Bioresour Technol 107:333–338
Wang JS, Bai YG, Yang PL, Shi PJ, Luo HY, Meng K, Huang HQ, Yin J, Yao B (2010) A new
xylanase from thermoalkaline Anoxybacillus sp E2 with high activity and stability over a broad
pH range. World J Microbiol Biotechnol 26(5):917–924
Wang Q, Xia T (2008) Enhancement of the activity and alkaline pH stability of Thermobifida fusca
xylanase A by directed evolution. Biotechnol Lett 30(5):937–944
Whitehead TR, Hespell RB (1989) Cloning and expression in Escherichia coli of a xylanase gene
from Bacteroides ruminicola 23. Appl Environ Microbiol 55(4):893–896
Winterhalter C, Liebl W (1995) Two extremely thermostable xylanases of the hyperthermophilic
bacterium Thermotoga maritima MSB8. Appl Environ Microbiol 61:1810–1815
7 Microbial Xylanases: Sources, Types, and Their Applications 213
Suttipun Keawsompong
S. Keawsompong (&)
Department of Biotechnology, Faculty of Agro-Industry, Kasetsart University,
Bangkok, Thailand
e-mail: [email protected]
8.1 Mannans
Mannans and heteromannans are widely distributed in nature as part of the hemi-
cellulose fraction in hardwoods and softwoods, seeds of leguminous plants
(Handford et al. 2003; Buckeridge et al. 2000), and in beans (Lundqvist et al. 2003).
Hemicelluloses are copolymers of both hexose and pentose sugars. The branched
structure allows hemicellulose to exist in an amorphous form that is more sus-
ceptible to hydrolysis. Within biomass, mannans or the hemicelluloses are situated
between the lignin and the collection of cellulose fibers underneath. Consistent with
their structure and side group substitutions, mannans seem to be interspersed and
covalently linked with lignins at various points while producing a coat around
underlying cellulose strands via hydrogen bonds, but as few H-bonds are involved
they are much more easily broken down than cellulose. The mannan layer with its
covalent linkage to lignin and its non-covalent interaction with cellulose maybe
important in maintaining the integrity of the cellulose in situ and in helping to
protect the fibers against degradation to cellulases (Puls and Schuseil 1993).
Mannan is the predominant hemicellulosic polysaccharide in softwoods from
gymnosperms, but is the minor hemicellulose in hardwood from angiosperms (Puls
and Schuseil 1993). Unsubstituted β-1,4-mannan, composed of a main chain of
β-mannose residues (Fig. 8.1), is an important structural component of some marine
algae and terrestrial plants such as ivory nut (Chanzy et al. 2004) and coffee bean
(Nunes et al. 2006). It resembles cellulose in the conformation of the individual
polysaccharide chains, and is water insoluble.
The galactomannans are mainly found in the seeds of the family of
Leguminoseae and are located in the endospermic part of the seeds (Dea and
Morrison 1975). Galactomannans consist of water-soluble 1,4-linked β-D-manno-
pyranosyl residues with side chains of single 1,6-linked α-D-galactopyranosyl
groups (Shobha et al. 2005).
Glucomannan is a water-soluble polysaccharide that is considered a dietary fiber.
It is a hemicellulose component in the cell walls of some plant species (Gonzáles
et al. 2004). Glucomannan is a food additive used as an emulsifier and thickener.
The polysaccharide consists of glucose and mannose in a proportion of 1:3 joined
by β-(1,4) linkages (Popa and Spiridon 1998). Hardwoods consist of glucomannan
with a glucose/mannose ratio of 1:1.5–2 (Hongshu et al. 2002).
Galactoglucomannan is a water-soluble hemicellulose, consisting of galactose,
glucose, and mannose. Many softwood species, e.g., Norway spruce are rich of
galactoglucomannans and can contain it up to 10–20 %. Galactoglucomannan
consists of a backbone of randomly distributed (1 → 4)-linked mannose and glu-
cose units with (1 → 6)-linked galactose units attached to mannose units. The
hydroxyl groups in locations C2 and C3 in mannose are partially substituted by
acetyl groups (Willför et al. 2007).
8 Mannanase 217
enzyme, cleaves β-1,4-linked mannosides and releases mannose from the nonre-
ducing ends of mannan and mannooligosacharides. β-Glucosidases, exo acting
enzyme, hydrolyzes β-1,4-glucopyranose at the nonreducing ends of oligosaccharides
that released from glucomannan and galactoglucomannan degradation. In addition,
the side chain sugars that attached at various points on mannan are removed by
α-galactosidase and acetyl mannan esterase. α-Galactosidase hydrolyzes α-1,6-linked
D-galactopyranose at side chain of galactomannan and galactoglucomannan. Acetyl
mannan esterase releases acetyl groups from galactoglucomannan.
However, mannan degradation by these enzymes is affected by the degree and
pattern of substitution of glucose and glucose residues in backbone of these
polysaccharides (Van Zyl et al. 2010). In addition, pattern of distribution of
O-acetyl group in glucomannan also affects the hydrolysis by the enzymes.
8.2.1 β-Mannanase
Fig. 8.3 The retaining mechanism of β-mannanase. Source Chauhan et al. (2012)
220 S. Keawsompong
condition for bacterial growth, and production and acidic pH for fungi (Chauhan
et al. 2012). However, bacterial β-mannanases are more thermostable than fungal
β-mannanases that are important for industrial application, such as β-mannanases
from Bacillus species have the advantages of high activity and convenient isolation
and thus have been used in research and industry (Araujo and Ward 1990).
β-Mannanase has been used in several industrial applications because of its broad
substrate specificity.
Alkaline β-mannanase with stable in detergents has been used in laundry segments
as stain removal boosters. Mannans are generally used as thickening agents in
several products such as hair gel, shampoo, conditioner, and toothpaste. The stains
containing mannan are difficult to remove so this enzyme can cleave into smaller
carbohydrate fragments that can reduce the stain cleaning process and remove
during the washing. Moreover, β-mannanase can be formulated as sanitization
products, contact lens cleanser and hard surface cleansers (Chauhan et al. 2012).
β-Mannanase has been used in the viscosity reduction of coffee bean extracts by
hydrolyzing the mannan component in the coffee extract (Chauhan et al. 2012). It
has been used in the maceration of fruit and vegetable materials and clarification of
222 S. Keawsompong
fruit juices (Moreira and Filho 2008). In addition, β-mannanase can be used in
enzymatic oil extraction of coconut. The enzymatic process can eliminate the
problems of alfatoxin contamination and oxidative rancidity of products (Chauhan
et al. 2012).
β-Mannanase can be used to improve the nutritional value of animal feed especially
poultry. Mannan polysaccharides are commonly found in feed ingredients such as
soyabean meal, guar meal, copra meal, palm kernel meal, and sesame meal. All
these meals have some common properties such as high fiber content, low
palatability, lack of several essential amino acids, and high viscosity coupled with
several anti-nutritional factors such as mannan, galactomannan, xylan, and arabi-
noxylan that limit the utilization in the animal intestine (Chauhan et al. 2012).
Moreover, they have been found to be highly deleterious to animal performance,
severely compromising weight gain and feed conversion as well as glucose and
water absorption (Dhawan and Kaur 2007). Therefore, incorporation of
β-mannanase in feed can help to cleave mannan and release of nutrients results in
increased villus height in duodenum and jejunum that leads to increase in surface
area and adsorption and decreased intestinal viscosity. So, it can improve both the
weight gain and feed conversion efficiency (Adibmoradi and Mehri 2007). Hemicell
supplied by ChemGen, USA is a fermentation product of B. lentus containing high
amount of β mannanases that degrade mannan in feed (Daskiran et al. 2004).
longer villi and a more shallow crypt (Yang et al. 2008; Baurhoo et al. 2009). These
studies showed the results that MOS could increase the energy of digestion and
production of digestive enzymes such as alkaline phosphatase, maltase, and leucine
aminopeptidase. Moreover, MOS could increase the goblet cells, mucus-producing
cells so the villi and intestinal surface could be more protected.
Mannose has been used as a component of medicine because of its properties such
as fast dissolving and structure forming (Chauhan et al. 2012). Mannose also has
been used as a remedy for urinary tract infection (Van Zyl et al. 2010). There is a
significant increase in using this sugar, so β-mannanase and other enzymes can be
used for the economical production of mannose from low-cost mannan substrates
such as palm kernel cake and copra meal. Guar gum has the positive effects on
some physiological functions like reducing plasma cholesterol and body fat without
reducing protein utilization and increased fecal excretion volume (Takeno et al.
1990). Therefore, a partially hydrolyzed guar gum (PHGG) with β-mannanase is
used in beverage form for treatment of several diseases such as irritable bowel
syndrome (IBS). PHGG can increase stool weight and decrease colon transit time
by providing non-digestible bulk, retaining water, and serving as a substrate for
microbial growth in the colon (Parisi et al. 2002). In addition, PHGG supplemented
with oral rehydration solution is also used for the treatment of acute diarrhea in
children by providing short chain fatty acids in large intestine and maintaining the
balance of salt and water (Alam et al. 2000).
The molecules in wood and soft wood are linked together in complexes known as
cellulose and hemicelluloses and together can represent up to 75 % of dried wood
weight. Moreover, sugar and nutrient is found in wood tissue and soft wood.
Getting access to these complexes is not easy. Surrounding the cellulose and
hemicelluloses is lignin which serves as the glue and protects the polysaccharides
from enzyme and microbial deconstruction. The lignin, cellulose, and hemicellulose
form strong bonds and their combined structure is named lignocellulose. The
process to overcome lignocellulosic recalcitrance and expose the cellulose and
hemicellulose, so that individual sugars can be released, is called pretreatment
shown in Fig. 8.4.
Pretreatment represents the necessary steps to convert the raw substrate into a
suitable form and includes size reduction by grinding, physical, chemical, or
enzymatic hydrolysis to increase substrate availability to release nutrient and sugar.
224 S. Keawsompong
Acid Pretreatment
Concentrated acid hydrolysis disrupts the hydrogen bonding of the cellulose chains,
converting the cellulose from a crystalline state to an amorphous state with little
glucose liberation (Kosaric and Vardar-Sukan 2001; Orozco et al. 2007). In 1965,
Oshima showed that crystalline cellulose achieves complete solubility at 72 %
H2SO4 or 42 % HCl at the relatively low temperatures (10–45 °C) (Oshima 1965;
Kosaric and Vardar-Sukan 2001). In the process of concentrated acidic hydrolysis,
dilute acid is used initially in the prehydrolysis digestion column for hemicellulose
removal prior to passing to the secondary counter current column to contact a strong
acidic solution With the concentrated acid method, glucose yields greater than 90 %
have been achieved with a rapid production rates whereas the increased acid
concentration of HCl, a reduction of xylose yield has been showed by the work of
Orozco et al. (2007).
Alkaline Hydrolysis
Alkali is seen as a swelling agent where the alkali acts indirectly with water being
the lysis agent (Kosaric and Vardar-Sukan 2001). By swelling the biomass, the
surface area is increased opening up the structure for water to migrate into the
material. Once inside the biomass, the water disrupts the hydrogen bonding
between the hemicellulose and the lignin carbohydrates (Balat et al. 2008). The
effect is the decrease in crystal and lignin disruption. The advantages of using alkali
over acidic methods are the removal of the lignin fraction without the degradation
8 Mannanase 225
of the other major constituents. The compromise is the increased length of reaction
times [hours or days in comparison to minutes for other processing methods (Balat
et al. 2008)].
Balat et al. (2008) pretreated corn stover with NaOH (2 %) and combined with
irradiation (500 gGy). They found that the glucose yields to increase from 20 % for
NaOH pretreatment to 43 % in combination (Balat et al. 2008). Although NaOH is
the dominant alkali approach, lime [calcium hydroxide, Ca(OH)2] can also be
employed as an alternative method. Ca(OH)2 has been used in the pretreatment of
several feedstocks in pilot scale applications, such as wheat straw (358 K for 3 h
and corn stover (373 K for 13 h). The attraction of lime/alkali pretreatment is the
removal of acetyl and various uronic acid substitutions on hemicellulose that lowers
the accessibility of the enzymes to the hemicellulose and cellulose surface (Sun and
Cheng 2002; Balat et al. 2008).
Steam Pretreatment
During hydrolysis, the polymers of hemicellulose are degraded into their mono-
meric subunits. Complete hemicellulose hydrolysis results in several hexoses and
pentoses. Main sugar in hemicellulose derived from soft wood is mannose
(Taherzadeh and Karimi 2007). The enzyme hydrolysis is catalyzed by hemicel-
lulose enzymes; without any pretreatment the conversion of native hemicellulose to
sugar is extremely slow, since hemicellulose is well protected by the lignin.
Therefore, pretreatment of these materials is necessary to increase the rate of
hydrolysis of hemicellulose (Galbe et al. 2007).
Efficiently releasing all sugars in biomass therefore requires the combined action
of a large number of different enzymes. Furthermore, some of these enzymes could
226 S. Keawsompong
β-Mannanase can use as slime control agent in water purification system, waste
water treatment, and cooling water treatment system. It has been used in the
enzymatic hydrolysis of galactomannan to enhance the flow of oil and gas in
drilling operation in the oil and gas industries (Chauhan et al. 2012).
8.4 Conclusions
References
9.1 Introduction
25
15 Canada
China
10 Europe
Brazil
5 USA
0
2007 2008 2009 2010 2011 2012 2013
Fig. 9.1 Country-wise global ethanol production (2007–2013). Adapted from www.afdc.energy.
gov/data/
Table 9.1 Average composition of biomass expressed as percentage on dry weight basis
Biomass Cellulose (%) Hemicellulose (%) Lignin (%)
Barley wood 40 20 15
Bermuda grass 25 35.7 6.4
Birch wood 40 33 21
Corn cobs 42 39 14
Corn stalks 35 15 19
Corn stover 38 26 19
Cotton seed hair 80–95 5–20 0–5
Flax sheaves 35 24 22
Forage sorghum 34 17 16
Grasses 25–40 35–50 10–30
Groundnut shells 38 36 16
Hardwood stem 40–55 34–40 18–25
Leaves 15–20 80–85 0–5
Miscanthus 43 24 19
Oat straw 41 16 11
Paper 85–99 0–5 0–15
Pine 41 10 27
Rice straw 32 24 13
Rice husk 36 15 19
Rye straw 31 25 7
Saw dust 55 14 21
Sorghum straw 33 18 15
Soybean stalks 34 25 20
Sugarcane 33 30 29
Sugarcane bagasse 42 25 20
Sweet sorghum 23 14 11
Switch grass 37 29 19
Salix 41.5 22–25 25
Softwood stem 45–50 25–35 25–35
Spruce 45 26 28
Wheat straw 30 24 18
Adapted from Jørgensen et al. (2007) and Kumar et al. (2009)
the other hand, Type II cell wall, present in commelinoid monocots like rice,
contains less of xyloglucans and shows predominance of glucuronoarabinoxylan
along with cellulose microfibrils (Yokoyama et al. 2004). Whatever the plant type,
the considerable amount of xyloglucan present is always around 20–25 % of dry
weight in dicots and 2–5 % in grasses (Hayashi 1989; Benko et al. 2008).
9 The Role and Applications of Xyloglucan Hydrolase … 235
Fig. 9.2 The typical structures of the subunits in XXXG-type xyloglucans. Adapted from http://
www.ccrc.uga.edu/
Table 9.3 List of hemicellulases that can be supplemented to cellulase enzyme to enhance
saccharification efficiency of lignocellulosic biomass
Enzyme GH family Enzyme Site of action
classification
number (EC)
β-Xylanase GH 10, GH EC 3.2.1.8 Hydrolyses β-1,4 glycosidic bond
11 in xylan
β-Xylosidase GH 52 EC 3.2.1.37 Hydrolyses xylobioses
Glucanase GH 16 EC 3.2.1.- Hydrolyses β-1,3; β-1,6; α-1,4; α-1,6
glucans
Mannanase GH 26 EC 3.2.1.78 Hydrolyses β-1,4-mannosidic linkage in
the main chain of mannan,
galactomannan, glucomannan and
galactoglucomannan
Xyloglucan GH 12, GH EC Hydrolyses xyloglucan (β-1,4 glucan
hydrolase 16 3.2.1.150,151,155 with α-1,6 linked xylose)
Arabinofuranosidase GH 43, GH EC 3.2.1.55 Removes arabinose substituents from
51, GH 54, α-arabinofuranoside, arabinoxylans,
GH 127 arabinogalactans
Glucouronidase GH 67, GH EC 3.2.1.139 Removes α-1,2 linked glucuronoyl or its
79 methyl ester in xylan
Acetyl xylan CE 1, CE 5, EC 3.1.1.72 Removes acetyl group from xylans or
esterase CE 16 other xylooligosaccharides
Feruloyl esterase CE 10 EC 3.1.1.73 Hydrolyses feruloyl esters
Glucuronoyl CE 15 EC 3.1.1.- Demethylates methyl glucuronoyl α-1,2
esterase linked to backbone xylose
Adapted from Sweeney and Xu (2012)
238 M. Saritha et al.
Depending upon the type and abundance of hemicellulose polymer present in the
biomass used as substrate, different combinations of hydrolytic enzymes listed in
Table 9.3 can be supplemented to cellulases. Presence of hemicellulases in
hydrolytic enzyme cocktail has been reported to enhance the saccharification yield
by releasing the sugars entrapped in hemicelluloses and also by increasing the
accessibility of the cellulase enzyme to the cellulose complex (Singh et al. 2014).
The hydrogen bonding present between cellulose and xyloglucan makes the
extraction of latter difficult. In order to access and modify the cellulose fibers
embedded and cross-linked by xyloglucans in the plant cell walls, remodeling of
xyloglucan is required. In nature, the enzymatic hydrolysis of xyloglucan and the
resulting xyloglucan solubilization is important in controlling wall strength during
cell growth (Takahashi et al. 2015). The enzymatic machinery comprising of
specific xyloglucanases that can digest xyloglucans to xyloglucan-oligosaccharides
is, thus, of utmost importance in improving accessibility of cellulose to hydrolytic
enzymes and in enhancing the conversion rate of polymers to sugars in order to
economize the saccharification process. In contrast to the cell wall degrading
enzymes, they cause cell wall modifications that are usually associated with the
weakening of extracellular barriers.
Xyloglucan depolymerization was earlier thought to be due to the action of
endoglucanases and other undefined hydrolases. Later, xyloglucan endotransgly-
colases (XETs) and xyloglucan-specific endo-β-1,4-glucanase were identified from
germinating Nasturtium seeds, where it catalyzed the seed storage xyloglucan
depolymerization (Rose et al. 2003). Subsequently, several proteins belonging to
the same class and with sufficient sequence homology were discovered. Presently,
xyloglucan hydrolases which carry out hydrolysis of xyloglucan are classified in the
xyloglucan endotransglucosylase (XET)/hydrolase (XEH) superfamily, known as
the XTH superfamily (Rose et al. 2003). The modes of action of these enzymes are
illustrated in the Fig. 9.3.
Several xyloglucan-active degrading enzymes of the XTH superfamily have been
reported by various researchers. These include xyloglucan-specific
endo-β-1,4-glucanase (EC 3.2.1.151), xyloglucan-specific exo-β-1,4-glucanase (EC
3.2.1.155), oligoxyloglucan reducing-end-specific cellobiohydrolase (EC 3.2.1.150),
xyloglucan endotransferase (EC 2.4.1.207), xyloglucan 4-glucosyltransferase (EC
2.4.1.168), xyloglucan 6-xylosyltransferase (EC 2.4.2.39), and xyloglucan
2-galactose transferase (EC 2.4.1.-), belonging to several glycosyl hydrolase
(GH) families: GH5, GH12, GH16, GH 26, GH44, and GH74 (Hayashi et al. 1980;
Fry et al. 1992; Pauly et al. 1999; Faik et al. 2002; Madson et al. 2003; Grishutin et al.
9 The Role and Applications of Xyloglucan Hydrolase … 239
2004; Yaoi et al. 2004; Baumann 2007b). The particulars of a few enzymes involved
in xyloglucan hydrolysis are given in Table 9.4.
As stated previously, xyloglucan hydrolases are believed to be cell
wall-loosening enzymes that are responsible for cell elongation by continuous
splicing of existing polymers and linking of new xyloglucan residues (Fry et al.
1992; Miedes et al. 2011). High expression of the enzyme in growing parts like
elongating tissues in Cicer arientinum (Romo et al., 2005), hypocotyls in pine
(Lorences 2004), germinating Nasturtium seeds (Stratilova et al. 2010), fiber
elongation in cotton (Michailidis et al. 2009), and even ripening of tomatoes
240 M. Saritha et al.
(Miedes and Lorences 2009) have been reported. Enhanced expression of the
enzyme in plants has also been reported on contact with pathogenic or beneficial
microorganisms. Mycorrhizal association in Medicago truncatula caused enhanced
expression of enzyme in roots both intimately and distantly associated with the
fungal partner (Maldonado-Mendoza et al. 2005) and Botryosphaeria dothidea
caused enhanced expression of the enzyme in apple tree (Bai et al. 2015).
However, due to their role in maintaining the cell wall structure, xyloglucan-
active enzymes may also be considered potential hemicellulose-repairing enzymes.
The XTHs produced during infection of apple fruit by Penicillium expansum has a
dual role of integrating newly secreted xyloglucan chains into an existing
wall-bound xyloglucan and restructuring existing cell wall material by catalyzing
transglucosylation between previously wall-bound xyloglucan molecules
(Muñoz-Bertomeu and Lorences 2014). Their activity gets inhibited as the infection
progresses, leading to hemicellulose degradation.
All glycosyl hydrolase enzymes act by causing splits in glycosidic bonds and
presumably, utilize either the inverting or the retaining mechanism. The inverting
mechanism is a direct displacement mechanism, resulting in the inverted configu-
ration on the anomeric carbon. Enzymes employing this reaction mechanism belong
to either GH 44 or GH 74 family and generally exhibit a distance of approximately
11 Å between the catalytic residues (Baumann 2007b). The cellulosomes of
Clostridium thermocellum are found to produce extracellular cellulases, along with
an inverting xyloglucanase CtXGH74A (Martinez-Fleites et al. 2006). Retaining
xyloglucan hydrolases utilize the double displacement mechanism for catalysis. The
cell wall active xyloglucan hydrolases belonging to the GH 16 family employ this
mechanism that involves the formation of a covalent glycosyl-enzyme intermediate
(Baumann 2007b). The only xyloglucan hydrolyzing enzyme from GH16 charac-
terized in detail is NXG1, originally isolated from Nasturtium cotyledons (Edwards
et al. 1985). Recently, close homologs of NXG1 analyzed by expression profiling
have been shown to be expressed during different stages of plant development.
These include the OsXTH19 from rice leaves (Yokoyama et al. 2004) and the
SlXTH6 from tomato shoot, seeds, and roots (Saladie et al. 2006). The GH16
transglycosylases which catalyze endotransglycosylation of xyloglucan that enables
cell expansion also employ the retaining mechanism (Sinnott 1990). The modes of
action of xyloglucanases have also been variously classified into the endo-mode
involving cleavage at multiple sites along the backbone, and the exo-mode
involving reducing end-specific cleavage (Feng et al. 2014).
9 The Role and Applications of Xyloglucan Hydrolase … 241
These cleave the ‘X–X’ and ‘X–G’ motifs and have been characterized as having
endo-processive modes of action, i.e., the enzymes act by sliding along the back-
bone during the catalytic action (Matsuzawa et al. 2014). These include the
endo-xyloglucanase (XcXGHA) from Xanthomonas citri pv. mangiferaeindicae
(Feng et al. 2014), and SaGH74A and SaGH74B from Streptomyces avermitilis
(Ichinose et al. 2012). In contrast, the OXG-RCBH from Geotrichum sp. M128
recognizes the reducing end of various xyloglucan-derived oligosaccharides and
releases two glycosyl residues from this site in the exo-mode (Ichinose et al. 2012).
A growing interest in xyloglucanases is mainly due to the possibility of their
application in a number of biotechnological processes, such as conversion of plant
waste, modification of xyloglucans for use in food and feed industries as well as
pulp and paper industry, production of novel surfactants from oligoxyloglucans,
and thermally reversible xyloglucan gels for drug delivery (Sinitsyna et al. 2010;
Damásio et al. 2012). During the conversion of biomass to value-added products, a
suitable cocktail of various hydrolytic enzymes is required to release all the sugars
present in the biomass. Enzymatic saccharification has renewed and centralized the
focus on different aspects of hemicellulases as they play an important role in
improving the economics of the overall process. The barriers, which include
development of robust enzyme formulations containing all accessory enzymes in
commercial cellulases, have to be alleviated for successful commercialization of
biofuels. An economically feasible enzyme technology for complete hydrolysis of
biomass polysaccharides into sugar monomers is very important for cost-effective
biofuel production. To achieve this, there is need to search for organisms with
hypercellulase activity for developing better quality cellulase formulations with
superior attributes such as higher efficiencies, thermostability, and lowered feed-
back inhibition and tolerance to inhibition by toxic byproducts from pretreatments,
using sophisticated biotechnological tools (Bon and Ferrara 2007).
Another feasible option is to supplement cellulase preparations with hemicel-
lulases and other accessory enzymes to enhance the sugar recovery from pretreated
lignocellulosic biomass. It has been reported that Trichoderma reesei, a widely
known cellulolytic filamentous fungus and the main industrial source of commer-
cially available cellulases has only one GH74 gene that codes for xyloglucanase,
while several other fungi have more than two genes for the same (Ichinose et al.
2012). In this context, microbial xyloglucan hydrolase preparations may be
important for designing enzyme cocktails for industrial biomass conversion,
because the degradation of xyloglucan could enhance cellulase accessibility to
cellulose (Kaida et al. 2009). There are several reports on improvement in sugar
yields during biomass bioconversion when cellulases were supplemented with
xyloglucan hydrolase. Supplementation of α xylosidase with commercial cellulases
enhanced the release of free glucose (82–88 %) and xylose (55–60 %) from
hydrogen peroxide pretreated corn stover (Jabbour et al. 2013). Xyloglucanase
addition has also been shown to enhance the hydrolysability of various lignocel-
lulosic substrates when added to a cellulase mixture (Nishitani 1995; Benko et al.
2008). Hydrolysis of glycosidic linkage of xyloglucans resulted in swelling of
cellulose microfibrils which enhanced cellulose accessibility and, subsequently,
244 M. Saritha et al.
9.5 Conclusion
References
Annual Energy Outlook (AEO) (2015) U.S. Energy Information Administration, U.S. Department
of Energy, Washington, DC. http://www.eia.gov/forecasts/aeo
Bai S, Dong C, Zhu J, Zhang Y, Dai H (2015) Identification of a xyloglucan-specific endo-(1-4)-
beta-D-glucanase inhibitor protein from apple (Malus × domestica Borkh.) as a potential
defense gene against Botryosphaeria dothidea. Plant Sci 231:11–19
Baumann MJ (2007a) Structural evidence for the evolution of xyloglucanase activity from
xyloglucanendo-transglycosylases: biological implications for cell wall metabolism. Plant Cell
19:1947–1963
Baumann MJ (2007b) Xyloglucan-active enzymes: properties, structures and applications.
Doctoral thesis submitted to School of Biotechnology, Royal Institute of Technology,
Stockholm, pp 10–32
Benko Z, Siika-aho M, Viikari L, Reczey K (2008) Evaluation of the role of xyloglucanase in the
enzymatic hydrolysis of lignocellulosic substrates. Enzyme Microb Technol 43:109–114
Bon EPS, Ferrara MA (2007) Bioethanol production via enzymatic hydrolysis of cellulosic
biomass. In: The role of agricultural biotechnologies for production of bioenergy in developing
countries. FAO. http://www.fao.org/biotech/seminaroct2007.html
Carpita NC, McCann M (2000) The cell wall. In: Buchanan B, Gruissem W, Jones RL
(eds) Biochemistry and molecular biology of plants. American Society of Plant Physiologists,
Rockville, pp 52–108
Chanliaud E, de Silva J, Strongitharm B, Jeronimidis G, Gidley M (2004) Mechanical effects of
plant cell wall enzymes on cellulose/xyloglucan composites RID A-7266-2011. Plant J 38:27–
37
Choudhary J, Saritha M, Nain L, Arora A (2014) Enhanced saccharification of steam-pretreated
rice straw by commercial cellulases supplemented with xylanase. J Bioprocess Biotech 4
(7):188–194
Claassen PAM, van Lier JB, Contreras LAM, van Niel EWJ, Sijtsma L, Stams AJM, de Vries SS,
Weusthuis RA (1999) Utilisation of biomass for the supply of energy carriers. Appl Microbiol
Biotechnol 52:741–755
Damásio ARL, Ribeiro LFC, Ribeiro LF, Furtado GP, Segato F, Almeida FBR, Crivellari AC,
Buckeridge MS, Souza TACB, Murakamie MT, Ward RJ, Prade RA, Polizeli MLTM (2012)
Functional characterization and oligomerization of a recombinant xyloglucan-specific
endo-β-1,4-glucanase (GH12) from Aspergillus niveus. Biochim Biophys Acta 1824:461–467
246 M. Saritha et al.
Edwards M, Dea ICM, Bulpin PV, Reid JSG (1985) Xyloglucan (amyloid) mobilization in the
cotyledons of Tropaeolum majus L seeds following germination. Planta 163:133–140
Enkhbaatar B, Temuujin U, Lim J-H, Chi W-J, Chang Y-K, Hong S-K (2012) Identification and
characterization of a xyloglucan-specific family 74 glycosyl hydrolase from Streptomyces
coelicolor A3(2). Appl Environ Microbiol 78(2):607–611
Ethanol Industry Outlook Report-Battling for the Barrel (2013) Renewable Fuels Association.
http://www.ethanolrfa.org
Faik A, Price NJ, Raikhel NV, Keegstra K (2002) An Arabidopsis gene encoding an
α-xylosyltransferase involved in xyloglucan biosynthesis. Proc Natl Acad Sci USA
99:7797–7802
Farrell AE, Plevin RJ, Turner BT, Jones AD, O’Hare M, Kammen DM (2006) Ethanol can
contribute to energy and environmental goals. Science 113:506–508
Feng T, Yan K-P, Mikkelsen MD, Meyer AS, Schols HA, Westereng B, Mikkelsen JD (2014)
Characterisation of a novelendo-xyloglucanase (XcXGHA) from Xanthomonas that accom-
modates a xylosyl-substituted glucose at subsite-1. Appl Microbiol Biotechnol 98:9667–9679
Fry SC (1989) The structure and functions of xyloglucan. J Exp Bot 40:1–11
Fry SC, Smith RC, Renwick KF, Martin DJ, Hodge SK, Matthews KJ (1992)
Xyloglucanendotransglycosylase, a new wall-loosening enzyme activity from plants.
Biochem J 282:821–828
Fry SC, York WS, Albersheim P, Darvill A, Hayashi T, Joseleau JP, Kato Y, Lorences EP,
Maclachlan GA, McNeil M, Mort AJ, Reid JSG, Seitz HU, Selvendran RR, Voragen AGJ,
White AR (1993) An unambiguous nomenclature for xyloglucan-derived oligosaccharides.
Physiol Plant 89:1–3
Furtado GP, Santos CR, Cordeiro RL, Ribeiro LF, de Moraes LAB, Damásio ARL,
Polizeli MLTM, Lourenzoni MR, Murakami MT, Ward RJ (2015) Enhanced
xyloglucan-specific endo-β-1,4-glucanase efficiency in an engineered CBM44-XegA chimera.
Appl Microbiol Biotechnol 99(12):5095–5107
Girio FM, Fonseca C, Carvalheiro F, Duarte LC, Marques S, Bogel-Lukasik R (2010)
Hemicelluloses for fuel ethanol: a review. Bioresour Technol 101:4775–4800
Gloster TM, Ibatullin FM, Macauley K, Eklof JM, Roberts S, Turkenburg JP, Bjornvad K,
Jorgensen PL, Danielsen S, Joha KS, Borchert TV, Wilson KS, Brumer H, Davies GJ (2007)
Characterization and three-dimensional structures of two Distinct bacterial xyloglucanases
from Families GH5 and GH12. J Biol Chem 282(26):19177–19189
Grishutin SG, Gusakov AV, Markov AV, Ustinov BB, Semenova MV, Sinitsyn AP (2004)
Specific xyloglucanases as a new class of polysaccharide-degrading enzymes. Biochim
Biophys Acta 1674:268–281
Guillen D, Sanchez S, Rodriguez-Sanoja R (2010) Carbohydrate-binding domains: multiplicity of
biological roles. Appl Microbiol Biotechnol 85(5):1241–1249
Hayashi T (1989) Xyloglucans in the primary-cell wall. Annu Rev Plant Physiol Plant Molec Biol
40:139–168
Hayashi T, Kato Y, Matsuda K (1980) Biosynthesis of xyloglucan in suspension-cultured soybean
cells. Plant Cell Physiol 21:1405–1418
Hiloidhari M, Das D, Baruah DC (2014) Bioenergy potential from crop residue biomass in India.
Renew Sust Energ Rev 32:504–512
Hu J (2014) The role of accessory enzymes in enhancing the effective hydrolysis of the cellulosic
component of pretreated biomass. Thesis submitted to The University of British Columbia,
Vancouver, pp 98–106
Ichinose H, Araki Y, Michikawa M, Harazono K, Yaoi K, Karita S, Kaneko S (2012)
Characterization of an endo-processive-type xyloglucanase having a β-1,4-glucan-binding
module and an endo-type xyloglucanase from Streptomyces avermitilis. Appl Environ
Microbiol 78(22):7939–7945
Jabbour D, Borrusch MS, Banerjee G, Walton JD (2013) Enhancement of fermentable sugar yields
by α-xylosidase supplementation of commercial cellulases. Biotechnol Biofuels 6:58–66
9 The Role and Applications of Xyloglucan Hydrolase … 247
Singh S, Pranaw K, Singh B, Tiwari R, Nain L (2014) Production, optimization and evaluation of
multicomponent holocellulase produced by Streptomyces sp. ssr-198. J Taiwan Inst Chem E
45:2379–2386
Sinitsyna OA, Fedorova EA, Pravilnikov AG, Rozhkova AM, Skomarovsky AA, Matys VYu,
Bubnova TM, Okunev ON, Vinetsky YuP, Sinitsyn AP (2010) Isolation and properties of
xyloglucanases of Penicillium sp. Biochemistry (Moscow) 75(1):41–49
Sinnott ML (1990) Catalytic mechanisms of enzymic glycosyl transfer. Chem Rev 90:1171–1202
Song S, Tang Y, Yang S, Yan Q, Zhou P, Jiang Z (2013) Characterization of two novel family 12
xyloglucanases from the thermophilic Rhizomucor miehei. Appl Microbiol Biotechnol 97
(23):10013–10024
Stratilova E, Ait-Mohand F, Rehulka P, Garajova S, Flodrova D, Rehulkova H, Farkas V (2010)
Xyloglucanendotransglycosylases (XETs) from germinating nasturtium (Tropaeolum majus)
seeds: isolation and characterization of the major form. Plant Physiol Biochem 48(4):207–215
Sweeney MD, Xu F (2012) Biomass converting enzymes as industrial biocatalysts for fuels and
chemicals: recent developments. Catalysts 2:244–263
Takahashi M, Yamamoto R, Sakurai N, Nakano Y, Takeda T (2015) Fungal
hemicellulose-degrading enzymes cause physical property changes concomitant with solubi-
lization of cell wall polysaccharides. Planta 241:359–370
Timell TE (1967) Recent progress in the chemistry of wood hemicelluloses. Wood Sci Technol
1:45–70
van den Brink J, de Vries RP (2011) Fungal enzyme sets for plant polysaccharide degradation.
Appl Microbiol Biotechnol 6:1477–1492
Vincken JP, Dekeizer A, Beldman G, Voragen AGJ (1995) Fractionation of xyloglucan fragments
and their interaction with cellulose. Plant Physiol 108:1579–1585
Vincken J-P, York WS, Beldman C, Voragen AGJ (1997) Two general branching patterns of
xyloglucan, XXXG and XXGC. Plant Physiol 114:9–13
Vlasenko E, Schülein M, Cherry J, Xu F (2010) Substrate specificity of family 5, 6, 7, 9, 12, and
45 endoglucanases. Bioresour Technol 101:2405–2411
Warner CD, Go RM, García-Salinas C, Ford C, Reilly PJ (2011) Kinetic characterization of a
glycoside hydrolase family 44 xyloglucanase/endoglucanase from Ruminococcus flavefaciens
FD-1. Enzyme Microb Technol 48(1):27–32
Wyman CE (1996) Handbook on bioethanol: production and utilization. Taylor Francis,
Washington, p 417
Yamatoya K, Shirakawa M (2003) Xyloglucan: structure, rheological properties, biological
functions and enzymatic modification. Current Trends Polym Sci 8:27–72
Yaoi K, Kondo H, Noro N, Suzuki M, Tsuda S, Mitsuishi Y (2004) Tandem repeat of a
seven-bladed β-propeller domain in oligoxyloglucan reducing-end-specific cellobiohydrolase.
Structure 12:1209–1217
Yaoi K, Mitsuishi Y (2004) Purification, characterization, cDNA cloning, and expression of a
xyloglucan endoglucanase from Geotrichum sp. M128. FEBS Lett 560:45–50
Yaoi K, Miyazaki K (2012) Cloning and Expression of isoprimeverose-producing oligoxyloglucan
hydrolase from actinomycetes species, Oerskovia sp. Y1. J Appl Glycosci 59(2):83–88
Yokoyama R, Rose JKC, Nishitani K (2004) A surprising diversity and abundance of
xyloglucanendotransglucosylase/hydrolases in rice. Classification and expression analysis.
Plant Physiol 134:1088–1099
Yuan S, Wu Y, Cosgrove DJ (2001) A Fungal endoglucanase with plant cell wall extension
activity. Plant Physiol 127(1):324–333
Part III
Lignocellulose Oxidureductases
Chapter 10
Role of Mushroom Mn-Oxidizing
Peroxidases in Biomass Conversion
10.1 Introduction
needed for physical treatments, they are useful because they lead to reduction of
particle size, commonly by grinding, pelleting or heating, and in such a way
improve digestibility of lignocellulosic biomass (Kuijk et al. 2015). Chemical
hydrolysis is based on usage of either acids (hydrochloric or sulphuric acid in high
concentrations) or bases (sodium or ammonium hydroxide) or not selective
chemicals (H2O2), and physico-chemical hydrolysis combine heat, moisture,
pressure, and chemicals (Hendriks and Zeeman 2009). However, numerous studies
have shown that these degradation processes are inefficient, very expensive, and
hazardous for the environment. Therefore preference is given to an alternative
environmental and economical friendly biological delignification, which unique
participants are bacteria and fungi owing to their ability to produce lignocellulolytic
enzymes. The main advantages of the process are energy saving and environmental
protection (Kuijk et al. 2015). Nevertheless, the process has specific demands and
some flaws. Namely, it can be realized only under aerobic conditions and effec-
tiveness depends on lignin structure, i.e. type and ratio of its constituents, branching
level, as well as lignin amount. Grabber (2005) showed that highly branched lignin
is less degradable than mainly linear lignin, Skyba et al. (2013) that syringyl-rich
lignin is more resistant to degradation by fungal enzymes, and Arora and Sharma
(2009) that negative correlation exists between lignin amount and dry matter
digestibility. Loss of dry matter by consumption of cellulose and hemicellulose by
fungi, slowness, and requirement for sterile conditions are the main drawbacks of
the process (Kuijk et al. 2015).
Filamentous fungi, among which Basidiomycetes stand out specially, are good
degraders of lignocellulose. According to lignocellulose decay pattern, Eriksson
et al. (1990) separated three groups of fungi: white rot, brown rot and soft rot, and
their activities depend on environmental conditions, plant species, and cell type.
White rot fungi are common in hardwood forests and include several hundred
species of Basidiomycetes and some Ascomycetes, which breakdown lignin
especially during the early phase of colonization, make access to the cellulose and
hemicellulose, and depolimerize them in fruiting stage (Hammel 1997; Kirk and
Cullen 1998; Sánchez 2009; Kuijk et al. 2015). Ginterová and Lazarová (1987)
showed that ratio among degraded lignin, cellulose, and hemicellulose at the end of
wheat straw colonization by Pleurotus ostreatus was 37.4:20:32.3 % and at the end
of fructification 41.8: 62.1: 60.8 %. Some species, such as Phanerochaete
chrysosporium, degrade lignin, cellulose, and hemicellulose simultaneously and
other species, for example Ceriporiopsis subvermispora, attack lignin first. White
rot fungi are the most efficient lignin mineralizators which have ability to cleave
ether bounds in non-phenolic lignin substructures whose proportion in the woody
biomass is 90 %. Contrary to these fungi, brown-rot species that are mainly present
in coniferous forests only modify lignin and their main activity is cellulose and
hemicellulose depolymerisation, while soft rot species degrade only cellulose under
conditions of high moisture and low lignin content (Sánchez 2009).
According to Leonowicz et al. (1999) and Sánchez (2009), white rot fungi
produce three groups of enzymes: (A) lignocellulolytic enzymes, which include
ligninases (lignin peroxidase, Mn-oxidizing peroxidases, laccase, horseradish
254 M. Stajić et al.
10.2.1.1 Structure
Mn-oxidizing peroxidases posses multiple substrate sites at both sides of the heme
pocket and could oxidize different substrates.
However, Fernández-Fueyo et al. (2014) demonstrated significant differences
between Pleurotus ostreatus MnPs and VPs in amino acid sequence length, amino
acid composition and distribution, as well as in substrate specificity and catalytic
efficiency on Mn2+. Thus, MnPs contains sequences composed of 331 and VPs of
339 residues mainly located in C-terminal region. In MnPs number of proline
residues varies from 25 to 26 and in VPs from 30 to 31, while number of lysine
residues is ranged from 7 to 10 in all isoforms except MnP4 where the number is
20. In the case of substrate and ability of Mn2+ oxidation, VPs have wider speci-
ficity and higher efficiency than MnPs. This could be explained by the fact that VPs,
besides Mn2+ oxidation site composing of three conserved acidic residues near the
internal heme propionate, also posses lignin oxidation site near tryptophan that is
substituted by an alanine or aspartic acid in MnPs. Exceptions are P. ostreatus
MnP1 and P. pulmonarius peroxidase which posses tryptophan but not activity
characterized for VPs, that Fernández-Fueyo et al. (2014) explained by change of
shape and charge of surface around tryptophan and in such a way substrate binding.
Analyzing the structures of P. ostreatus MnP4 and VP1 these authors noted a few
differences: (i) in domain close to the Ca2+ ion; (ii) in the main heme access
channel, which is wider in VP1 than in MnP4 due to the presence of three substrate
oxidation sites; (iii) in shape and charge of lower channel lip, as wide reservoir with
negative charge in VP1 and as narrow groove of less charged surface in MnP4;
(iv) in number of lysine residues, MnP4 has the minimum one unit more than VP1,
which influences isoelectic points, and (v) in number of hydrogen bonds and salt
bridges that are more numerous in MnP4 than in VP1.
radicals can cleave Cα–Cβ and β-aryl ether bonds in abundant non-phenolic lignin
substructures. Contrary to fatty acids, which are normally presented in the fungal
environment, thiols are not produced by fungi and SH-groups are released in small
amount by protein degradation during cell lyses (Hofrichter 2002). Likewise, Mn3+
can oxidize the oxalate from chelate generating CO2 and anion radical, which in the
presence of O2 forms another CO2 molecule and superoxide radical (O– 2 ) that at low
wood pH values transforms into perhydroxyl radical (HOO) leading to lipid per-
oxidation and releasing of ligninolytic peroxyl radicals (Fig. 10.1). However, more
efficient mechanism of mineralization of non-phenolic lignin moieties means their
modification by cellobiose dehydrogenase to structures which are accessible to
MnPs. Chelates of Mn3+ with the acids can also react with each other and convert to
alkyl radicals which then react spontaneously with oxygen and form some other
radicals, for example superoxide, that could be used by native enzyme in the
absence of H2O2 (Hofrichter 2002). This author also emphasized that high con-
centrations of H2O2 or O2 lead to formation of Compound III, catalytically inactive
form, which cause reversible inactivation of MnPs.
Martinez et al. (1996) first purified and characterized two VPs after submerged
cultivation of Pleurotus eryngii in glucose/peptone/yeast extract medium. Later
Martinez (2002) showed that these enzymes can oxidize both Mn2+ and aromatic
compounds, and that present hybrids of MnPs and lignin peroxidases. These
enzymes occur naturally and until now it is shown that various species of the genera
Pleurotus and Bjerkandera, Lentinus edodes, Panus tigrinus, etc. can synthesize
them (Hofrichter 2002; Hammel and Cullen 2008).
10 Role of Mushroom Mn-Oxidizing Peroxidases … 259
At the beginning of twenty-first century, humankind faces with high rate of world
population growth and in such a way with food insufficient, environmental pollu-
tion, and climate change. It should be emphasized that 70 % of agricultural prod-
ucts are not utilized and become ballast and potential pollutant of environment.
Numerous fungi have essential role in conversion of this enormous amount of
lignocellulosic waste in food, feed, paper pulp, biofuels, and various chemicals.
However, all of the conversions require delignification without or with minimal loss
of carbohydrates, which could be achieved by selection of the most efficient fungal
species and strains, the optimal lignocellulosic biomass, as well as by optimization
of cultivation conditions (Kuijk et al. 2015). Nutritional value and availability are
the main criteria for selection of lignocellulosic residues. However, composition of
plant waste of the same origin and consequently its degradation can vary because of
existence of numerous cultivars of the same crop, for example number of wheat
cultivars is still grown. It was demonstrated by Arora and Sharma (2009) who noted
that Phlebia floridensis more efficiently delignificated wheat straw from north-
eastern India than straw from other climatic regions. As efficiency of fungal
enzymatic system depends on substrate type, selection of degrader is conditioned
by lignocellulosic residue. Thus, Ceriporiopsis subvermispora, Pleurotus eryngii,
P. ostreatus, Lentinus edodes, Hericium clathroides, and Trametes versicolor are
excellent selective degraders, namely they mineralize more than 20 % of lignin and
less than 5 % of cellulose in wheat straw (Tuyen et al. 2012; Knežević et al. 2013a,
b; Stajić et al. 2013; Kuijk et al. 2015). Numerous studies confirmed that C.
subvermispora and P. eryngii are the most effective mineralizers of wheat straw
lignin and showed that Mn-oxidizing peroxidases have the essential role in the
process of delignification (Fernandez-Fueyo et al. 2012; Tuyen et al. 2012).
However, Asiegbu et al. (1996) and Chi et al. (2007) demonstrated that combi-
nation of few fungal species can be more effective in depolymerization of lignin in
some plant wastes than their monocultures. Thus, extent of spruce sawdust delig-
nification by co-culture of Pleurotus sajor-caju, T. versicolor, and Phanerochaete
chrysosporium was significantly higher than by each monoculture separately (16
and 0–5 %, respectively), and co-culture of C. subvermispora and P. ostreatus
caused higher loss of lignin from aspen wood than in its colonization with the
monocultures. However, due to divergence in genotype, significant differences in
lignocellulolytic enzyme production and selectivity of lignocellulose degradation
exist among strains of the same mushroom species (Simonić et al. 2010; Knežević
et al. 2013b). For example, percents of degraded wheat straw lignin and cellulose
by some P. ostreatus strains were almost the same (Salvachua et al. 2011), while
other strains predominantly depolymerized lignin (Adamović et al. 1998). Solid
state fermentation, small lignocellulosic particles of suitable shape, high oxygen
level, moderate water content, certain inoculum type (spores or spawn) in moderate
amount, and the presence of copper, manganese, linoleic acid, and veratryl alcohol
260 M. Stajić et al.
are good conditions for biological treatment of lignocellulose by white rot fungi
(Kuijk et al. 2015).
10.3.1 Food
between 123 and 262 kg of fruiting bodies per ton of wheat straw (Kuijk et al.
2015).
Nowadays, number of cultivated mushroom species is relatively small, 100
species are cultivated economically, 60 commercially, and only 10 in industrial
scale, but current trend is growth of world production in favor of new species, such
as Agaricus blazei, Agrocybe aegerita, Armillaria mellea, Auricularia fuscosuc-
cinea, Coprinus comatus, Dictyophora duplicata, Lepista nuda, Stropharia
rugosaannulata, Tremella cinnabarina, T. aurantialba, and Tricholoma giganteum
(Chang and Miles 2004).
10.3.2 Feed
Most agricultural and industrial wastes that are used for animal feeding are rich in
low digestible fibers and poor in nutrients, especially in proteins and vitamins.
Besides high nutritionally valued food, result of mushroom cultivation is also
production of enormous amount of spent substrate that can be an important feed
ingredient. Mushroom mycelium is rich in protein, essential amino acids, and chitin
and could be important nitrogen sources, while β-glucans and other extra- and
intracellular polysaccharides could be an additional glucose sources and
immunostimulators for ruminants (Reis et al. 2012; Cheung 2013). Thus, protein
content in wheat straw after P. ostreatus cultivation can be increased by 89 % that
leads to overcoming of the main limitation of straw feed, i.e. nitrogen lack (Kuijk
et al. 2015). The authors also emphasized that the protein increases rumen
microorganism population and in such a way cellulose and hemicellulose
digestibility and energy releasing. Spend mushroom compost is also rich in vita-
mins, minerals, antimicrobial agents, probiotics, and other biologically active
compounds that modify animal metabolism, enhance growth, improve immune
system, and increase resistance to various diseases (Enshasy and Hatti-Kaul 2013).
Mushroom contribution to feed production is also transformation of natural
recalcitrant lignin into digestible compounds owing to well developed lignocellu-
lolytic enzyme system, especially Mn-oxidizing peroxidases. Although ruminants
posses cellulolytic microorganisms in rumen that can depolymerize cellulose and
hemicellulose from plant cell wall, connection of these carbohydrates for lignin by
covalent or other linkages is negatively correlated with their digestibility under
rumen anaerobic conditions. Rodrigues et al. (2008) showed that cultivation of
various white rot mushrooms on wheat straw could increase degradability of neutral
detergent fibers in rumen in even 13 % due to their strong ligninolytic enzymes
which act to lignin. Contrary to wheat straw that was not biologically treated, spent
wheat straw composts of Pleurotus ostreatus and P. sajor-caju were well consumed
and digested by sheep (Calzada et al. 1987; Fazaeli et al. 2006). Similar efficient
consumption and decomposition of crude protein, hemicellulose, and cellulose from
spend Coprinus fimetarius rice straw compost and spent Ganoderma sp. wheat
straw compost by goats were reported by Rai et al. (1989) and Shrivastava et al.
262 M. Stajić et al.
(2012). However, cows are more selective and their feed can be enriched with
maximum 17 % of spend P. ostreatus wheat straw substrate (Adamović et al.
1998).
Kuijk et al. (2015) reviewed results of numerous studies and also reported that
addition of selected microelements and inducers (copper, manganese, linoleic acid,
di-rhamnolipid, and veratryl alcohol) to lignocellulosic residues enhance activity of
fungal Mn-oxidizing peroxidases and indirectly fiber digestibility by animals.
However, some mushroom metabolites could be toxic for animals and humans and
special attention has to be given to that aspect of plant waste bioconversion by
fungi.
Since nowadays forest preservation is compulsory, plant residues that are produced
in enormous amount, but not used, can be important natural sources for paper-
making. Good sample for extraordinary production of some agricultural wastes was
given by González Alriols et al. (2009). Namely, they noted huge quantity of
unutilized Elaeis guineensis residues after oil extraction in Malaysia, even 15
million tons per year, which present serious ballast. In a few last decades, white rot
fungi or their enzymes have the main role in the process, known as biopulping. The
main reasons for that are its healthy, environmental, and economical friendly
relationships contrary to traditional usage of alkaline sulfide, as well as improve-
ment of pulp quality and quantity (Bajpai et al. 2001). Yang et al. (2008) showed
that pretreatment of eucalyptus woody biomass with Trametes hirsuta led to
increasing fiber internal bond strength in 32 %. Fungal ligninolytic enzymes, pri-
marily peroxidases, catalyze one-electron oxidation of lignin phenolic groups
producing phenoxy radicals on fiber surfaces and in such a way enhancing inter
fiber adhesion (Singh and Singh 2014). Thus, this modified lignin could replace
traditionally harmful adhesives that is current trend and one more possible appli-
cation of fungal enzymatic system.
Besides enhancement of the bond strength and fiber adhesion, wood pretreat-
ment with white rot fungi removes pitches rich in lipophylic compounds (resin and
fatty acids and triglycerides), which can reduce paper and effluent quality and cause
technical problems (Martinez-Inigo et al. 2000; Dorado et al. 2001; Gutiérrez et al.
2006; Van Beek et al. 2007). Martinez-Inigo et al. (2000) reported that Phlebia
radiate and Poria subvermispora were the most active in the removal of lipophylic
extractives and in such a way pitch from eucalypt wood, and Dorado et al. (2001)
that Bjerkandera sp. and Trametes versicolor eliminated up to 90 % of these
compounds from Pinus sylvestris wooden biomass. Letter Van Beek et al. (2007)
showed that T. versicolor was also effective in removal of resin acids and
triglycerides from spruce wood chips, which content was reduced in 40 and 100 %,
respectively, and that Pycnoporus cinnabarinus caused losses of sterols, resin acids,
and triglycerides from various pulp systems in 65–100 %. Further study showed
10 Role of Mushroom Mn-Oxidizing Peroxidases … 263
that this ability of white rot fungi is based on production of lipase, resinase,
lipoxigenases, and laccases (Gutiérrez et al. 2009). As pitch control is combined
with lignin removal from the pulp, white rot fungi play important role in pulp
bleaching (bio-bleaching) and thus in improving paper brightness, environment
protection, and energy saving (Bajpai et al. 2001; Jerusik 2010). However, the
process should be controlled in order to avoid cellulose degradation.
Besides papermaking, white rot fungi and their enzymes have one more role,
bioremediation of waste water. Namely, huge amount of water, approximately 45–
227 × 103 litters, is used for production of a ton of product resulting in the
releasing of the same quantity of effluent and further in zooplankton and fish dying,
slime and scum formation, esthetic problems, and disruption of the environment
(Pokhrel and Viraraghavan 2004). Fungi are characterized by better survival in
effluents and production of strong extracellular ligninolytic enzymes, especially
lignin- and Mn-oxidizing peroxidases that degrade various phenolic compounds
among which colors are the most resistant. Thus, numerous studies showed high
effectiveness of Trametes pubescens, Phanerochaete chrysosporium, and Pleurotus
ostreatus in the process (Prasad and Gupta 1997; Choudhury et al. 1998; González
et al. 2010; Zhang et al. 2012). Prasad and Gupta (1997), Choudhury et al. (1998),
and Saxena and Gupta (1998) noted high level of color degradation, even 80–
93.8 %, in effluents from paper industry treated by T. versicolor, Ph. chrysosporium
(sole or in combination with Pycnoporus sanguineus), Pleurotus ostreatus, and
Heterobasidion annosum.
10.3.4 Biofuels
Miscanthus sinensis, which production is 5–6 times higher than cereal one, or from
dried citrus waste that amount is about 800,000 tons only in USA (Zechendorf
1999; Tabka et al. 2006). However, in order to cellulose to be broken down to the
glucose and further ferment, depolymerization of recalcitrant network around cel-
lulose fibers, built of lignin, is necessary. Thus, it comes again to white rot fungi
and their peroxidases which are the main participants in the first step of bioethanol
production known as pretreatment, i.e. the essential catalyzers of beginning phase
of lignin degradation. This pretreatment also depends on fungal species and strain,
cultivation conditions, enzyme production and activity, and oxidative mechanisms
(Dias et al. 2010; Wan and Li 2010; Salvachua et al. 2011). Dias et al. (2010)
reported that various white rot fungi degraded between 2 and 65 % of wheat straw
lignin during its solid state fermentation and in such a way increased
cellulose/lignin ratio. However, Salvachua et al. (2011) emphasized that mode of
lignocellulosic mineralization is different from species to species, for example
contrary to Panus tigrinus and Phlebia radiata that degraded lignin and polysac-
charides simultaneously, Pleurotus eryngii and Phellinus robustus removed lignin
selectively and faster. Likewise, according to these authors, delignification extent is
not always in correlation with ligninolytic enzymes production, as well as with fiber
digestibility and sugar production. Thus, contrary to P. radiata and P. robustus,
characterized both by high Mn-oxidizing peroxidase activities and effective lignin
removal, good lignin degraders Bjerkandera adusta and Coriolopsis rigida had low
active enzymes. In the case of effectiveness of cellulose and hemicellulose enzy-
matic hydrolysis in the second step of bioethanol production, Pycnoporus coccineus
was high effective hemicellulose and weak cellulose degrader (98 and 31 %,
respectively), Bjerkandera adusta was almost equal depolymerizer of these poly-
mers (43 and 54 %, respectively), Stereum hirsutum mineralized only significant
portion of cellulose (43 %), the highest glucose yield was obtained by cultivation of
Poria subvermispora (69 %) and Irpex lacteus (66 %) and the lowest by P. eryngii
and P. robustus.
Namely, one enzyme rarely possesses all catalytic properties required for effective
lignocellulosic conversion. Therefore recombination can be used for unification of
several beneficial characteristics from a few enzymes into superior one with high
stability and broad substrate promiscuity. Finding of appropriate organism for
expression of the gene leads to obtaining of significant amount of the recombinant
(Alcalde 2015). Thus, Ogawa et al. (1998) showed that expression of mnp genes of
Pleurotus ostreatus, excellent Mn-oxidizing peroxidase producer, in Coprinus
cinereus, characterized with fast growth, led to increase of lignin degradation level
in only 16 days, and Pérez-Boada et al. (2002) and Salame et al. (2012) obtained
active versatile peroxidase by expression of mnp4 gene in Escherichia coli but in
lower yield. Considering that lignocellulolytic enzymes act synergistically, genetic
engineering could be used for fusion of certain enzymes and also for generation of
strain deficient in some enzyme. With this aim, Chalamcherla et al. (2010) getting
cellulase-deficient P. ostreatus strain characterized by low cellulose and high lignin
degradation, as well as by high capacity of application in various biotechnological
processes. Nowadays, special attention is also given to optimization of cultivation
conditions for production of certain lignocellulolytic enzyme and study of enzy-
matic system of fungal species living in extreme environment as well as their genes
encoding proteins with unknown function among which some can be lignocellu-
lolytic enzyme (Guerriero et al. 2015). Everything mentioned show great possi-
bilities of usage and managing of fungi and their enzymes with the aim of biomass
conversion into high-valued products and environment protection.
References
Calzada JF, Franco LF, de Arriola MC, Rolz C, Ortiz MA (1987) Acceptability, body weight
changes and digestibility of spent wheat straw after harvesting of Pleurotus sajor-caju. Biol
Waste 22:303–309
Camarero S, Sarkar S, Ruiz-Dueñas FJ, Martinez MJ, Martinez AT (1999) Description of a
versatile peroxidase involved in natural degradation of lignin that has both Mn-peroxidase and
lignin-peroxidase substrate binding sites. J Biol Chem 274:10324–10330
Chalamcherla VL, Singaracharya MA, Lakshmi MV (2010) Amino acids profile of the
lignocellulosic feed treated with cellulase-free lignolytic mutants of Pleurotus ostreatus.
Bioresources 5:259–267
Chang ST, Miles PG (2004) Mushrooms. Cultivation, nutritional value, medicinal effect, and
environmental impact. CRC Press LLC, London
Cheung PCK (2013) Mini-review on edible mushrooms as source of dietary fiber: preparation and
health benefits. Food Sci Hum Well 2:162–166
Chi YJ, Hatakka A, Maijala P (2007) Can co-culturing of two white-rot fungi increase lignin
degradation and the production of lignin-degrading enzymes? Int Biodeter Biodegr 59:32–39
Choudhury S, Sahoo N, Manthan M, Rohela RS (1998) Fungal treatment of pulp and paper mill
effluents for pollution control. J Ind Pollut Control 14:1–13
Datta A, Bettermann A, Kirk TK (1991) Identification of a specific manganese peroxidase among
ligninolytic enzymes secreted by Phanerochaete chrysosporium during wood decay. Appl
Environ Microbiol 57:1453–1460
Demirbas A (2009) Progress and recent trends in biodisel fuels. Energ Convers Manag 50:14–34
Dias AA, Freitas GS, Marques GSM, Sampaio A, Fraga IS, Rodrigues MAM, Evtuguin DV,
Bezerra RMF (2010) Enzymatic saccharification of biologically pre-treated wheat straw with
white-rot fungi. Bioresource Technol 101:6045–6050
Dorado J, van Beek TA, Claassen FW, Sierra-Alvarez R (2001) Degradation of lipophilic wood
extractive constituents in Pinus sylvestris by the white rot fungi Bjerkandera sp. and Trametes
versicolor. Wood Sci Technol 35:117–125
Enshasy HAE, Hatti-Kaul R (2013) Mushroom immunomodulators: unique molecules with
unlimited applications. Trends Biotechnol 31:668–677
Eriksson KEL, Blanchette RA, Ander P (1990) Microbial and enzymatic degradation of wood and
wood components. Springer, Berlin
Evans CS, Gallagher IM, Atkey PT, Wood DA (1991) Localisation of degradative enzymes in
white-rot decay of lignocellulose. Biodegradation 2:93–106
Fazaeli H, Azizi A, Amile M (2006) Nutritive value index of treated wheat straw with Pleurotus
fungi fed to sheep. Pakistan J Biol Sci 9:2444–2449
Fernandez-Fueyo E, Ruiz-Dueñas FJ, Ferreira P, Floudas D, Hibbett DS, Canessa P et al (2012)
Comparative genomics of Ceriporiopsis subvermispora and Phanerochaete chrysosporium
provide insight into selective ligninolysis. Proc Nat Acad Sci USA 109:5458–5463
Fernández-Fueyo E, Ruiz-Dueñas FJ, Martínez MJ, Romero A, Hammel KE, Medrano FJ,
Martínez AT (2014) Ligninolytic peroxidase genes in the oyster mushroom genome:
heterologous expression, molecular structure, catalytic and stability properties, and
lignin-degrading ability. Biotechnol Biofuels 7:2
Ginterová A, Lazarová A (1987) Degradation dynamics of lignocellulose materials by
wood-rotting Pleurotus fungi. Folia Microbiol 32:434–437
González Alriols M, Tejado A, Blanco M, Mondragon I, Labidi J (2009) Agricultural palm oil tree
residues as raw material for cellulose, lignin and hemicelluloses production by ethylene glycol
pulping process. Chem Eng J 148:106–114
González LF, Sarria V, Sánchez OF (2010) Degradation of chlorophenols by sequential
biological-advanced oxidative process using Trametes pubescens and TiO2/UV.
BioresourceTechnol 101:3493–3499
Grabber JH (2005) How do lignin composition, structure, and cross-linking affect degradability? A
review of cell wall model studies. Crop Sci 45:820–831
Guerriero G, Hausman JF, Strauss J, Ertan H (2015) Destructuring plant biomass: focus on fungal
and extremophilic cell wall hydrolases. Plant Sci 234:180–193
10 Role of Mushroom Mn-Oxidizing Peroxidases … 267
Medeiros BM, Bento VA, Nunes LLA, Oliveira CS (1999) Optimization of some variables that
affect the synthesis of laccase by Pleurotus ostreatus. Bioprocess Eng 21:483–487
Muñoz C, Guillen F, Martínez TA, Martínez JM (1997a) Induction and characterization of laccase
in the ligninolytic fungus Pleurotus eryngii. Cur Microbiol 34:1–5
Muñoz C, Guillen F, Martínez TA, Martínez JM (1997b) Laccase isoenzymes of Pleurotus
eryngii: characterization, catalytic properties, and participation in activation of molecular
oxygen and Mn2+ oxidation. Appl Environ Microbiol 63:2166–2174
Nigam PS, Singh A (2011) Production of liquid biofuels from renewable resources. Prog Energ
Combust 37:52–68
Ogawa K, Yamazaki T, Kajiwara S, Watanabe A, Asada Y, Shishido K (1998) Molecular breeding
of the basidiomycete Coprinus cinereus strains with high lignin-decolorization and -
degradation activities using novel heterologous protein expression vectors. Appl Microbiol
Biotechnol 49:285–289
Pérez-Boada M, Doyle WA, Ruiz-Dueñas FJ, Martínez MJ, Martínez AT, Smith AT (2002)
Expression of Pleurotus eryngii versatile peroxidase in Escherichia coli and optimisation of
in vitro folding. Enzyme Microb Tech 30:518–524
Perez-Boada M, Ruiz-Dueñas FJ, Pogni R, Basosi R, Choinowski T, Martinez MJ, Piontek K,
Martinez AT (2005) Versatile peroxidase oxidation of high redox potential aromatic
compounds: site-directed mutagenesis, spectroscopic and crystallographic investigation of
three long-range electron transfer pathways. J Mol Biol 354:385–402
Pokhrel D, Viraraghavan T (2004) Treatment of pulp and paper mill wastewater—a review. Sci
Total Environ 333:37–58
Prasad GK, Gupta RK (1997) Decolourization of pulp and paper mill effluent by two white-rot
fungi. Indian J Environ Health 39:89–96
Rai SN, Walli TK, Gupta BN (1989) The chemical composition and nutritive value of rice straw
after treatment with urea or Coprinus fimetarius in a solid state fermentation system. Anim
Feed Sci Tech 26:81–92
Reis FS, Barros L, Martins A, Ferreira ICFR (2012) Chemical composition and nutritional value of
the most widely appreciated cultivated mushrooms: an inter-species comparative study. Food
Chem Toxicol 50:191–197
Rodrigues MAM, Pinto P, Bezerra RMF, Dias AA, Guedes CVM, Cardoso VMG et al (2008)
Effect of enzyme extracts isolated from white-rot fungi on chemical composition and in vitro
digestibility of wheat straw. Anim Feed Sci Technol 141:326–338
Ruiz-Dueñas FJ, Guillen F, Camarero S, Perez-Boada M, Martinez MJ, Martinez AT (1999)
Regulation of peroxidase transcript levels in liquid cultures of the ligninolytic fungus Pleurotus
eryngii. Appl Environ Microbiol 65:4458–4463
Ruiz-Dueñas FJ, Morales M, Garcia EMY, Martinez MJ, Martinez AT (2009) Substrate oxidation
sites in versatile peroxidase and other basidiomycetes peroxidases. J Exp Bot 60:441–452
Salame TM, Knop D, Levinson D, Mabjeesh SJ, YardenO Hadar Y (2012) Release of Pleurotus
ostreatus versatile-peroxidase from Mn2+ repression enhances anthropogenic and natural
substrate degradation. PLoS ONE 7:e52446
Salvachua D, Prieto A, Lopez-Abelairas M, Lu-Chau T, Martinez AT, Martinez MJ (2011) Fungal
pretreatment: an alternative in second-generation ethanol from wheat straw. Bioresource
Technol 102:7500–7506
Sánchez C (2009) Lignocellulosic residues: biodegradation and bioconversion by fungi.
Biotechnol Adv 27:185–194
Saxena N, Gupta RK (1998) Decolourization and delignification of pulp and paper mill effluent by
white rot fungi. Indian J Exp Biol 36:1049–1051
Shrivastava B, Nandal P, Sharma A, Jain KK, Khasa YP, Das TK, Mani V, Kewalramani NJ,
Kundu SS, Kuhad RC (2012) Solid state bioconversion of wheat straw into digestible and
nutritive ruminant feed by Ganoderma sp. rckk02. Bioresource Technol 107:347–351
Simonić J, Vukojević J, Stajić M, Glamočlija J (2010) Intraspecific diversity within Ganoderma
lucidum in laccase and Mn-dependent peroxidases production during plant residues
fermentation. Appl Biochem Biotechnol 162:408–415
10 Role of Mushroom Mn-Oxidizing Peroxidases … 269
Singh AP, Singh T (2014) Biotechnological applications of wood-rotting fungi: a review. Biomass
Bioenerg 62:198–206
Skyba O, Douglas CJ, Mansfield SD (2013) Syringyl-rich lignin renders poplars more resistant to
degradation by wood decay fungi. Appl Environ Microbiol 79:2560–2571
Stajić M, Persky L, Friesem D, Hadar Y, Wasser SP, Nevo E, Vukojević J (2006a) Effect of
different carbon and nitrogen sources on laccase and peroxidases activity by selected Pleurotus
species. Enzyme Microb Tech 38:65–73
Stajić M, Persky L, Hadar Y, Friesem D, Duletić-Laušević S, Wasser SP, Nevo E (2006b) Effect of
copper and manganese ions on activities of laccase and peroxidases in three Pleurotus species
grown on agricultural wastes. Appl Biochem Biotechnol 128:87–96
Stajić M, Vukojević J, Duletić-Laušević S (2009) Biology of Pleurotus eryngii and the role in
biotechnological processes: a review. Crit Rev Biotechnol 29:55–66
Stajić M, Kukavica B, Vukojević J, Simonić J, Veljović-Jovanović S, Duletić-Laušević S (2010)
Wheat straw conversion by enzymatic system of Ganoderma lucidum. Bioresources 5:2362–
2373
Stajić M, Vukojević J, Knežević A, Milovanović I (2013) Influence of trace elements on
ligninolytic enzyme activity of Pleurotus ostreatus and P. pulmonarius. Bioresources 8:3027–
3037
Stuardo M, Vasquez M, Vicuna R, Gonzalez B (2004) Molecular approach for analysis of model
fungal genes encoding ligninolytic peroxidases in wood-decaying soil system. Lett Appl
Microbiol 38:43–49
Tabka MG, Herpoel-Gimbert I, Monod F, Asther M, Sigoillot JC (2006) Enzymatic sacchari-
fication of wheat straw for bioethanol production by a combined cellulose xylanase and
feruloyl esterase treatment. Enzyme Microb Tech 39:897–902
Thurston C (1994) The structure and function of fungal laccase. Microbiol 140:19–26
Tuyen VD, Cone JW, Baars JJP, Sonnenberg ASM, Hendriks WH (2012) Fungal strain and
incubation period affect chemical composition and nutrient availability of wheat straw for
rumen fermentation. Bioresource Technol 111:336–342
Urzúa U, Larrondo LF, Lobos S, Larrain J, Vicuña R (1995) Oxidation reactions catalyzed by
manganese peroxidase isoenzymes from Ceriporiopsis subvermispora. FEBS Lett 371:132–
136
Van Beek TA, Kuster B, Chassen FW, Tienvieri T, Bertaud F, Lenon G, Petit-Conil M,
Sierra-Alvarez R (2007) Fungal biotreatment of spruce wood with Trametes versicolor for
pitch control: influence of extractive contents, pulping process parameters, paper quality and
effluent toxicity. Bioresource Technol 98:302–311
Varela E, Guillén F, Martínez AT, Martínez MJ (2001) Expression of Pleurotus eryngii
aryl-alcohol oxidase in Aspergillus nidulans: purification and characterization of the
recombinant enzyme. Biochim Biophys Acta 1546:107–113
Wan C, Li Y (2010) Microbial pretreatment of corn stover with Ceriporiopsis subvermispora for
enzymatic hydrolysis and ethanol production. Bioresource Technol 101:6398–6403
Wariishi H, Akilswaran L, Gold MH (1988) Manganese peroxidase from the basidiomycetes
Phanerochaete chrysosporium: spectral characterization of the oxidized states and the catalytic
cycle. Biochemistry 27:5365–5370
Yang Q, Zhan H, Wang S, Fu S, Li K (2008) Modification of Eucalyptus CTMP fibers with white
rot fungus Trametes hirsute—effects on fiber morphology and paper physical strengths.
Bioresource Technol 99:8118–8124
Yang B, Wyman C (2008) Pre-treatment: the key to unlocking low-cost cellulosic ethanol.
Biofuels Bioprod Bioref 2:26–40
Zechendorf B (1999) Sustainable development: How can biotechnology contribute? Feature
17:219–225
Zhang S, Jiang M, Zhou Z, Zhao M, Li Y (2012) Selective removal of lignin in steam-exploded
rice straw by Phanerochaete chrysosporium. Int Biodeter Biodegr 75:89–95
Chapter 11
Role and Application of Versatile
Peroxidase (VP) for Utilizing
Lignocellulose in Biorefineries
Abstract Within the last few years, the relatively new heme peroxidase, versatile
peroxidase (VP), attracted some attention as a model enzyme and as an industrial
biocatalyst. VPs are interesting for efficient delignification processes due to their
ability to degrade a broad spectrum of recalcitrant substrates without using medi-
ators. Lignin is very complex, highly aromatic, and one of the major compounds of
lignocellulose, which is an important feedstock for biorefineries, particularly when
originated from wood. In utilizing lignocellulose (e.g., second generation biofuels),
lignin modification, and removal must first be addressed. Moreover, lignin degra-
dation for value-added bio-products is also of great interest. An industrial imple-
mentation of VPs would be beneficial in this context, but involves several scientific
and technical challenges.
11.1 Introduction
importance, with the latter of these being of high economic potential and therefore
particularly interesting (Holladay et al. 2007; Huang and Ramaswamy 2013).
However, the contributing factors affecting the probability of their successful
implementation include the structural complexity of native lignin, its recalcitrance,
and its uniqueness (the woody plant’s fingerprint) leading to inconsistent, poly-
disperse product streams. These facts still constitute a major scientific and technical
challenge (Holladay et al. 2007). Consequently, further fundamental research and
development is needed for lignin degradation in both general and targeted usage.
Prerequisites for efficient lignin modification and degradation in biorefineries
are:
(i) Conservation of important lignin related aromatic structures.
(ii) Subsequent processes like saccharification of cellulose by cellulases should
not be negatively influenced.
(iii) Mild conditions.
(iv) Moderate waste streams/effluent treatment and moderate operating costs.
Ligninolytic heme peroxidases (POXs) have gained increased importance in this
context, with particular attention on the use of versatile peroxidases (VPs). VPs are
relatively new POXs and were first identified as MnPs in the late nineties. Within
the last few years they have attracted attention “as a model enzyme and as a source
of industrial/environmental biocatalyst” (Ruiz-Dueñas et al. 2009a). The main
reason for this is their capability to digest/convert a broad spectrum of complex
substrates, ranging from azo dyes like RB5 to (non)phenolic lignin, without the
necessity of mediators (Pogni et al. 2005; Martínez et al. 2009). Nonetheless,
several factors hinder their short-term commercialization and industrial imple-
mentation. This chapter outlines the main issues for industrial applications, starting
first with a brief overview of the involvement of POXs in natural delignification.
Fig. 11.1 Summary of the natural degradation process of lignocellulosic biomass and the
importance of ligninolytic heme peroxidases (modified based on Dashtban et al. 2010). Here, the
reactions of the first POX intermediates are demonstrated. Lac laccase, LiP lignin peroxidase, MnP
manganese peroxidase, VP versatile peroxidase; Lac-I, LiP-I, MnP-I, and VP-I are corresponding
protein radicals, generally known as compound I. S stands for an appropriate substrate of low
molecular weight resulting in relatively reactive radicals which ultimately mediate the cleavage of
lignin bonds and lignin modifications
Direct degradation refers to when the ligninolytic enzyme is in its active state
(characterized by an appended suffix-I) and directly attacks native lignin linkages.
However, when low molecular weight (LMW) substrate (S) is present/provided,
this interposes an intermediate step where enzyme action results first in
substrate-related diffusible radicals mediating nonenzymatic reactions, which then
cause the cleavage of lignin bonds. This is called indirect lignin degradation,
because the LMW substrates serve as so-called mediators. Lac rely on phenolic
mediators for promoting their oxidation capabilities “towards more recalcitrant
non-phenolic lignin compounds” (Cañas and Camarero 2010). For oxidative ability,
the enzyme redox potential is a limited criterion (Ayala 2010). Lac exhibits lower
redox potential properties (≤ 0.8 V) than POX (> 1 V) (Bourbonnais and Paice
1990; Cañas and Camarero 2010). Lignin comprises 80–90 % non-phenolic moi-
eties, where just 10–20 % are phenolics (Kawai et al. 1999). The cleavage of
non-phenolics is thus essential for efficient lignin modification and degradation. The
redox potential of non-phenolic lignin compounds is also > 1 V (Cañas and
Camarero 2010). Consequently, POX have recently gained increased attention for
delignification purposes (Cañas and Camarero 2010) and received higher interest as
11 Role and Application of Versatile Peroxidase (VP) for Utilizing… 275
As initially mentioned, VPs are relative new POXs and are attractive (model)
biocatalysts (Ruiz-Dueñas et al. 2009a). One key advantage is their independence
from mediators. For industrial application, mediators are used in high amounts.
Therefore, mediators must be (i) cheap, (ii) non-toxic, (iii) readily available,
276 N. Busse and P. Czermak
(iv) in large quantities, and/or (v) easy recyclable (Rodríguez Couto and Toca
Herrera 2006; Cañas and Camarero 2010).
Research work concerning delignification processes by VPs are limited, particu-
larly in kinetic studies with representative lignin model compounds (LMCs)
like adlerol (1-(3,4-dimethoxyphenyl)-2-(2-methoxy-phenoxy)-1,3-propanediol), a
non-phenolic β-O-4 lignin model dimer. The reason for this is the underlying, highly
complex, POX reaction mechanism which will be demonstrated in the following, prior
to focusing on mediator-independent adlerol degradation.
! E ½Fe
k1
E0 ½Fe3 þ þ H2 O2 I 4þ
¼ OP þ þ H2 O ð11:1Þ
with k as the reaction rate constant and the index for describing the pathway
number.
Fig. 11.2 Simplified “POX ping–pong” mechanism. E0–EII represents the three main peroxidase
oxidation states. The symbols S and S•+ stand for an appropriate substrate (e.g., lignin, lignin
model compounds such as adlerol) and its corresponding radical cation. The reaction will be
initiated by H2O2 (pathway 1), and followed by two consecutive, one-electron reduction steps
(pathway 2–3). Reprinted from Busse et al. (2013) with slight modification
11 Role and Application of Versatile Peroxidase (VP) for Utilizing… 277
! E ½Fe
k2
EI ½Fe4 þ ¼ OP þ þ S II 4þ
¼ O þ S þ ð11:2Þ
! E ½Fe
k3
EII ½Fe4 þ ¼ O þ S 0 3þ
þ S þ þ H2 O ð11:3Þ
In Eq. (11.2), substrates causing EI reduction can also reduce EII (Eq. 11.3) as
described by Dunford (2010) for horseradish peroxidase (HRP). The same applies
for ligninolytic enzymes, such as LiP, and phenolic substrates (Schoemaker 1990).
Here, the overall net enzyme reaction can be simply expressed by Eq. (11.4).
! 2S
r
þ
2S + H2 O2 þ 2H2 O ð11:4Þ
Assuming the POX cycle is not inhibited, and considering k3 is rate limiting
(k3 = kcat because of k2 ≫ k3), r results in Eq. (11.6) (Rasmussen et al. 1995).
X
II
½ET ¼ ½Ei ð11:7Þ
i¼0
Fig. 11.3 Schematic drawing of nonenzymatic radical reactions–radical chain reactions. The
reaction is initiated by the cation radical intermediate S•+, a product of enzyme activity
(e) Ether-bond cleavage [e.g., addition of solvent (H2O)] (Palmer et al. 1987).
(f) Hydroxylation of benzylic methylene groups, e.g., through oxygen incorpo-
ration via O2 or H2O in the solvent (Fig. 11.3).
(g) Phenol formation by nucleophilic attack, e.g., the addition of a solvent such as
H2O (Palmer et al. 1987).
(h) Aromatic ring cleavage caused by reactions with perhydroxy radicals HOO•/
HO•2 (Fig. 11.3); details will follow in the next section (Palmer et al. 1987).
The highly reactive free-radical species (in Fig. 11.3 denoted as R• and R•1) can bind
to molecular oxygen (O2), forming final products as well as superoxide anion
radicals (O•−2 ) (Schoemaker 1990) via degradation of organic peroxy radical
intermediates (ROO• or R1OO• in Fig. 11.3) (Palmer et al. 1987). O•− 2 ultimately
undergoes rapid disproportionation forming H2O2 and O2 (Harman et al. 1986;
Schoemaker 1990). Here, the O•− 2 is in a pH-dependent equilibrium with its pro-
tonated counterpart HOO• (Bielski et al. 1985). At low pH conditions, HOO•
predominates, existing as a powerful oxidant (Halliwell and Gutteridge 1985) and
thus contributing to the substrate degradation processes. This occurs because
additional bond cleavage is initiated as a consequence of the substrate (S) oxidizing
to its radical cation, where HOO• is reduced to H2O2 (Palmer et al. 1987).
Alternatively, R1OO• can further react with substrate molecules RH, resulting in
new free radical R• and organic peroxide R1OOR′ formation. If R′ represents
hydrogen, organic hydroperoxide (R1OOH) will instead be produced. As shown in
Fig. 11.3, R1OOH competes with the co-substrate H2O2 for reaction with E0–EI and
with S for EI conversion, depending on their relative concentrations. Reactions of
R1OOH with EI lead to undesired enzyme inactivations. In a low-oxygen atmo-
sphere, R1OOH formation diminishes (Acosta et al. 1988), but polymerization
increases (Palmer et al. 1987). R• products may also attack the peroxidase, again
leading to inactivation (Nicell et al. 1993). However, these radicals can also
undergo dimerization and polymerization processes (e.g., reaction with other free
radicals such as R•1) (Palmer et al. 1987), or can act as suitable substrates S for
maintaining the catalytic cycle. The latter is, however, limited in applicability for
radical cation intermediates S•+, as demonstrated later.
With consideration of further enzymatic reactions, both O•− •
2 and HOO (de-
pending on pH) can additionally serve as electron donors, instead of reducing
substrates (competitive reaction). Consequently, all three enzyme states, EI, EII, and
EIII may be, for instance, reduced by O•−2 under O2 production (Bielski et al. 1985;
Kettle et al. 2007) according to Eqs. (11.8)–(11.11), based on Kettle et al. (2007).
The reaction in Eq. (11.9) is supposed to prevent enzyme inhibition caused by poor
reducing substrates, in order to maintain activity (Kettle et al. 2007).
EI þ O
2 ! E þ O2
II
ð11:8Þ
EII þ O
2 þH
þ
! E0 þ O2 þ H2 O ð11:9Þ
280 N. Busse and P. Czermak
EIII þ O
2 þ 2H
þ
! E0 þ H2 O2 þ O2 ð11:10Þ
or
EIII þ O
2 þ 2H
þ
! EI þ H2 O þ O2 ð11:11Þ
At this point, it becomes obvious that peroxidase activity at least leads to radical
cation formation by reacting with a suitable substrate, which triggers several
non-enzymatic reactions, similar to a “domino effect.” Moreover, the complexity of
the peroxidase reaction mechanism is evident, which explains why even the
degradation of simple aromatic monomers (e.g., VA), except pathway I in Fig. 11.2
(Dunford 2010), is not fully understood, and that further fundamental research is
still needed. Indeed, in a recent review, Busse et al. (2013) found that the mech-
anism is far more complex and strongly depends on (i) reaction conditions, (ii) POX
type, (iii) the presence of inhibitors, (iv) the used substrate.
Fig. 11.4 Graphical illustration of the adlerol (β-O-4 model compound, structure I) biocatalysis
based on LiP. Several degradation variants (partly in a summarized manner) are shown derived
from proposed reaction schemes by Tien and Kirk (1984), Kirk and Farrell (1987), Lundell et al.
(1993b) and Schoemaker et al. (1994b). I adlerol, II adlerol-derived radical cation, III veratryl
alcohol radical (VA), IV veratraldehyde (VAld) with 60–70 % major product, V Cβ-cation, VI
glycoaldehyde, VII guaiacol, VIII Cβ-centered radical, IX hemiacetal intermediate (unstable),
X keton, a second major product (15 %). Reprinted from Busse et al. (2013) with slight
modifications
282 N. Busse and P. Czermak
In Fig. 11.4, the basic adlerol degradation steps, induced by (fungal) heme per-
oxidases, are schematically explained in conjunction to the previously described,
(non)enzymatic reactions. The catalytic cycle is initiated by H2O2 to form EI as usual.
EI further reacts with adlerol (structure I) to EII, resulting in the production of an
adlerol-derived cation radical intermediate (structure II) (Lundell et al. 1993b). Once
the radical cation is generated, it will be rapidly fragmented, non-enzymatically
(Hatakka et al. 1991). For this, various theories have been under consideration, as
summarized in Fig. 11.4. The reason for this is most likely due to both (i) S•+
chemistry (as already mentioned above), and (ii) pH-dependent product distribution
(Schoemaker 1990). Despite the non-enzymatic reaction, veratraldehyde (VAld,
structure IV) is the main fission product in the end. This same procedure is repeated
when another adlerol molecule is oxidized by EII (refer to pathway 3, Fig. 11.4).
In case adlerol will not serve as the reducing substrate for some reason, com-
ponent II (S•+), III (VA•), and VII (guaiacol) could act as substrates for driving the
cycle back to E0. The latter causes polymerization of its radical products (Chance
and Maehly 1955), which are phenoxy radicals (Lundell et al. 1993b). A reduction
of EII–E0 by component III also results in veratraldehyde (VAld) formation
(Schoemaker 1990; Schoemaker et al. 1994b). However, this is only an occurance
under low-oxygen conditions (Schoemaker et al. 1994b), since the existence of III
is supposed to be short-lived by reacting with O2 immediately. It is important here
to remember that the mediator function of radical products of component III can
cause co-oxidation (thus degradation) of adlerol as well. It is believed that efficient
mediation can only be realized if the degrading substrate will bind close to the
active site of the enzyme (Schoemaker et al. 1994b).
Under substrate (adlerol) limited conditions, Tien and Kirk (1984) demonstrated
a non-stereospecifity of their used LiP, due to complete cleavage of the β-O-4 lignin
model compound. This is an important feature, because lignin has a highly irregular
structure (Schoemaker 1990). Conversely, the involvement of nonenzymatic radical
(chain) reactions is obvious in assisting enzymatic degradation.
Up until now, the normal POX cycle has been introduced (pathway 1–3 in
Figs. 11.2, 11.4). In Fig. 11.5, the supposed VP reaction mechanism is summa-
rized. EI and EII in Fig. 11.5 are expected to be tryptophan radical (Trp•) containing
proteins. Furthermore, this tryptophan radical is suggested to be located on the
enzyme surface (≈10–11 Å to the heme center (Pogni et al. 2005; Ruiz-Dueñas
et al. 2009a)) for substrate oxidation purposes. This means there is no direct contact
between the heme (Fe3+) co-factor and the reducing aromatic substrates (e.g., VA
(non-phenolic monomer) or complex lignin-derived polymers (macromolecules))
avoiding steric hindrance. Interactions between substrate and heme group are
11 Role and Application of Versatile Peroxidase (VP) for Utilizing… 283
Fig. 11.5 Supposed H2O2-dependent reaction mechanism of the crude VP from B. adusta for
adlerol degradation. E0–EIII are enzyme intermediates. The symbols S and S•+ stand for an
appropriate substrate (e.g., adlerol) and the corresponding radical cation. O•−
2 denotes superoxide
radical anions. Pathway 1–3 depicts the usual POX reaction cycle with reversible steps, as
described by equation (a) and (b) (Ruiz-Dueñas et al. 2009b; Busse et al. 2013). The cycle will be
initiated by H2O2 (pathway 1). Pathway 4, 5, and 7 show important side reactions with H2O2. The
latter is in competition with pathway 6. Enzyme deactivation (Ei: inactivated enzyme) is found to
occur via EIII, as sketched in pathway 7. Pathway 8 illustrates a spontaneous unimolecular decay of
EIII. Reprinted from Busse et al. (2013)
enabled via long-range electron transfer (LRET) (Ruiz-Dueñas et al. 2009a). The
involvement of LRET in lignin degradation processes through ligninolytic enzymes
is generally accepted (Busse et al. 2013).
Pathway 1 is the only reaction which has been understood in the past (Dunford
2010), and therefore particular evidence for pathway 3 is still pending (Busse et al.
2013). Thus, for the adlerol bioconversion by a VP type from B. adusta, pathway 3
in Fig. 11.5 was assumed based on transient state kinetics made through VA oxi-
dation studies using a recombinant VP from E. coli (Ruiz-Dueñas et al. 2009b).
Moreover, reversible binary enzyme–substrate (E–S) complex (precursor complex
(Dunford 2010)) formation that occur before irreversible product release are cer-
tainly possible during the catalytic cycle of VP (concerns pathway 2 and 3,
Fig. 11.5) (Ruiz-Dueñas et al. 2009b). Both EI and EII exhibit reactions described
by Eqs. (11.12, 11.13) (Ruiz-Dueñas and Martínez 2009; Ruiz-Dueñas et al.
2009b).
284 N. Busse and P. Czermak
$ E S ! E ½Fe
Km2 k2
EI ½Fe4 þ ¼ OTrp þ S I II 3þ
Trp þ S þ ð11:12Þ
Trp þ S $ E S ! E ½Fe
Km3 k3
EII ½Fe3 þ II 0 3þ
þ S þ ð11:13Þ
vmax, S1, S2, KmS1 , and KmS2 stand for maximum reaction velocity, co-substrate H2O2,
substrate S (e.g., adlerol), dissociation constant for S1 and dissociation constant
for S2.
Maintaining constant S2 concentration, while varying S1, Eq. (11.14) can be
simplified to Eq. (11.15).
max ½S1
vapp
r¼ ð11:15Þ
Kmapp þ ½S1
vmax ½S2
max ¼
vapp ð11:16Þ
KmS2 þ ½S2
KmS1 ½S2
Kmapp ¼ ð11:17Þ
KmS2 þ ½S2
vapp app
max and Km are the apparent kinetic parameters of maximum reaction velocity and
dissociation constant, respectively. Equations (11.14)–(11.17) are valid in the
absence of inhibitions/deactivation phenomena and product(s) (Bisswanger 2008;
Cornish-Bowden 2012). Due to the well-known H2O2-sensitivity of POX,
inactivation/deactivation reactions are substantial side effects of crucial importance.
Figure 11.5 concerns the major reaction pathways that have a co-substrate H2O2 in
addition to substrate S. Both substrates compete for reactions with EI and EII (refer
to Fig. 11.5 pathway 2, 4, and 3, 5, respectively). When a certain S/H2O2/E ratio is
unbalanced, POX deviate from their normal catalytic cycle, and instead promote
reactions via compound III (EIII), diminishing enzyme activity (pathway 5, 7)
(Arnao et al. 1990a, b; Wariishi and Gold 1990; Busse et al. 2013). EIII is another
third enzyme intermediate and exhibits non, or sometimes barley any catalytic
activity (Schoemaker 1990). Subsequently, EIII can react further with H2O2,
resulting in an enzyme deactivation Ei (pathway 7 (Wariishi and Gold 1990;
Goodwin et al. 1994; Busse et al. 2013)), also known as suicide inactivation
(Valderrama et al. 2002). Research investigations have shown that VP deactivation
11 Role and Application of Versatile Peroxidase (VP) for Utilizing… 285
!E
kdðobsÞ
E i ð11:18Þ
kiapp ½S1
kdðobsÞ ¼ ð11:19Þ
Kiapp ½S1
kiapp and Kiapp denote the apparent maximal rate of enzyme inactivation and the
apparent concentration of inhibitor for reaching half-maximal rate of inactivation,
respectively.
In the presence of substrate S, another competition with H2O2 occurs, since EIII
may also react with S by returning back to E0 (pathway 6, Fig. 11.5) under radical
cation intermediate and H2O2 formation (assumed based on Yokota and Yamazaki
(1965) and Acosta et al. (1988)). Alternatively, radical cation products S•+ are able to
revert EIII–E0. However, it does not necessarily mean that all potential substrates are
suitable for such EIII conversion. On the contrary, for a LiP H2 isozyme, for instance,
neither phenolics nor their corresponding phenoxyl radical products nor VA were
capable of returning EIII back to the native state E0, but VA•+ worked fine (Chung
and Aust 1995). The validity of VP and adlerol and its radical cation products still
needs to be clarified. Whatever the case, once an appropriate substrate like adlerol is
added, H2O2-dependent enzyme deactivation is competitively inhibited, and
Eq. (11.19) yields Eq. (11.20). This implies that the substrate does in fact possess
protective properties (Busse et al. 2013). For illustration refer to Fig. 11.6.
kapp ½S1
kdðobsÞ ¼ i ð11:20Þ
Kiapp 1 þ ½SS22 þ ½S1
Km
Depending on the environmental circumstances, e.g., the used buffer system, the
right term in Eqs. (11.19) and (11.20) must be extended by summation of k0i
(Copeland 2002). Thus, k0i denotes the enzyme inactivation rate caused by other
factors apart from H2O2. Furthermore, reaction pathway 5 and 7 (Fig. 11.5), and thus
enzyme deactivation, are pH-dependent and will be favored by an acidic environ-
ment (Cai and Tien 1992; Busse et al. 2013). VPs, and all other POX, commonly
have their reaction optimum around pH 3–4 (Wong 2009; Hofrichter et al. 2010;
Liers et al. 2010; Busse et al. 2013) and are consequently quite susceptible to H2O2
(catalytic amounts of > 30–50 µM (Goodwin et al. 1994; Busse et al. 2013)). This is
a major drawback of POX compared to Lac, particularly when attempting to reach
industrial applications, as will be discussed next. Spontaneous decay of EI (is
unstable (Dunford 2010)) to EII contributes to additional loss in activity (reaction
path is omitted in Fig. 11.5). The unimolecular reaction described by pathway 8
286 N. Busse and P. Czermak
Fig. 11.6 H2O2-dependent deactivation kinetics in the absence of the substrate adlerol
(experimental data (squares), computer simulation (red solid line)) and in the presence of
4 mM adlerol (experimental data (triangles), computer simulation (black solid line)). Reaction
conditions: T: 30 ± 1 °C, pH: 4.0 (100 mM sodium tartrate buffer), n: 375 rpm, enzyme
concentration: *0.06 mg mL−1, V: 10 mL or 50 mL. Each examination was carried out in
duplicates (parallel) with a maximal standard deviation of 10 % from the mean. The inset plot
illustrates the corresponding half-life (t1/2) course as a function of initial H2O2 concentration.
Reprinted from Busse et al. (2013)
(Fig. 11.5), EIII decay to E0 along with O•−2 release, (Nakajima and Yamazaki 1987;
Arnao et al. 1990b) would imply enzyme regeneration or loss of activity.
Finally, it is obvious that balancing the S/E/H2O2 ratio is an important task for
minimizing enzyme deactivations, and thus avoiding excessive H2O2 concentra-
tions. It is in fact impossible to fully eliminate enzyme deactivation, but it can be
delayed. Maintaining enzyme consumption as low as possible is therefore a major
tasks, since POX are currently still costly.
Table 11.2 Hosts for homologous and heterologous VP production (modified based on de Weert
and Lokman (2010))
Organism Host strain Yield References
Homologous
P. ostreatus P. ostreatus TM2-18 21 mg L-1 Tsukihara et al.
7300 U L-1 (2006)
Heterologous
B. adusta E. coli strain BL21 *12 mg L−1 with max. Mohorčič et al.
(DE3)pLysS 0.25 mU mg−1a (2009)
P. eryngii Aspergillus nidulans 466 U L−1 Eibes et al.
(2009)
a
Specific activity was determined based on VA degradation
288 N. Busse and P. Czermak
Liquid lignin product streams (technical lignins) generated via chemical pretreat-
ment of lignocellulosic biomass, for instance, is an alternative substrate for further
processing and finishing by POX. The initial chemical oxidation should preferably
lead to lower molecular weight and more biodegradable intermediates (Mantzavinos
and Psillakis 2004). However, it has to be taken into account that such a pretreatment
can also generate inhibitory products. Moreover, the pH may have to be adjusted.
This might be difficult to achieve. Depending on the isolation method, the resulting
technical lignin may not be soluble at an acidic pH. In general, most technical lignins
are soluble in organic solvents (e.g., dimethyl sulfoxide (DMSO), dioxane,
dimethylformamide, and pyridine) (Saake and Lehnen 2012). VP remains active in
organic solvents like DMSO, at least in the presence of an inorganic substrate
(Rodakiewicz-Nowak et al. 2006). Nonetheless, with the goal of reaching industrial
application, further addition of organic solvents may be uneconomic and in fact
problematic, relating to the recovery of the organic solvent as well as resulting
ecological aspects.
11 Role and Application of Versatile Peroxidase (VP) for Utilizing… 289
For the provision of optimal operating conditions (e.g., T, buffer system, S/H2O2/E
ratio, presence of additives, operation mode), the choice of an appropriate reactor
design is an important issue of bioprocess engineering in order to guarantee the
economic implementation of VP and POX for delignification purposes.
For a rational reactor design, several aspects must be taken into account. Due to
complex and rather uncontrollable radical reactions as seen above, such aspects
include first and foremost an efficient maximization of catalytic turnover numbers
(kcat = vmax/[ET]) in order to avoid an excess of disruptive factors like increased H2O2
and phenol concentrations. Phenols cause polymerization and formation of phenoxyl
radicals affecting the catalytic cycle. A continuous removal of such compounds is
therefore highly recommended. Moreover, reuse of the enzyme is also beneficial.
Combining the enzymatic conversion process by connecting an external cross-
flow ultrafiltration (UF) unit to an enzyme membrane reactor system, as sketched in
Fig. 11.7, represents a promising technological approach for delignification
(Flaschel et al. 1983; Lema et al. 2010; Huang and Ramaswamy 2013). Such
configurations provide several important advantages:
(i) Continuous operation mode.
(ii) Complete biocatalyst retention and recycling.
(iii) Application of free/native enzyme is possible (homogeneous catalysis),
avoiding time-consuming enzyme immobilization accompanied by mass
transfer limitations (Busse et al. 2016; López et al. 2007).
(iv) Replacement of fresh active enzyme depending on deactivation in order to
maintain specified substrate conversion levels (Busse et al. 2016).
(v) Physical biocatalyst immobilization.
(vi) Rapid fractionation of biocatalyst and unreacted high molecular weight
substrates from desired product.
(vii) Continuous removal of undesired by-products and enzyme inhibitors, shift-
ing reaction toward product yield maximization (Busse et al. 2016; López
et al. 2007).
Membrane filtration processes like tubular UF membranes are already well estab-
lished for lignin recovery and separation in biorefineries (Jönsson 2013) and in the
pulp and paper industry (Ebrahimi et al. 2009). This is due to the following eco-
nomic aspects (Huang and Ramaswamy 2013; Jönsson 2013):
(i) Low chemical consumption.
(ii) Fractionation is controllable by pore size.
(iii) Feed volume reduction for further processing (Jönsson and Wimmerstedt 1985).
(iv) Any T and pH adjustments are superfluous.
290 N. Busse and P. Czermak
Despite the fact that tubular UF membranes exhibit low packing densities and high
energy demands (Leiknes 2009), they have “the largest foot-print” (Jönsson 2013).
Moreover, these membranes require low pretreatment (Jönsson 2013), tolerate feed
streams of high fouling potential and of high pollution (e.g., solids) and are easy to
clean (Melin and Rautenbach 2007; Leiknes 2009). Nevertheless, proteins are
serious membrane foulants (Blatt et al. 1970; Porter 1972; Field 1996), besides
lignin-containing feed streams (Sridhar and Bhattacharya 1991; Pfromm and
Watkins 1997; Ebrahimi et al. 2016; Busse et al. 2016). Consequently, constant
enzyme addition, and/or an increased amount of lignin-derived fragments, result in
secondary layer formation on the membrane surface (membrane fouling) limiting
filtration performance, e.g., permeability or throughput, solute retention (Busse
et al. 2016), and membrane service life. Fouling is a common issue in membrane
filtration processes and is mainly affected by (i) membrane properties (e.g., selected
pore size, material, configuration), (ii) hydrodynamic conditions, and (iii) feed
composition (Mulder 1996; Cheryan 1998; Melin and Rautenbach 2007). Filtration
performance is a crucial parameter, since it determines the efficiency of enzymatic
conversion, protein load, and reactor size, along with hydraulic retention time
(refer to Fig. 11.7) (Busse et al. 2016). Thus, the maintenance of constant high
volumetric flow rates (or fluxes) is a major task throughout filtration processes.
Using ceramic membranes is advantageous in comparison to polymer mem-
branes. Major aspects are:
(i) High thermal and chemical stability which is beneficial for cleaning.
(ii) High mechanical stability, particularly when filtering abrasive media.
(iii) Less susceptible to fouling and low adhesion potential for molecular organic
substances (Ebrahimi et al. 2015; Fan et al. 2015).
(iv) High membrane service life/lifetime (6 years for ceramic membranes and 1.5
for polymeric) (Arkell et al. 2014).
Conversely, polymeric membranes are less expensive, easy to configure and scal-
able (Van der Bruggen 2009).
Several applications have already been reported using POX in membrane biore-
actor systems. However, these have just been for treating xenobiotics and pollutions
(e.g., azo dye Orange II by MnP, nonylphenol by VP), which use mostly polymeric
UF membranes (López et al. 2007; Méndez-Hernández et al. 2015). The choice for
the optimal material ultimately depends on individual needs (Arkell et al. 2014).
Key prerequisite for implementing an effective process control, and thus ensuring
optimal operation condition, is an efficient and reliable online monitoring system
(except T and pH, since they are standard). Although H2O2 concentration is such a
crucial parameter, the availability of commercial H2O2 online sensors, with suffi-
cient sensitivity over a wide pH and T range, is limited. Contrarily, there are several
fluorescence probes, such as some based on oxidation of N-
acetyl-3,7-dihydroxy-phenoxazine (e.g., Amplex Red®) by HRP, for measuring
H2O2 within a range of 10 nM to 100 µM (Gomes et al. 2005). Disadvantages of
using such probes are: (i) sample preparation is required, (ii) HRP inactivation at c
(H2O2) > 100 µM (Towne et al. 2004), thus first requiring the finding of an
appropriate pre-dilution, (iii) incubation times of 15–30 min before receiving first
results, and (iv) pH is also limited, ranging from 7.5 to 8.5. Similar conditions are
true concerning any HRP/POX-related biosensor for H2O2 measurements. In the
current state, amperometric sensors equipped with platinum electrodes appear rather
suitable, since they are not susceptible to H2O2. Sample preparations are also
superfluous, among other things. Particularly, the membrane protected PEROX
H2O2 online sensors from ProMinent Dosiertechnik in Heidelberg, Germany, meet
useful requirements, including aspects such as a wide pH-/T- and c(H2O2)-range
with a comparatively low response time. First successful applications of this have
already been reported for chloroperoxidase catalyzed oxidations (Seelbach et al.
1997; van Deurzen et al. 1997).
Another important parameter is the actual enzyme activity, which could be
detected by simply monitoring the production of certain fission products and/or
292 N. Busse and P. Czermak
substrate consumption. When complex substrates such as lignin are involved and
product formation is hardly manageable, as already illustrated above, lignin degra-
dation is usually monitored by UV absorption measurements at λ = 280 nm. Highly
concentrated lignin solutions appear dark brown. The color intensity is an indirect
indicator of lignin amount (Kinstre 1993). Conversely, the release of lignin-derived
high molecular weight fragments from LC biomass would result in an intensification
of color in the culture medium. A precise assignment of these fragments is not
possible, and further analytical methods will be required which are limited in
applicability, partly very costly, and not yet sufficiently standardized. An extensive
overview will be given by Lin and Dence (1992). Moreover, a fundamental expertise
is mandatory for process optimization and data interpretation. At present, there is no
simple, standalone analytical method. Based on these facts, alternative (in)direct
measurement techniques must be found. Such techniques should be (i) cheap,
(ii) easy to implement, online, (iii) robust, (iv) quick and precise and (v) low main-
tenance. Measuring dissolved oxygen (DO), especially via optical sensors, is cur-
rently the most promising parameter for process monitoring due to pseudo-catalase
activity of POX at H2O2 excess (or low c(S)/c(H2O2) ratios) resulting in O2 formation
(Acosta et al. 1988; Wariishi and Gold 1989; Vlasits et al. 2010). Excessive c(H2O2)
causes considerable enzyme deactivation (Busse et al. 2013). On the contrary, O2
consumption is an indicator for high c(S)/c(H2O2) ratios (López et al. 2007). Thus, O2
consumption is also indicative for substrate degradation processes as seen in
Figs. 11.3 and 11.4, whereas O2 can also be generated simultaneously concerning
HO•2/O•− 2 release (Figs. 11.3 and 11.5). Recent studies by López et al. (2007), treating
xenobiotic compounds such as Organe II by MnP, have shown that DO could be
valuable for implementing process control strategies (e.g., feedback control).
However, based on the variety of radical reactions, O2 formation/consumption is not
trivial (Palmer et al. 1987). Further research is therefore needed for a better (process)
understanding, particularly in the field of delignification. Moreover, the presence of
substrates (e.g., VA and adlerol) capable of bond cleavage seems to be a precondition
for significant O2 consumption (Palmer et al. 1987).
underlying reaction conditions and not by the immobilization method. Despite sev-
eral studies, the proof-of-concept is still outstanding. Additionally, in the case of
woody biomass, good enzyme penetration must be further guaranteed, affected by
porosity and the pore size, of the substrate to be treated.
Another promising/future based strategy may be enzymes from insects and
insect-associated microorganisms. Insects are the most diverse taxonomic animal
class, colonizing almost every ecological niche. To succeed in these sometimes
extreme niches, insects have established diverse biological and chemical systems,
e.g., the production of defense molecules (Gross et al. 2008; Schlipalius et al. 2012),
protein stabilization (Bale 2002), or lytic enzymes (Landureau and Jolles 1970; Mika
et al. 2013). Insects also house symbiotic microorganisms as digestive helpers,
fodder, or both (Scharf et al. 2011). This is possible due to the special metabolic
systems, or exogenous enzymes, of the associated organisms. For instance, insects
like bark beetles, ambrosia beetles and termites, are able to feed on wood, making
them xylophagous, and are well known as wood pests. The function of their
digestive systems is not yet completely clarified. The enzymatic apparatus for the
oxidation of lignin and the hydrolysis of cellulose is only marginally understood.
One reason might be the diverse sources of the key enzymes, which may be pro-
duced either by gut-inhabiting microorganisms (Morrison et al. 2009), by symbiotic
fungi cultivated by the insects (Currie 2001), or by the insects themselves as
endogenous enzymes (Pauchet et al. 2010). Nevertheless, a lot of work has been
done in attempts to understand the lignocellulolytic system of the insects and the
degree of participation by symbionts (Geib et al. 2008; Scharf et al. 2011).
Currently, POX are still costly for industrial use. Nonetheless, there is no doubt
that once limitations are resolved, particularly H2O2 instability and efficient pro-
duction, the potential for large-scale application of VPs raises significantly (Casella
et al. 2010). However, many questions with respect to the enzymatic delignification
mechanism still remain in place, i.e., substrate preparation/pretreatment, E0
recovery, control of inhibitory reactions/factors besides H2O2, involvement of
accessory factors and enzymes, etc. Moreover, for better process control, devel-
opments focusing on simple and fast online monitoring systems will also be nec-
essary. For these purposes, further fundamental research is needed.
Acknowledgments The researchers would like to thank the Hessen State Ministry of Higher
Education, Research and the Arts for the financial support within the Hessen initiative for scientific
and economic excellence (LOEWE).
References
Andrawis A, Johnson KA, Tien M (1988) Studies on compound I formation of the lignin
peroxidase from Phanerochaete chrysosporium. J Biol Chem 263(3):1195–1198
Arkell A, Olsson J, Wallberg O (2014) Process performance in lignin separation from softwood
black liquor by membrane filtration. Chem Eng Res Des 92(9):1792–1800. doi:10.1016/j.
cherd.2013.12.018
Arnao MB, Acosta M, del Rio JA (1038) García-Cánovas F (1990a) Inactivation of peroxidase by
hydrogen peroxide and its protection by a reductant agent. Biochim Biophys Acta
(BBA) Protein Struct Mol Enzymol 1:85–89. doi:10.1016/0167-4838(90)90014-7
Arnao MB, Acosta M, del Río JA, Varón R (1041) García-Cánovas F (1990b) A kinetic study on
the suicide inactivation of peroxidase by hydrogen peroxide. Biochim Biophys Acta Protein
(BBA) Struct Mol Enzymol 1:43–47. doi:10.1016/0167-4838(90)90120-5
Ayala M (2010) Redox potential of peroxidases. In: Torres E, Ayala M (eds) Biocatalysis based on
heme peroxidases. Springer, Berlin, pp 61–77. doi:10.1007/978-3-642-12627-7_4. ISBN
978-3-642-12626-0
Ayala M, Pickard MA, Vazquez-Duhalt R (2008) Fungal enzymes for environmental purposes, a
molecular biology challenge. J Mol Microbiol Biotechnol 15(2–3):172–180. doi:10.1159/
000121328
Bale JS (2002) Insects and low temperatures: from molecular biology to distributions and
abundance. Philos Trans B 357(1423):849–862. doi:10.1098/rstb.2002.1074
Bielski BH, Cabelli DE, Ravindra LA, Ross AB (1985) Reactivity of HO2/O2-radicals in aqueous
solution. J Phys Chem Ref Data 14(4):1041–1100
Bisswanger H (2008) Enzyme kinetics: section 2. In: Enzyme kinetics. Wiley-VCH Verlag GmbH
& Co. KGaA, pp 124–193. ISBN 9783527319572
Blanchette RA (1991) Delignification by wood-decay fungi. Annu Rev Phytopathol 29(1):
381–403. doi:10.1146/annurev.py.29.090191.002121
Blatt WF, Dravid A, Michaels AS, Nelsen L (1970) Solute polarization and cake formation in
membrane ultrafiltration: causes, consequences, and control techniques. In: Flinn JE
(ed) Membrane science and technology. Springer, USA, pp 47–97. doi:10.1007/978-1-4684-
1851-4_4. ISBN 978-1-4684-1853-8 (print), 978-1-4684-1851-4 (online)
Böckle B, Martínez MJ, Guillén F, Martínez ÁT (1999) Mechanism of peroxidase inactivation in
liquid cultures of the ligninolytic fungus Pleurotus pulmonarius. Appl Environ Microbiol 65
(3):923–928
Bourbonnais R, Paice MG (1990) Oxidation of non-phenolic substrates: an expanded role for laccase
in lignin biodegradation. FEBS Lett 267(1):99–102. doi:10.1016/0014-5793(90)80298-W
Busse N, Wagner D, Kraume M, Czermak P (2013) Reaction kinetics of versatile peroxidase for
the degradation of lignin compounds. Am J Biochem Biotechnol 9(4):365–394. doi:10.3844/
ajbbsp.2013.365.394
Busse N, Fuchs F, Kraume M, Czermak P (2016) Treatment of enzyme initiated delignification
reaction mixtures with ceramic ultrafiltration membranes: Experimental investigations and
modeling approach. Sep Sci Technol 51(9):1546–1565. doi:10.1080/01496395.2016.1167739
Cai D, Tien M (1992) Kinetic studies on the formation and decomposition of compounds II and
III. Reactions of lignin peroxidase with H2O2. J Biol Chem 267(16):11149–11155
Camarero S, Sarkar S, Ruiz-Dueñas FJ, Martınez ́ MJ, Martínez ÁT (1999) Description of a
versatile peroxidase involved in the natural degradation of lignin that has both manganese
peroxidase and lignin peroxidase substrate interaction sites. J Biol Chem 274(15):10324–
10330. doi:10.1074/jbc.274.15.10324
Cañas AI, Camarero S (2010) Laccases and their natural mediators: biotechnological tools for
sustainable eco-friendly processes. Biotechnol Adv 28(6):694–705. doi:10.1016/j.biotechadv.
2010.05.002
Carabajal M, Kellner H, Levin L, Jehmlich N, Hofrichter M, Ullrich R (2013) The secretome of
Trametes versicolor grown on tomato juice medium and purification of the secreted
oxidoreductases including a versatile peroxidase. J Biotechnol 168(1):15–23. doi:10.1016/j.
jbiotec.2013.08.007
11 Role and Application of Versatile Peroxidase (VP) for Utilizing… 295
Casella L, Monzani E, Nicolis S (2010) Potential applications of peroxidases in the fine chemical
industries. In: Torres E, Ayala M (eds) Biocatalysis based on heme peroxidases. Springer,
Berlin, pp 111–153. doi:10.1007/978-3-642-12627-7_6. ISBN 978-3-642-12626-0
Chance B, Maehly AC (1955) Assay of catalases and peroxidases. In: Methods in enzymology, vol
2. Academic Press, London, pp 764–775. doi: 10.1016/S0076-6879(55)02300-8
Cheryan M (1998) Ultrafiltration and microfiltration handbook. CRC Press, London. ISBN
978-1-56676-598-5
Chung N, Aust SD (1995) Inactivation of lignin peroxidase by hydrogen peroxide during the
oxidation of phenols. Arch Biochem Biophys 316(2):851–855. doi:10.1006/abbi.1995.1114
Conesa A, van den Hondel CAMJJ, Punt PJ (2000) Studies on the production of fungal
peroxidases in Aspergillus niger. Appl Environ Microbiol 66(7):3016–3023. doi:10.1128/aem.
66.7.3016-3023.2000
Copeland RA (2002) Enzymes: a practical introduction to structure, mechanism, and data analysis.
Wiley-VCH, London. ISBN 0-471-35929-7, 0-471-22063-9
Cornish-Bowden A (2012) Fundamentals of enzyme kinetics. Wiley-Blackwell, London. ISBN
9783527330744
Currie CR (2001) A community of ants, fungi, and bacteria: a multilateral approach to studying
symbiosis. Annu Rev Microbiol 55(1):357–380. doi:10.1146/annurev.micro.55.1.357
Dashtban M, Schraft H, Syed TA, Qin W (2010) Fungal biodegradation and enzymatic
modification of lignin. Int J Biochem Mol Biol 1(1):36–50
de Montellano PRO (2010) Catalytic mechanisms of heme peroxidases. In: Torres E, Ayala M
(eds) Biocatalysis based on heme peroxidases. Springer, Berlin, pp 79–107. doi:10.1007/978-
3-642-12627-7_5. ISBN 978-3-642-12626-0
de Weert S, Lokman BC (2010) Heterologous expression of peroxidases. In: Torres E, Ayala M
(eds) Biocatalysis based on heme peroxidases. Springer, Berlin, pp 315–333. doi:10.1007/978-
3-642-12627-7_12. ISBN 978-3-642-12626-0
Donohoe BS, Decker SR, Tucker MP, Himmel ME, Vinzant TB (2008) Visualizing lignin
coalescence and migration through maize cell walls following thermochemical pretreatment.
Biotechnol Bioeng 101(5):913–925. doi:10.1002/bit.21959
Dunford HB (1991) Horsradish peroxidase: structure and kinetic properties. In: Everse J
(ed) Peroxidases in chemistry and biology, vol 2. CRC Press, London, pp 1–24. ISBN
0849369649
Dunford HB (2010) Peroxidases and catalases: biochemistry, biophysics, biotechnology and
physiology. Wiley, London. ISBN 9780470224762
Ebrahimi M, Kerker S, Wienold A, Neul H, Hilpert M, Mund P, Czermak P (2009) Processing and
characterization of ceramic membranes for the efficient removal of COD and residual lignin
from bleaching effluents. International conference and exhibition for filtration and separation
technology. Filtech Wiesbaden, Germany
Ebrahimi M, Busse N, Kerker S, Schmitz O, Hilpert M, Czermak P (2016) Treatment of
the bleaching effluent from sulfite pulp production by ceramic membrane filtration. Membranes
6(1). doi:10.3390/membranes6010007
Ebrahimi M, Kerker S, Daume S, Geile M, Ehlen F, Unger I, Schütz S, Czermak P (2015)
Innovative ceramic hollow fiber membranes for recycling/reuse of oilfield produced water.
Desalination Water Treat 55(13):3554–3567. doi:10.1080/19443994.2014.947780
Eibes GM, Lú-Chau TA, Ruiz-Dueñas FJ, Feijoo G, Martínez MJ, Martínez AT, Lema JM (2009)
Effect of culture temperature on the heterologous expression of Pleurotus eryngii versatile
peroxidase in Aspergillus hosts. Bioprocess Biosyst Eng 32(1):129–134. doi:10.1007/s00449-
008-0231-7
Eriksson K-EL, Blanchette RA, Ander P (1990) Microbial and enzymatic degradation of wood and
wood components. Springer, Berlin. ISBN 354051600X
Fan R, Ebrahimi M, Quitmann H, Czermak P (2015) Lactic acid production in a membrane
bioreactor system with thermophilic Bacillus coagulans: fouling analysis of the used ceramic
membranes. Sep Sci Technol 50(14):2177–2189. doi:10.1080/01496395.2015.1031401
296 N. Busse and P. Czermak
Field RW (1996) Mass transport and the design of membrane systems. In: Scott K, Hughes R
(eds) Industrial membrane separation technology. Springer, Netherlands, pp 67–113. doi:10.
1007/978-94-011-0627-6_4
Flaschel E, Wandrey C, Kula M-R (1983) Ultrafiltration for the separation of biocatalysts.
Downstream processing. Springer, Berlin, pp 73–142. doi:10.1007/978-3-662-39694-0_2
Geib SM, Filley TR, Hatcher PG, Hoover K, Carlson JE, Jimenez-Gasco MdM, Nakagawa-Izumi
A, Sleighter RL, Tien M (2008) Lignin degradation in wood-feeding insects. Proc Natl Acad
Sci 105(35):12932–12937. doi:10.1073/pnas.0805257105
Glasser WG, Davé V, Frazier CE (1993) Molecular weight distribution of (semi-) commercial
lignin derivatives. J Wood Chem Technol 13(4):545–559. doi:10.1080/02773819308020533
Gomes A, Fernandes E, Lima JLFC (2005) Fluorescence probes used for detection of reactive
oxygen species. J Biochem Biophys Methods 65(2–3):45–80. doi:10.1016/j.jbbm.2005.10.003
Goodwin DC, Barr DP, Aust SD, Grover TA (1994) The role of oxalate in lignin
peroxidase-catalyzed reduction: protection from compound III accumulation. Arch Biochem
Biophys 315(2):267–272. doi:10.1006/abbi.1994.1499
Gross J, Schumacher K, Schmidtberg H, Vilcinskas A (2008) Protected by fumigants: beetle
perfumes in antimicrobial defense. J Chem Ecol 34(2):179–188. doi:10.1007/s10886-007-
9416-9
Halliwell B, Gutteridge JMC (1985) The importance of free radicals and catalytic metal ions in
human diseases. Mol Aspects Med 8(2):89–193. doi:10.1016/0098-2997(85)90001-9
Hammel KE, Jensen KA, Mozuch MD, Landucci LL, Tien M, Pease EA (1993) Ligninolysis by a
purified lignin peroxidase. J Biol Chem 268(17):12274–12281
Harman LS, Carver DK, Schreiber J, Mason RP (1986) One- and two-electron oxidation of
reduced glutathione by peroxidases. J Biol Chem 261(4):1642–1648
Hatakka A (1994) Lignin-modifying enzymes from selected white-rot fungi: production and role
from in lignin degradation. FEMS Microbiol Rev 13(2–3):125–135. doi:10.1111/j.1574-6976.
1994.tb00039.x
Hatakka A (2005) Biodegradation of lignin. In: Biopolymers online. Wiley-VCH Verlag GmbH &
Co. KGaA. doi:10.1002/3527600035.bpol1005. ISBN 9783527600038
Hatakka A, Hammel K (2011) Fungal biodegradation of lignocelluloses. In: Hofrichter M
(ed) Industrial applications, vol 10. Springer, Berlin, pp 319–340. doi:10.1007/978-3-642-
11458-8_15. ISBN 978-3-642-11457-1
Hatakka A, Lundell T, Tervilä-Wilo ALM, Brunow G (1991) Metabolism of non-phenolic β-O-4
lignin model compounds by the white-rot fungus Phlebia radiata. Appl Microbiol Biotechnol
36(2):270–277. doi:10.1007/BF00164433
Hofrichter M, Ullrich R, Pecyna M, Liers C, Lundell T (2010) New and classic families of secreted
fungal heme peroxidases. Appl Microbiol Biotechnol 87(3):871–897. doi:10.1007/s00253-
010-2633-0
Holladay J, White J, Bozell J, Johnson D (2007) Top value-added chemicals from biomass—
volume II—results of screening for potential candidates from biorefinery lignin Pacific
Northwest National Laboratory. Richland, WA, U.S. Department of Energy, 87
Howard RL, Abotsi E, Jansen van Rensburg EL, Howard S (2003) Lignocellulose biotechnology:
issues of bioconversion and enzyme production. Afr J Biotechnol 2(12):602–619
Huang H-J, Ramaswamy S (2013) Overview of biomass conversion processes and separation and
purification technologies in biorefineries. In: Ramaswamy S, Huang H-J, Ramarao BV
(eds) Separation and purification technologies in biorefineries. Wiley, London, pp 3–36. ISBN
978-0-470-97796-5 (Hardcover)
Johnson TM, Pease EA, Li JKK, Tien M (1992) Production and characterization of recombinant
lignin peroxidase isozyme H2 from Phanerochaete chrysosporium using recombinant
baculovirus. Arch Biochem Biophys 296(2):660–666. doi:10.1016/0003-9861(92)90624-6
Jönsson A-S (2013) Microfiltration, ultrafiltration and diafiltration. In: Ramaswamy S, Huang H-J,
Ramarao BV (eds) Separation and purification technologies in biorefineries. Wiley, London,
pp 205–231. ISBN 978-0-470-97796-5 (Hardcover)
11 Role and Application of Versatile Peroxidase (VP) for Utilizing… 297
Jönsson AS, Wimmerstedt R (1985) The application of membrane technology in the pulp and
paper industry. Desalination 53(1–3):181–196. doi:10.1016/0011-9164(85)85060-8
Kawai S, Asukai M, Ohya N, Okita K, Ito T, Ohashi H (1999) Degradation of a non-phenolic
β-O-4 substructure and of polymeric lignin model compounds by laccase of Coriolus
versicolor in the presence of 1-hydroxybenzotriazole. FEMS Microbiol Lett 170(1):51–57.
doi:10.1111/j.1574-6968.1999.tb13354.x
Kettle AJ, Anderson RF, Hampton MB, Winterbourn CC (2007) Reactions of superoxide with
myeloperoxidase. Biochemistry 46(16):4888–4897. doi:10.1021/bi602587k
Kinstre RB (1993) An overview of strategies for reducing the environmental impact of
bleach-plant effluents. Tappi J 76(5):105–113
Kirk TK, Farrell RL (1987) Enzymatic “combustion”: the microbial degradation of lignin. Annu
Rev Microbiol 41:465–505
Landureau JC, Jolles P (1970) Lytic enzyme produced in vitro by insect cells: lysozymelor
chitinase? Nature 225(5236):968–969. doi:10.1038/225968a0
Lange J-P (2007) Lignocellulose conversion: an introduction to chemistry, process and economics.
Biofuels Bioprod Biorefin 1(1):39–48. doi:10.1002/9783527621118.ch2
Leiknes T (2009) Wastewater treatment by membrane bioreactors. In: Membrane operations.
Wiley-VCH Verlag GmbH & Co. KGaA, pp 363–395. doi: 10.1002/9783527626779.ch16.
ISBN 9783527626779
Lema J, López C, Eibes G, Taboada-Puig R, Moreira MT, Feijoo G (2010) Reactor
engineering. In: Torres E, Ayala M (eds) Biocatalysis based on heme peroxidases. Springer,
Berlin, pp 245–290. doi:10.1007/978-3-642-12627-7_10. ISBN 978-3-642-12626-0
Liers C, Bobeth C, Pecyna M, Ullrich R, Hofrichter M (2010) DyP-like peroxidases of the jelly
fungus Auricularia auricula-judae oxidize nonphenolic lignin model compounds and
high-redox potential dyes. Appl Microbiol Biotechnol 85(6):1869–1879. doi:10.1007/
s00253-009-2173-7
Lin SY, Dence CW (1992) Methods in lignin chemistry. Springer, Berlin. doi:10.1007/978-3-642-
74065-7. ISBN 9783642740671
Longoria A, Tinoco R, Torres E (2010) Enzyme technology of peroxidases: immobilization,
chemical and genetic modification. In: Torres E, Ayala M (eds) Biocatalysis based on heme
peroxidases. Springer, Berlin, pp 209–243. doi:10.1007/978-3-642-12627-7_9. ISBN
978-3-642-12626-0
López C, Moreira MT, Feijoo G, Lema JM (2007) Dynamic modeling of an enzymatic membrane
reactor for the treatment of xenobiotic compounds. Biotechnol Bioeng 97(5):1128–1137.
doi:10.1002/bit.21311
Lundell T, Schoemaker H, Hatakka A, Brunow G (1993a) New mechanism of the Ca–Cb cleavage
in non-phenolic arylglycerol b-aryl ether lignin substructures catalyzed by lignin peroxidase.
Holzforschung Int J Biol Chem Phys Technol Wood 47:219. doi:10.1515/hfsg.1993.47.3.219
Lundell T, Wever R, Floris R, Harvey P, Hatakka A, Brunow G, Schoemaker H (1993b) Lignin
peroxidase L3 from Phlebia radiata. Pre-steady-state and steady-state studies with veratryl
alcohol and a non-phenolic lignin model compound 1-(3,4-dimethoxyphenyl)-2-
(2-methoxyphenoxy)propane-1,3-diol. Eur J Biochem 211(3):391–402
Lundell TK, Mäkelä MR, Hildén K (2010) Lignin-modifying enzymes in filamentous
basidiomycetes—ecological, functional and phylogenetic review. J Basic Microbiol 50(1):
5–20. doi:10.1002/jobm.200900338
Mantzavinos D, Psillakis E (2004) Enhancement of biodegradability of industrial wastewaters by
chemical oxidation pre-treatment. J Chem Technol Biotechnol 79(5):431–454. doi:10.1002/
jctb.1020
Martínez ÁT (2007) High redox potential peroxidases. In: Polaina J, MacCabe A (eds) Industrial
enzymes. Springer, Netherlands, pp 477–488. doi:10.1007/1-4020-5377-0_27. ISBN
9781402053771
Martínez MJ, Ruiz-Dueñas FJ, Guillén F, Martínez ÁT (1996) Purification and catalytic properties
of two manganese peroxidase isoenzymes from Pleurotus eryngii. Eur J Biochem 237(2):
424–432. doi:10.1111/j.1432-1033.1996.0424k.x
298 N. Busse and P. Czermak
Martínez ÁT, Ruiz-Dueñas FJ, Martínez MJ, del Río JC, Gutiérrez A (2009) Enzymatic
delignification of plant cell wall: from nature to mill. Curr Opin Biotechnol 20(3):348–357.
doi:10.1016/j.copbio.2009.05.002
Melin T, Rautenbach R (2007) Membranverfahren Grundlagen der Modul- und
Anlagenauslegung. Springer, Berlin. ISBN 978-3-540-34327-1
Méndez-Hernández JE, Eibes G, Arca-Ramos A, Lú-Chau TA, Feijoo G, Moreira MT, Lema JM
(2015) Continuous removal of nonylphenol by versatile peroxidase in a two-stage membrane
bioreactor. Appl Biochem Biotechnol 175(6):3038–3047. doi:10.1007/s12010-014-1474-1
Mester T, Field JA (1998) Characterization of a novel manganese peroxidase-lignin peroxidase
hybrid isozyme produced by Bjerkandera species strain BOS55 in the absence of manganese.
J Biol Chem 273(25):15412–15417. doi:10.1074/jbc.273.25.15412
Mika N, Zorn H, Rühl M (2013) Insect-derived enzymes: a treasure for industrial biotechnology
and food biotechnology. In: Vilcinskas A (ed) Yellow biotechnology II. Springer, Berlin,
pp 1–17. doi:10.1007/10_2013_204. ISBN 3642399010
Mohorčič M, Benčina M, Friedrich J, Jerala R (2009) Expression of soluble versatile peroxidase of
Bjerkandera adusta in Escherichia coli. Bioresour Technol 100(2):851–858. doi:10.1016/
j.biortech.2008.07.005
Moreira PR, Bouillenne F, Almeida-Vara E, Xavier Malcata F, Frère JM, Duarte JC (2006)
Purification, kinetics and spectral characterisation of a new versatile peroxidase from a
Bjerkandera sp. isolate. Enzyme Microb Technol 38(1–2):28–33. doi:10.1016/j.enzmictec.
2004.12.035
Morrison M, Pope PB, Denman SE, McSweeney CS (2009) Plant biomass degradation by gut
microbiomes: more of the same or something new? Curr Opin Biotechnol 20(3):358–363.
doi:10.1016/j.copbio.2009.05.004
Mulder MHV (1996) Basic principles of membrane technology. Kluwer, The Netherlands. ISBN
0-7923-4247-X (HB)
Nakajima R, Yamazaki I (1987) The mechanism of oxyperoxidase formation from ferryl
peroxidase and hydrogen peroxide. J Biol Chem 262(6):2576–2581
Nicell JA, Bewtra JK, Biswas N, Taylor E (1993) Reactor development for peroxidase catalyzed
polymerization and precipitation of phenols from wastewater. Water Res 27(11):1629–1639.
doi:10.1016/0043-1354(93)90127-4
Palmer JM, Harvey PJ, Schoemaker HE (1987) The role of peroxidases, radical cations and
oxygen in the degradation of lignin [and discussion]. Philos Trans R Soc Lond A 321
(1561):495–505. doi:10.1098/rsta.1987.0027
Pauchet Y, Wilkinson P, Chauhan R, French-Constant RH (2010) Diversity of beetle genes
encoding novel plant cell wall degrading enzymes. PLoS One 5(12):e15635. doi:10.1371/
journal.pone.0015635
Pease EA, Aust SD, Tien M (1991) Heterologous expression of active manganese peroxidase from
Phanerochaete chrysosporium using the baculovirus expression system. Biochem Biophys Res
Commun 179(2):897–903. doi:10.1016/0006-291X(91)91903-P
Pérez-Boada M, Ruiz-Dueñas FJ, Pogni R, Basosi R, Choinowski T, Martínez MJ, Piontek K,
Martínez ÁT (2005) Versatile peroxidase oxidation of high redox potential aromatic
compounds: site-directed mutagenesis, spectroscopic and crystallographic investigation of
three long-range electron transfer pathways. J Mol Biol 354(2):385–402. doi:10.1016/j.jmb.
2005.09.047
Pfromm PH, Watkins EJ (1997) Fouling of electrodialysis membranes by organic macromolecules.
IPST technical paper series number 676 12
Pogni R, Baratto MC, Giansanti S, Teutloff C, Verdin J, Valderrama B, Lendzian F, Lubitz W,
Vazquez-Duhalt R, Basosi R (2005) Tryptophan-based radical in the catalytic mechanism of
versatile peroxidase from Bjerkandera adusta. Biochemistry 44(11):4267–4274. doi:10.1021/
bi047474l
Porter MC (1972) Concentration polarization with membrane ultrafiltration. Product R&D 11
(3):234–248. doi:10.1021/i360043a002
11 Role and Application of Versatile Peroxidase (VP) for Utilizing… 299
Rasmussen CB, Dunford HB, Welinder KG (1995) Rate enhancement of compound I formation of
barley peroxidase by ferulic acid, caffeic acid and coniferyl alcohol. Biochemistry 34
(12):4022–4029. doi:10.1021/bi00012a021
Rodakiewicz-Nowak J, Jarosz-Wilkołazka A, Luterek J (2006) Catalytic activity of versatile
peroxidase from Bjerkandera fumosa in aqueous solutions of water-miscible organic solvents.
Appl Catal A 308:56–61. doi:10.1016/j.apcata.2006.04.009
Rodríguez Couto S, Toca Herrera JL (2006) Industrial and biotechnological applications of
laccases: a review. Biotechnol Adv 24(5):500–513. doi:10.1016/j.biotechadv.2006.04.003
Rühl M, Kilaru S, Navarro-Gonzaléz M, Hoegger P, Kharazipour A, Kües U (2007) Production
of laccase and other enzymes for the wood industry. In: Kües U (ed) Wood production,
wood technology, and biotechnological impacts. Universitätsverlag Göttingen, Göttingen,
pp 469–507. ISBN 0849369649
Rühl M, Fischer C, Kües U (2008) Ligninolytic enzyme activities alternate with mushroom
production during industrial cultivation of Pleurotus ostreatus on wheatstraw-based substrate.
Curr Trends Biotechnol Pharm 2(4):478–492. ISSN 0973-8916
Ruiz-Dueñas FJ, Martínez ÁT (2009) Microbial degradation of lignin: how a bulky recalcitrant
polymer is efficiently recycled in nature and how we can take advantage of this. Microb
Biotechnol 2(2):164–177. doi:10.1111/j.1751-7915.2008.00078.x
Ruiz-Dueñas FJ, Morales M, García E, Miki Y, Martínez MJ, Martínez ÁT (2009a) Substrate
oxidation sites in versatile peroxidase and other basidiomycete peroxidases. J Exp Bot 60
(2):441–452. doi:10.1093/jxb/ern261
Ruiz-Dueñas FJ, Pogni R, Morales M, Giansanti S, Mate MJ, Romero A, Martínez MJ, Basosi R,
Martínez ÁT (2009b) Protein radicals in fungal versatile peroxidase: catalytic tryptophan
radical in both compound I and compound II and studies on W164Y, W164H, and W164S
variants. J Biol Chem 284(12):7986–7994. doi:10.1074/jbc.M808069200
Saake B, Lehnen R (2012) Lignin. In: Ullmann’s encyclopedia of industrial chemistry.
Wiley-VCH Verlag GmbH & Co. KGaA. doi: 10.1002/14356007.a15_305.pub3
Sánchez C (2009) Lignocellulosic residues: biodegradation and bioconversion by fungi.
Biotechnol Adv 27(2):185–194. doi:10.1016/j.biotechadv.2008.11.001
Scharf ME, Karl ZJ, Sethi A, Boucias DG (2011) Multiple levels of synergistic collaboration in
termite lignocellulose digestion. PLoS One 6(7):e21709. doi:10.1371/journal.pone.0021709
Schlipalius DI, Valmas N, Tuck AG, Jagadeesan R, Ma L, Kaur R, Goldinger A, Anderson C,
Kuang J, Zuryn S, Mau YS, Cheng Q, Collins PJ, Nayak MK, Schirra HJ, Hilliard MA,
Ebert PR (2012) A core metabolic enzyme mediates resistance to phosphine gas. Science 338
(6108):807–810. doi:10.1126/science.1224951
Schoemaker HE (1990) On the chemistry of lignin biodegradation. Recl Trav Chim Pays-Bas 109
(4):255–272. doi:10.1002/recl.19901090402
Schoemaker HE, Piontek K (1996) On the interaction of lignin peroxidase with lignin. Pure Appl
Chem 68:2089–2096. doi:10.1351/pac199668112089
Schoemaker HE, Lundell TK, Floris R, Glumoff T, Winterhalter KH, Piontek K (1994a) Do
carbohydrates play a role in the lignin peroxidase cycle? Redox catalysis in the endergonic
region of the driving force. Bioorganic Med Chem 2(6):509–519. doi:10.1016/0968-0896(94)
80021-9
Schoemaker HE, Lundell TK, Hatakka AI, Piontek K (1994b) The oxidation of veratryl alcohol,
dimeric lignin models and lignin by lignin peroxidase: the redox cycle revisited. FEMS
Microbiol Rev 13(2–3):321–331. doi:10.1111/j.1574-6976.1994.tb00052.x
Schüttmann I, Bouws H, Szweda RT, Suckow M, Czermak P, Zorn H (2014) Induction,
characterization, and heterologous expression of a carotenoid degrading versatile peroxidase
from Pleurotus sapidus. J Mol Catal B Enzym 103:79–84. doi:10.1016/j.molcatb.2013.08.007
Seelbach K, van Deurzen MPJ, van Rantwijk F, Sheldon RA, Kragl U (1997) Improvement of the
total turnover number and space-time yield for chloroperoxidase catalyzed oxidation.
Biotechnol Bioeng 55(2):283–288. doi:10.1002/(sici)1097-0290(19970720)55:2<283:aid-
bit6>3.0.co;2-e
300 N. Busse and P. Czermak
Singh D, Chen S (2008) The white-rot fungus Phanerochaete chrysosporium: conditions for the
production of lignin-degrading enzymes. Appl Microbiol Biotechnol 81(3):399–417. doi:10.
1007/s00253-008-1706-9
Sridhar S, Bhattacharya PK (1991) Limiting flux phenomena in ultrafiltration of kraft black liquor.
J Membr Sci 57(2–3):187–206. doi:10.1016/S0376-7388(00)80678-X
Stöcker M (2008) Bio- und BTL-Kraftstoffe in der Bioraffinerie: katalytische Umwandlung
Lignocellulose-reicher Biomasse mit porösen Stoffen. Angew Chem 120(48):9340–9351.
doi:10.1002/ange.200801476
Sun Y, Cheng J (2002) Hydrolysis of lignocellulosic materials for ethanol production: a review.
Bioresour Technol 83(1):1–11. doi:10.1016/S0960-8524(01)00212-7
Tien M (1987) Properties of ligninase from Phanerochaete chrysosporium and their possible
applications. Crit Rev Microbiol 15(2):141–168. doi:10.3109/10408418709104456
Tien M, Kirk TK (1984) Lignin-degrading enzyme from Phanerochaete chrysosporium:
Purification,characterization, and catalytic properties of a unique H2O2-requiring oxygenase.
PNAS 81(8):2280–2284.doi:10.1073/pnas.81.8.2280
Towne V, Will M, Oswald B, Zhao Q (2004) Complexities in horseradish peroxidase-catalyzed
oxidation of dihydroxyphenoxazine derivatives: appropriate ranges for pH values and
hydrogen peroxide concentrations in quantitative analysis. Anal Biochem 334(2):290–296.
doi:10.1016/j.ab.2004.07.037
Tsukihara T, Honda Y, Sakai R, Watanabe T, Watanabe T (2006) Exclusive overproduction of
recombinant versatile peroxidase MnP2 by genetically modified white rot fungus, Pleurotus
ostreatus. J Biotechnol 126(4):431–439. doi:10.1016/j.jbiotec.2006.05.013
Valderrama B, Ayala M, Vazquez-Duhalt R (2002) Suicide inactivation of peroxidases and the
challenge of engineering more robust enzymes. Chem Biol 9(5):555–565. doi:10.1016/S1074-
5521(02)00149-7
Van der Bruggen B (2009) Fundamentals of membrane solvent separation and pervaporation. In:
Membrane operations. Wiley-VCH Verlag GmbH & Co. KGaA, pp 45–61. doi: 10.1002/
9783527626779.ch3. ISBN 9783527626779
van Deurzen MPJ, Seelbach K, van Rantwijk F, Kragl U, Sheldon RA (1997) Chloroperoxidase:
use of a hydrogen peroxide-stat for controlling reactions and improving enzyme performance.
Biocatal Biotransform 15(1):1–16. doi:10.3109/10242429709003606
Viamajala S, Donohoe B, Decker S, Vinzant T, Selig M, Himmel M, Tucker M (2010) Heat and
mass transport in processing of lignocellulosic biomass for fuels and chemicals. In: Singh OV,
Harvey SP (eds) Sustainable biotechnology. Springer, Netherlands, pp 1–18. doi:10.1007/978-
90-481-3295-9_1. ISBN 978-90-481-3294-2
Vlasits J, Jakopitsch C, Bernroitner M, Zamocky M, Furtmüller PG, Obinger C (2010)
Mechanisms of catalase activity of heme peroxidases. Arch Biochem Biophys 500(1):74–81.
doi:10.1016/j.abb.2010.04.018
Wang Y, Vazquez-Duhalt R, Pickard MA (2003) Manganese-lignin peroxidase hybrid from
Bjerkandera adusta oxidizes polycyclic aromatic hydrocarbons more actively in the absence of
manganese. Can J Microbiol 49(11):675–682. doi:10.1139/w03-091
Wariishi H, Gold MH (1989) Lignin peroxidase compound III: formation, inactivation, and
conversion to the native enzyme. FEBS Lett 243(2):165–168. doi:10.1016/0014-5793(89)
80122-X
Wariishi H, Gold MH (1990) Lignin peroxidase compound III. Mechanism of formation and
decomposition. J Biol Chem 265(4):2070–2077
Wong DW (2009) Structure and action mechanism of ligninolytic enzymes. Appl Biochem
Biotechnol 157(2):174–209. doi:10.1007/s12010-008-8279-z
Yokota K, Yamazaki I (1965) The activity of the horseradish peroxidase compound 3. Biochem
Biophys Res Commun 18:48–53. doi:10.1016/0006-291X(65)90880-6
Zorn H, Peters T, Nimtz M, Berger RG (2005) The secretome of Pleurotus sapidus.
PROTEOMICS 5(18):4832–4838. doi:10.1002/pmic.200500015
Chapter 12
Fungal Aryl-Alcohol Oxidase
in Lignocellulose Degradation
and Bioconversion
Abbreviations
AAO Aryl-alcohol oxidase
DFF 2,5-Diformylfuran
FDCA 2,5-Furandicarboxylic acid
FFCA 5-Formylfurancarboxylic acid
GMC Glucose-methanol-choline oxidase/dehydrogenase superfamily
HMF 5-Hydroxymethylfurfural
HMFCA 5-Hydroxymethylfurancarboxylic acid
LiP Lignin peroxidase
Table 12.1 Comparison of the catalytic efficiencies of P. eryngii and B. adusta AAO oxidising
representative alcohol and aldehyde substrates
kcat/Km (s−1 mM−1)
P. eryngii B. adusta
Benzyl alcohol 47 ± 9 18 ± 1
OH
p-Fluorobenzyl alcohol F 59 ± 6 22 ± 2
OH
m-Fluorobenzyl alcohol 13 ± 1 47 ± 2
OH
F
Vanillyl alcohol HO 0 31 ± 1
OH
O
because they are widespread and abundant (forests cover 27 % of world’s area) and,
thus, they are cheap and can be easily stored. As a consequence, the conversion of
these materials into biofuels in biorefineries is of great interest. However, not only
biofuels, that is heat and power, are important, but other by-products obtained
during the biorefinery processes are also being carefully examined aiming at using
them for the production of valorised chemicals (Bozell and Petersen 2010). AAO is
a candidate for the enzymatic delignification of plant biomass (in synergy with
other oxidoreductases) and for the production of aromatic aldehydes and acids that
can originate from this renewable resource.
In this chapter, AAO involvement in lignocellulose decay owing to its capacity
to produce H2O2, as well as its potential applications in different industrial pro-
cesses, are discussed.
Lignocellulosic biomass (in woody and nonwoody vascular plants) accounts for the
60 % of the total carbon fixed by land photosynthesis, and is constituted of three
main polymers: cellulose, hemicellulose and lignin (Higuchi 1997). Both cellulose
and lignin are the two most abundant polymers on Earth. Lignin is located in the
middle lamella, where it attains its highest concentration, and the secondary wall of
vascular plants, together with the above polysaccharides. Its main functions are
conferring rigidity to the plants, waterproofing vessels and providing protection
against desiccation, pathogens and irradiation (Gellerstedt and Henriksson 2008).
Since lignin is very recalcitrant, it constitutes the main sink of carbon in land
ecosystems and, as a consequence, its degradation is a key step to complete the
carbon cycle. In fact, before it started to be biodegraded and mineralised, at the end
of the Carboniferous period (*300 million year ago), the carbon that the plants
fixed accumulated and therefore gave rise to the coal deposits we currently use as
source of fossil fuels (Floudas et al. 2012).
The only organisms capable of degrading wood are some saprotrophic agari-
comycetes, feeding on the simple sugars produced when cellulose and
12 Fungal Aryl-Alcohol Oxidase … 305
degradation. Moreover, the enzyme was located in the hyphal sheath (which is
formed by secreted fungal polysaccharide) during lignocellulose degradation
(Barrasa et al. 1998), as it had been previously reported for ligninolytic peroxidases
and laccases (Gallagher et al. 1989; Green et al. 1992).
The activity of AAO is supported by both non-phenolic and phenolic aryl
alcohols that can derive from fungal metabolism (de Jong et al. 1994; Gutiérrez
et al. 1994) and from lignin degradation (Kirk and Farrell 1987; Shimada and
Higuchi 1991) and are substrates for this enzyme in cooperation with related
dehydrogenases. It was seen that H2O2 is produced when adding aromatic alcohols,
as well as the corresponding aldehydes and acids, to the mycelium of P. eryngii
(Guillén et al. 1994). Thus, it was postulated that the redox cycling of these
compounds results in a continuous supply of H2O2. The product of the reaction with
AAO (aldehyde or acid) needs to be reduced in order to be re-used by the enzyme.
Therefore, it is hypothesised that the oxidised species is transported to the intra-
cellular space, where aryl-alcohol or aryl-aldehyde dehydrogenases, both
NADPH-dependent enzymes, convert it into alcohol again (de Jong et al. 1994;
Guillén and Evans 1994). Supply of reduced NADPH must be supported by the
high carbon availability for a fungus under ligninolytic conditions (Fig. 12.1).
P. eryngii AAO’s preferential substrate is p-methoxybenzyl alcohol
(Table 12.1), which is a secondary metabolite detected in fungal cultures grown
+
Fungal NADP NADPH
hypha Secondary Aryl-alcohol Catabolism
metabolism
t b li dehydrogenase
p-methoxybenzyl
alcohol H2O
+
glucose CO2
p-anisaldehyde AAO +
O2
H2O2 + Fe2+ vanillin
H2O Fenton reaction
Peroxidase OH·
Fig. 12.1 Scheme of the natural role of AAO in decay of plant cell-wall producing H2O2 for:
(i) activation of lignin-degrading peroxidases; and (ii) formation of cellulose-depolymerising
hydroxyl radical (also causing lignin oxidation)
12 Fungal Aryl-Alcohol Oxidase … 307
both in glucose and lignin media (Gutiérrez et al. 1994). It is known that aromatic
alcohols, aldehydes and acids are derived from the shikimic acid pathway of fungal
secondary metabolism (Turner and Aldridge 1983). It was seen that the oxidised
product, p-methoxybenzaldehyde (also known as p-anisaldehyde), was much more
abundant in the Pleurotus cultures than its alcoholic counterpart (Gutiérrez et al.
1994). High levels of related compounds, such as 3-chloro-p-anisaldehyde, were
also found to be abundant in B. adusta cultures (de Jong et al. 1992) and agree with
catalytic efficiencies towards the corresponding alcohol for B. adusta AAO, for
which seems to be the preferential substrate (Table 12.1). These chlorinated
derivatives are also minor extracellular metabolites in P. eryngii and P. ostreatus
(Gutiérrez et al. 1994; Okamoto et al. 2002).
These findings further supported the hypothesis postulating that AAO was acting
in the extracellular space involved in the said redox-cycling process for H2O2
production.
Based on macroscopic and chemical composition features, the wood-degrading
processes were split into two different types involving different decay mechanisms.
Some fungi leave a whitish residue, and thus were named white-rot fungi, while
others produce a brown residue and were called brown-rot fungi (Martínez et al.
2005; Schwarze et al. 2000; Zabel and Morrell 1992). Owing to genomic and
enzymatic studies, it is now thought that AAO could act as an auxiliary enzyme in
both processes due to its H2O2-producing activity and presence in the sequenced
genomes of Agaricomycotina responsible for the two types of wood-decay
processes.
(as described below for brown-rot decay) and also of lignin (Fig. 12.1) (Bes et al.
1983; Forney et al. 1982; Gómez-Toribio et al. 2009).
White-rot fungi can be classified into two groups according to their gross
degradation patterns. Some of them degrade lignin and cellulose simultaneously as
it is the case for P. chrysosporium, one of the most studied lignin-degrading
organisms. Instead, others, such as Ceriporiopsis subvermispora, degrade lignin
before cellulose (Otjen and Blanchette 1986). The main differences between the
sequenced genomes of these two fungi appear to be related to (i) the peroxidase
repertoire, and (ii) the genes involved in lipid metabolism (Fernández-Fueyo et al.
2012). The latter is related to the fact that free radicals from unsaturated lipids are
supposed to also play a role in lignin attack (Bao et al. 1994). Selective white-rot
fungi are the most interesting for industrial (biotechnological) applications in which
carbohydrates are the raw material, since they release cellulose from the lignin
matrix without significantly consuming it.
The analysis of genomes of white-rot fungi in which AAO appeared to be
produced, along with peroxidases, further confirmed its involvement in the
white-rot process as an auxiliary enzyme producing H2O2. For instance, Floudas
et al. (2012) analysed 24 basidiomycete genomes to search for enzymes involved in
the degradation of lignin. On the one hand, all white-rot genomes studied possessed
AAO genes with the only exception of Auricularia delicata. On the other hand,
AAO appears to be the most common H2O2-producing GMC, since methanol
oxidase genes are not as abundant, glucose oxidase genes are absent, and pyranose
2-oxidase genes are only found in two of the genomes studied (Table 12.2).
Table 12.2 Inventory of peroxidase and GMC genes in eleven white-rot Agaricomycotina
genomes
AD PST FM DS TV SH BA PB PC GS CS
Peroxidases 21 21 33 21 39 11 34 16 18 14 25
GMCs AAO 0 4 2 9 3 14 11 3 3 9 4
CDH 1 1 1 1 1 1 1 1 1 1 1
MOX 3 3 2 4 4 7 5 6 3 4 1
P2O 2 0 0 0 2 0 1 1 1 0 0
From Floudas et al. (2012) and Ruiz-Dueñas et al. (2013) and Ferreira et al. (2015)
AAO, aryl-alcohol oxidase; CDH, cellobiose dehydrogenase; MOX, methanol oxidase and P2O,
pyranose 2-oxidase
AD, Auricularia delicata; PST, Punctularia strigosozonata; FM, Fomitiporia mediterranea; DS,
Dichomitus squalens; TV, Trametes versicolor; SH, Stereum hirsutum; BA, Bjerkandera adusta;
PB, Phlebia brevispora; PC, Phanerochaete chrysosporium; GS, Ganoderma sp. in the
Ganoderma lucidum complex and CS, Ceriporiopsis subvermispora
12 Fungal Aryl-Alcohol Oxidase … 309
Table 12.3 Inventory of peroxidase and GMC genes (see Table 12.2 for enzyme abbreviations)
in six brown-rot Agaricomycotina genomes
CP GT FP WC DSP RP
Peroxidases 0 0 1 1 0 1
GMCs AAO 0 2 1 0 0 2
CDH 1 1 0 0 0 0
MOX 2 1 4 4 1 4
P2O 0 0 0 0 0 0
From Floudas et al. (2012) and Ferreira et al. (2015)
CP, Coniophora puteana; GT, Gloeophyllum trabeum; FP, Fomitopsis pinicola; WC, Wolfiporia
cocos; DSP, Dacryopinax sp. and RP, Rhodonia placenta
310 J. Carro et al.
studies have shown that brown-rot fungi have AAO too (Martinez et al. 2009),
suggesting that the latter oxidase might play a role in Fenton chemistry. Floudas
et al. (2012) analysed seven brown-rot fungal genomes and found AAO genes in
several of them, although they were not as abundant as those of methanol oxidases
(Table 12.3).
One of the main steps of pulp and paper industry is the treatment of wood chips to
separate cellulosic fibres (the actual raw material used) from the lignin forming the
middle lamella, a process called pulping. The concern about environment and
energy wasting has stimulated studies on the use of microorganisms to accomplish
this task. The resulting biopulping process is thought to be energetically and
environmentally more favourable than chemical and/or mechanical treatments
(Blanchette et al. 1992; Rasmussen et al. 2010).
The biopulping process starts with the colonisation of wood xylem and par-
enchyma by the fungi. After hyphae have grown on the substrate, the fungus will
start producing its ligninolytic enzymatic system in order to degrade the middle
lamellae and separate fibres (Breen and Singleton 1999). Since the loss of cellulose
is undesirable, the organisms of choice are white-rot fungi showing preference for
lignin degradation rather than cellulose degradation, whose task is to make cellu-
losic fibres accessible for the papermaking process (Scott and Swaney 1998).
Several selective ligninolytic fungi, such as C. subvermispora and P. eryngii,
together with the model white-rot fungus P. chrysosporium, and brown-rot fungi
such as Postia placenta, among others, have been investigated for biopulping of
wood and annual plants (Akhtar et al. 1997; Camarero et al. 1998; Ferraz et al.
2008; Giles et al. 2014; Masarin et al. 2009; Vicentim et al. 2009). AAO is sup-
posed to act as an auxiliary enzyme providing H2O2 in these processes, as it is the
case of natural lignin degradation.
Pulp bleaching, which is the removal of chromophores in order to obtain white
paper pulp, is another process in which AAO has shown to contribute. In a study in
which two flax pulps were treated with fungal enzymes—laccases, peroxidases and
feruroyl esterases—the ability of AAO from Pleurotus pulmonarius CBS 507.85,
which is a natural hyperproducer of this enzyme, to aid in this process was tested
(Sigoillot et al. 2005). The results showed that the presence of AAO along with
laccase improved the bleaching process probably due to the ability of AAO to
prevent the repolymerisation of the phenoxy radicals released by laccases by using
them as electron acceptors (as an alternative for O2). In a similar way, AAO can be
combined with ligninolytic peroxidases, in the presence of a substrate enabling it to
release the H2O2 required by the former enzymes.
12 Fungal Aryl-Alcohol Oxidase … 311
12.4 Bioconversions
White-rot fungi are among the most versatile flavour and aroma producers in nature
(Fraatz and Zorn 2011; Lapadatescu et al. 2000). These compounds are mainly of
aromatic nature and synthesised through biotransformations by plant, enzymatic or
microbial processes (Serra et al. 2005). Since there exists demand for naturally
produced compounds, the biotechnological production of flavours and aromas
attracts much attention (Krings and Berger 1998) due to the great economic
importance this industry has.
One of the most important flavours is vanillin, naturally produced by orchids of
the Vanilla genus, but this source only represents the 1 % of the commercial vanilla
flavour. As a consequence, several methods of obtaining vanillin have been
developed (Priefert et al. 2001) that use bioconversion of lignin and phenyl-
propanoids, such as eugenol (Overhage et al. 2003). The ability of B. adusta AAO
to oxidise vanillyl alcohol reported by Romero et al. (2009), could be exploited in
the biotechnological production of this flavour. It has been suggested that AAO
may be used to avoid formation of vanillyl alcohol as a by-product diminishing the
yield of vanillin formation (Fig. 12.2a) by the fungi of the genus Pycnoporus
(Lomascolo et al. 2011).
(c) HO O
d t AAO
B adusta
B.
646±45 s-1 mM-1
y g AAO
P. eryngii
5230±620 s-1 mM-1
O O
312 J. Carro et al.
Many of the drugs and potential drug candidates possess chiral centres and most of
them need to be commercialised as enantiomers rather than racemates given that
enantiomers often carry out different activities within biological systems (Carey
et al. 2006). Therefore, chiral intermediates for pharmaceuticals are synthesised
through enantioselective asymmetric reactions (Patel 2013). As an alternative,
deracemization of chiral mixtures is used by the pharmaceutical industry to obtain
pure enantiomers. Among chiral compounds, some secondary alcohols are used as
chiral intermediates and analytical reagents, and the development of synthesis
procedures for the production of enantiomerically enriched (enantioenriched)
alcohols has gained importance in the pharmaceutical industry.
Biological systems are generally chiral and, as a consequence, many enzymes
are regio- and enantioselective. These properties are regarded to be a consequence
of the active sites’ architecture and the enzyme’s mechanism. Therefore, many
microorganisms and enzymes offer attractive alternatives for easy production
(asymmetrical synthesis or deracemization) of enantiomeric compounds of interest
in the fine chemicals and pharmaceutical sectors (de Albuquerque et al. 2015;
Matsuda et al. 2009). The AAO catalytic mechanism consists in a hydride
abstraction from the benzylic position of the alcohol by the oxidised flavin, in a
reaction aided by an active-site histidine acting as a catalytic base to form the
alkoxide intermediate (Hernández-Ortega et al. 2012a). Due to active site
12 Fungal Aryl-Alcohol Oxidase … 313
architecture, and the simultaneous nature of the hydride and proton abstractions,
hydride transfer by AAO is stereoselective (only takes place from the pro-
R position) as shown using the two α-monodeuterated enantiomers of p-methox-
ybenzyl alcohol (Hernández-Ortega et al. 2012b). Taking advantage of this infor-
mation, Hernández-Ortega et al. (2012b) assayed the transformation of racemic
secondary alcohols using P. eryngii AAO. They saw that the enzyme was able to
oxidise the racemic 1-(p-methoxyphenyl)-ethanol, although it showed an apparent
efficiency orders of magnitude smaller than the one for p-methoxybenzyl alcohol. In
this reaction, AAO enantioselectivity towards the (S) isomer was shown using
chiral HPLC, which allows for the isolation of (R) isomer from chiral mixtures
(Fig. 12.3). Moreover, the AAO enantioselectivity towards another secondary
alcohol 1-(p-fluorophenyl)-ethanol was estimated as a S/R ratio of 21, in reactions
using the individual enantiomers.
Therefore, it is plausible that AAO could be used for the isolation of isomers
from racemates taking advantage of its kinetic behaviour. However, AAO activity
on secondary alcohols is low, due to some hindrances among such substrates and
the residues forming the active site (Fernández et al. 2009). A mutated variant
(F501A), in which the side chain of a bulky aromatic residue was removed to make
room in the cavity, was created with the purpose of facilitating oxidation of sec-
ondary alcohols. The removal of this side chain resulted in a stereoselectivity S/
R ratio on 1-(p-fluorophenyl)-ethanol three-fold higher than that of the wild-type
enzyme. Hence, improved variants by directed evolution, as it has been done with
galactose oxidase for these purposes (Escalettes and Turner 2008), or further
site-directed mutagenesis would result in better deracemization reactions.
(a) (b)
(R)-isomer (S)-isomer (R)-isomer
A225
A225
30 35 40 45 50 55 30 35 40 45 50 55
Retention time (min) Retention time (min)
Fig. 12.4 Oxidative pathways from HMF (1) to FDCA (5). FFCA (4) formation can take place
through two alternative intermediate compounds: DFF (2) or HMFCA (3)
12 Fungal Aryl-Alcohol Oxidase … 315
AAO
13±1 s-1 mM-1
1 2
O2 H2O2 47%
AAO
9±1 s-1 mM-1 UPO
2’ 4 5
53% O2 H2O2 H2O2 H2O
Fig. 12.5 Sequential enzymatic cascade for the production of FDCA (5) from HMF (1) using
P. eryngii AAO and A. aegerita UPO via DFF (2) and its hydrated gem-diol counterpart (2′)
(abundance of DFF and its gem-diol at equilibrium shown as percentage), and FFCA (4). AAO
catalytic efficiencies are indicated
316 J. Carro et al.
12.5 Conclusions
The finite character of fossil fuels makes it necessary to find new resources,
preferentially renewable ones, which allow for the substitution of the classical
resources. In this way, lignocelluloses are a remarkable material due to their
ubiquitous and renewable character. Furthermore, the concern about the environ-
ment impels us to search for new catalytic procedures that take advantage from
natural processes, instead of the chemical environmentally polluting and
energy-wasting ones. Hence, the use of organisms or enzymes to carry out catalytic
industrial processes is gaining importance.
AAO is a very promiscuous enzyme that catalyses the oxidation of a great deal
of polyunsaturated alcohols (and hydrated aldehydes). Its applicability on some
biotechnology-based industrial processes has been reviewed here and it has showed
to be promising as a biocatalyst in several conversions. AAO, because of its
involvement in the lignocellulose decay process, has potential to be applied to
lignocellulose biorefineries. AAO has so far shown its potential applications in
paper pulp manufacture and bioconversion of lignocellulose-derived compounds
(such as HMF), synthesis of flavours, and deracemization of chiral alcohols
(Fig. 12.6). In spite of the fact that the use of enzymes (among which AAO) for
these purposes is not yet sufficiently developed, much progress has been done the
recent years in this field as one can infer from the huge amount of publications
available on enzymatic biocatalysis.
Oxidation
O id ti off D
Deracemization
i ti Fl
Flavour and
d C ll l
Cellulose
furfurals of chiral alcohols aroma synthesis industries
Applications
pp cat o s
AAO
Physiology
H2O2 production
Substrate of Production of
peroxidases hydroxyl radical
Cellulose
depolymerisation
and lignin oxidation
Fig. 12.6 Scheme depicting the physiological role of AAO in lignocellulose degradation and its
potential applications in the cellulose and other chemical sectors
12 Fungal Aryl-Alcohol Oxidase … 317
Acknowledgments This work was supported by the HIPOP (BIO2011-26694) and the NOESIS
(BIO2014-56388-R) projects of the Spanish Ministry of Economy and Competitiveness, and the
INDOX (KBBE-2013-7-613549) and EnzOx2 (H2020-BBI-PPP-2015-RIA-720297) European
projects. J.C. acknowledges a FPU fellowship of the Spanish Ministry of Education, Culture and
Sport.
References
Akhtar M, Blanchette RA, Kirk TK (1997) Fungal delignification and biomechanical pulping of
wood. In: Scheper T (ed) Advances in biochemical engineering/biotechnology. Springer,
Berlin, pp 160–195
Arantes V, Jellison J, Goodell B (2012) Peculiarities of brown-rot fungi and biochemical Fenton
reaction with regard to their potential as a model for bioprocessing biomass. Appl Microbiol
Biotechnol 94:323–338
Ayers AR, Ayers SB, Eriksson K-E (1978) Cellobiose oxidase, purification and partial
characterization of a hemoprotein from Sporotrichum pulverulentum. Eur J Biochem
90:171–181
Baldrian P, Valaskova V (2008) Degradation of cellulose by basidiomycetous fungi. FEMS
Microbiol Rev 32:501–521
Bao WJ, Usha SN, Renganathan V (1993) Purification and characterization of cellobiose
dehydrogenase, a novel extracellular hemoflavoenzyme from the white-rot fungus
Phanerochaete chrysosporium. Arch Biochem Biophys 300:705–713
Bao WL, Fukushima Y, Jensen KA, Moen MA, Hammel KE (1994) Oxidative degradation of
non-phenolic lignin during lipid peroxidation by fungal manganese peroxidase. FEBS Lett
354:297–300
Barrasa JM, Gutiérrez A, Escaso V, Guillén F, Martínez MJ, Martínez AT (1998) Electron and
fluorescence microscopy of extracellular glucan and aryl-alcohol oxidase during wheat-straw
degradation by Pleurotus eryngii. Appl Environ Microbiol 64:325–332
Bes B, Ranjeva R, Boudet AM (1983) Evidence for the involvement of activated oxygen in fungal
degradation of lignocellulose. Biochimie 65:283–289
Blanchette RA, Burnes TA, Eerdmans MM, Akhtar M (1992) Evaluating isolates of
Phanerochaete chrysosporium and Ceriporiopsis subvermispora for use in biological pulping
processes. Holzforschung 46:109–115
Bourbonnais R, Paice MG (1988) Veratryl alcohol oxidases from the lignin degrading
basidiomycete Pleurotus sajor-caju. Biochem J 255:445–450
Bozell JJ, Petersen GR (2010) Technology development for the production of biobased products
from biorefinery carbohydrates—the US Department of Energy’s “Top 10” revisited. Green
Chem 12:539–554
Breen A, Singleton FL (1999) Fungi in lignocellulose breakdown and biopulping. Curr Opin
Biotechnol 10:252–258
Camarero S, Barrasa JM, Pelayo M, Martínez AT (1998) Evaluation of Pleurotus species for
wheat-straw biopulping. J Pulp Pap Sci 24:197–203
Camarero S, Böckle B, Martínez MJ, Martínez AT (1996) Manganese-mediated lignin degradation
by Pleurotus pulmonarius. Appl Environ Microbiol 62:1070–1072
Carey JS, Laffan D, Thomson C, Williams MT (2006) Analysis of the reactions used for the
preparation of drug candidate molecules. Org Biomol Chem 4:2337–2347
Carro J, Ferreira P, Rodríguez L, Prieto A, Serrano A, Balcells B, Ardá A, Jiménez-Barbero J,
Gutiérrez A, Ullrich R, Hofrichter M, Martínez AT (2015) 5-Hydroxymethylfurfural
conversion by fungal aryl-alcohol oxidase and unspecific peroxygenase. FEBS J 282
(16):3218–3229
318 J. Carro et al.
Cavener DR (1992) GMC oxidoreductases. A newly defined family of homologous proteins with
diverse catalytic activities. J Mol Biol 223:811–814
Daniel G, Volc J, Filonova L, Plihal O, Kubátová E, Halada P (2007) Characteristics of
Gloeophyllum trabeum alcohol oxidase, an extracellular source of H2O2 in brown rot decay of
wood. Appl Environ Microbiol 73:6241–6253
Daniel G, Volc J, Kubátová E (1994) Pyranose oxidase, a major source of H2O2 during wood
degradation by Phanerochaete chrysosporium, Trametes versicolor, and Oudemansiella
mucida. Appl Environ Microbiol 60:2524–2532
de Albuquerque NCP, de Gaitani CM, de Oliveira ARM (2015) A new and fast DLLME-CE
method for the enantioselective analysis of zopiclone and its active metabolite after fungal
biotransformation. J Pharm Biomed Anal 109:192–201
de Jong E, Cazemier AE, Field JA, de Bont JAM (1994) Physiological role of chlorinated aryl
alcohols biosynthesized de novo by the white rot fungus Bjerkandera sp. strain BOS55. Appl
Environ Microbiol 60:271–277
de Jong E, Field JA, Dings JAFM, Wijnberg JBPA, de Bont JAM (1992) De novo biosynthesis of
chlorinated aromatics by the white-rot fungus Bjerkandera sp. BOS55. Formation of
3-chloro-anisaldehyde from glucose. FEBS Lett 305:220–224
Dijkman WP, Fraaije MW (2014) Discovery and characterization of a 5-hydroxymethylfurfural
oxidase from Methylovorus sp strain MP688. Appl Environ Microbiol 80:1082–1090
Dijkman WP, Groothuis DE, Fraaije MW (2014) Enzyme-catalyzed oxidation of
5-hydroxymethylfurfural to furan-2,5-dicarboxylic acid. Angew Chem 126:6633–6636
Dijkman WP, Binda C, Fraaije MW, Mattevi A (2015) Structure-based enzyme tailoring of
5-hydroxymethylfurfural oxidase. ACS Catal 5:1833–1839
Eriksson K-E, Pettersson B, Volc J, Musílek V (1986) Formation and partial characterization of
glucose-2-oxidase, a H2O2 producing enzyme in Phanerochaete chrysosporium. Appl
Microbiol Biotechnol 23:257–262
Escalettes F, Turner NJ (2008) Directed evolution of galactose oxidase: generation of
enantioselective secondary alcohol oxidases. ChemBioChem 9:857–860
Evans CS, Dutton MV, Guillén F, Veness RG (1994) Enzymes and small molecular mass agents
involved with lignocellulose degradation. FEMS Microbiol Rev 13:235–240
Faison BD, Kirk TK (1983) Relationship between lignin degradation and production of reduced
oxygen species by Phanerochaete chrysosporium. Appl Environ Microbiol 46:1140–1145
Farmer VC, Henderson MEK, Russell JD (1960) Aromatic-alcohol-oxidase activity in the growth
medium of Polystictus versicolor. Biochem J 74:257–262
Fernández IS, Ruiz-Dueñas FJ, Santillana E, Ferreira P, Martínez MJ, Martínez AT, Romero A
(2009) Novel structural features in the GMC family of oxidoreductases revealed by the crystal
structure of fungal aryl-alcohol oxidase. Acta Crystallogr D Biol Crystallogr 65:1196–1205
Fernández-Fueyo E, Ruiz-Dueñas FJ, Ferreira P, Floudas D, Hibbett DS, Canessa P, Larrondo L,
James TY, Seelenfreund D, Lobos S, Polanco R, Tello M, Honda Y, Watanabe T, Watanabe T,
Ryu JS, Kubicek CP, Schmoll M, Gaskell J, Hammel KE, St. John FJ, Vanden
Wymelenberg A, Sabat G, Bondurant SS, Syed K, Yadav J, Doddapaneni H,
Subramanian V, Lavín JL, Oguiza JA, Perez G, Pisabarro AG, Ramírez L, Santoyo F,
Master E, Coutinho PM, Henrissat B, Lombard V, Magnuson JK, Kües U, Hori C, Igarashi K,
Samejima M, Held BW, Barry K, LaButti K, Lapidus A, Lindquist E, Lucas S, Riley R,
Salamov A, Hoffmeister D, Schwenk D, Hadar Y, Yarden O, de Vries RP, Wiebenga A,
Stenlid J, Eastwood DC, Grigoriev IV, Berka R, Blanchette RA, Kersten P, Martínez AT,
Vicuña R, Cullen D (2012) Comparative genomics of Ceriporiopisis subvermispora and
Phanerochaete chrysosporium provide insight into selective ligninolysis. Proc Natl Acad
Sci USA 109:5458–5463
Ferraz A, Guerra A, Mendonça R, Masarin F, Vicentim MP, Aguiar A, Pavan PC (2008)
Technological advances and mechanistic basis for fungal biopulping. Enzyme Microb Technol
43:178–185
Ferreira P, Carro J, Serrano A, Martínez AT (2015) A survey of genes encoding H2O2-producing
GMC oxidoreductases in 10 Polyporales genomes. Mycologia 107:1105–1119
12 Fungal Aryl-Alcohol Oxidase … 319
Ruiz-Dueñas FJ, Martínez AT (2009) Microbial degradation of lignin: how a bulky recalcitrant
polymer is efficiently recycled in nature and how we can take advantage of this. Microb
Biotechnol 2:164–177
Sannia G, Limongi P, Cocca E, Buonocore F, Nitti G, Giardina P (1991) Purification and
characterization of a veratryl alcohol oxidase enzyme from the lignin degrading basidiomycete
Pleurotus ostreatus. Biochim Biophys Acta 1073:114–119
Schwarze FWMR, Baum S, Fink S (2000) Dual modes of degradation by Fistulina hepatica in
xylem cell walls of Quercus robur. Mycol Res 104:846–852
Scott GM, Swaney R (1998) New technologies for papermaking: biopulping economics. Tappi J
81:153–157
Serra S, Fuganti C, Brenna E (2005) Biocatalytic preparation of natural flavours and fragrances.
Trends Biotechnol 23:193–198
Shimada M, Higuchi T (1991) Microbial, enzymatic and biomimetic degradation of lignin. In:
Hon DNS, Shiraishi N (eds) Wood and cellulosic chemistry. Marcel Dekker, New York,
pp 557–619
Sigoillot C, Camarero S, Vidal T, Record E, Asther M, Pérez-Boada M, Martínez MJ, Sigoillot
J-C, Asther M, Colom J, Martínez AT (2005) Comparison of different fungal enzymes for
bleaching high-quality paper pulps. J Biotechnol 115:333–343
Turner WB, Aldridge DC (1983) Fungal metabolites II. Academic Press, London
Ullrich R, Nuske J, Scheibner K, Spantzel J, Hofrichter M (2004) Novel haloperoxidase from the
agaric basidiomycete Agrocybe aegerita oxidizes aryl alcohols and aldehydes. Appl Environ
Microbiol 70:4575–4581
van Deurzen MPJ, van Rantwijk F, Sheldon RA (1997) Chloroperoxidase-catalyzed oxidation of
5-hydroxymethylfurfural. J Carbohydr Chem 16:299–309
Vicentim MP, Faria RD, Ferraz A (2009) High-yield kraft pulping of Eucalyptus grandis Hill ex
Maiden biotreated by Ceriporiopsis subvermispora under two different culture conditions.
Holzforschung 63:408–413
Waldner R, Leisola MSA, Fiechter A (1988) Comparison of ligninolytic activities of selected
white-rot fungi. Appl Microbiol Biotechnol 29:400–407
Yelle DJ, Ralph J, Lu F, Hammel KE (2008) Evidence for cleavage of lignin by a brown rot
basidiomycete. Environ Microbiol 10:1844–1849
Zabel R, Morrell J (1992) Wood microbiology: decay and its prevention. Academic Press, London
Zhao H, Holladay JE, Brown H, Zhang ZC (2007) Metal chlorides in ionic liquid solvents convert
sugars to 5-hydroxymethylfurfural. Science 316:1597–1600
Chapter 13
Monosaccharide Oxygenase
Abstract Oxygenases are enzymes that catalyze the insertion of oxygen atom into
an organic substrate. With the intention to perform this kind of reactions, those
enzymes want to prompt molecular oxygen to conquer its spin-forbidden response
with the organic substrate. In most cases, oxygenases make use of organic cofactors
to transfer electron to molecular oxygen for its activation. Due to chemo-, regio-,
and/or enantio-selective activity both mono- or dioxygenases are attractive bio-
catalysts. Oxygenases having many organic substrates, of these monosaccharide
reactive oxygenases are focused in this chapter. Most abundant protein on earth
known as Rubisco monooxygenase and 5-carbon monosaccharide, Sialic acid
monooxygenase is well studied but any dioxygenase capable to incorporate O2
atom in monosachharide is still unknown.
Biochemical reactions in which electrons are exchanged starting with one molecule
then onto the next are catalyzed by a wide assortment of enzymes, alluded to as
oxidoreductases or redox catalysts (EC 1.x.x.x) (Dixon and Webb 1979). Taking
into account the kind of reactions they catalyze, oxidoreductases have been sepa-
rated into 22 diverse EC-subclasses. A more broad order was proposed by Xu
(2005), in which these proteins were separated into four subgroups: (i) oxidases,
Oxydizing enzymes
(Oxygen serve as electron acceptor)
Di-oxygenase
Mono-Oxygenase
13.1.1 Oxygenases
Catalyze the incorporation of oxygen atom into the substrate. They are classified
into two groups: monooxygenase and dioxygenases. Dioxygenases consolidate
both oxygen atoms into the substrate, while monooxygenases incorporate a single
oxygen atom as a hydroxyl group into the substrate and the second oxygen atom is
reduced to water. So as to do this sort of reactions, these catalysts need to actuate
molecular oxygen to overcome its spin-forbidden reaction with the organic sub-
strate. Mostly, monooxygenases use (in)organic cofactors to exchange electrons to
molecular oxygen for its activation.
13.1.2 Monosaccharides
Oxidoreductases are getting more attention these days for selective oxygenation of
aromatic compounds and a wide range of reactions have been documented using
these enzymes from various microbial sources. Monooxygenases are an emerging
class of biotechnologically relevant enzymes due to their chemo-, regio-, and/or
enantio-selective, making those appealing biocatalysts. Monooxygenases incorpo-
rate one atom from O2 into a substrate that accepts oxygen and using another
substrate that furnishes the two H atoms to reduce the other oxygen to water.
Monooxygenases responds with O2 and supplement one atom of oxygen into the
substrate and second oxygen atom forms water with a reduced co-substrate.
326 A. Kumar et al.
S þ O2 þ RH2 ! SO þ H2 O þ R
+ +
NAD(P)H+H +O2 NAD (P) + H2O
(Styrene monooxygenase)
Fig. 13.3 Asymmetric chemical synthetic routes that catalyze selective oxygen insertion are
usually mediated by heavy metals with complex ligands and/or explosive peroxidic oxidizing
agents. A way to circumvent the use of these hazardous materials is by applying the
flavin-dependent Type I Baeyer–Villiger monooxygenases. These enzymes are able to catalyze
a broad range of oxidative reactions such as Baeyer–Villiger and heteroatom oxidations
328 A. Kumar et al.
Fig. 13.4 Proposed catalytic cycle of non-heme iron-dependent monooxygenases (Valentine et al.
1999)
13 Monosaccharide Oxygenase 329
Dioxygenases introduce both gas atom of associate gas molecule into a substrate. In
most instances, the two oxygen atoms react with one substrate molecule (in-
tramolecular dioxygenase)
S þ O2 ! SO2
Some dioxygenase, however, incorporate one atom into every of the oxygen
molecule into completely different molecules of the same substrate.
S2 þ O2 ! 2SO
S þ S0 þ O2 ! SO þ S0 O
Toluene
Industrial synthesis of
Toluene 2,3-cis-dihydrodiol
6C-Methyl-D-mannose
The enzyme oxygen complex adds to the substrate with the elimination of a
proton and the E−Fe2+ complex is regenerated.
Fig. 13.7 The likely mechanism of oxygenase reactions catalyzed by way of RuBP carboxylase–
oxygenase reactions
334 A. Kumar et al.
CO2 is likewise the sole carbon source for the predominant life-forms on this
planet, and its efficient incorporation into organic matter is directly related to the
productivity of important ecosystems, including agriculturally significant plants.
For maximum organisms, ribulose 1,5-bisphosphate (RuBP)
carboxylase/oxygenase (RubisCO) catalyzes the number one step of CO2 fixation
(Gutteridge and Gatenby 1995; Hartman and Harpel 1994; Tabita 1995); in spite of
the reality that it’s miles the most considerable protein located on earth (Ellis 1979),
RubisCO’s catalytic performance is seriously limited by way of the ability to
catalyze a competing O2 fixation response. This ends in inefficient CO2 fixation and
occasional productiveness. Right now, the molecular foundation for CO2/O2 dis-
crimination is not completely understood; however, the relative capacity of this
enzyme to choose either carboxylation or oxygenation is not immutable, but, varies
for different source of RubisCO (Jordan and Ogren 1981; Read and Tabita 1994).
The pyrenoid is a spherical structure in the chloroplast stroma, discovered more
than 130 years ago (Schmitz 1882; Vaucher 1803; Brown 1967). Pyrenoids have
been found in nearly all of the major oceanic eukaryotic primary producers and
mediate *28–44 % of global carbon fixation (Field et al. 1998; Behrenfeld et al.
2001; Rousseaux and Gregg 2013; Mann 1996; Thierstein and Young 2004; Meyer
and Griffiths 2013). A pyrenoid typically consists of a matrix surrounded by a
starch sheath and traversed by membrane tubules continuous with the photosyn-
thetic thylakoid membranes (Engel et al. 2015). This matrix is thought to consist
primarily of tightly packed Rubisco and its chaperone, Rubisco activase (McKay
and Gibbs 1991). In higher plants and non-pyrenoid-containing photosynthetic
eukaryotes, Rubisco is instead soluble throughout the chloroplast stroma. The
molecular mechanism by which Rubisco aggregates to form the pyrenoid matrix
remains enigmatic.
Two mechanisms for Rubisco accumulation in the pyrenoid have been proposed:
(i) Rubisco holoenzymes could bind each other directly through hydrophobic
residues (Meyer et al. 2012), or (ii) a linker protein may link Rubisco holoenzymes
together (Engel et al. 2015; Meyer et al. 2012). The second model is based on
analogy to the well-characterized prokaryotic carbon-concentrating organelle, the
b-carboxysome, where Rubisco aggregation is mediated by a linker protein con-
sisting of repeats of a domain resembling the Rubisco small subunit (Long et al.
2007).
The reaction, which occurs in C4 plants and bacteria, uses the high energy
phosphate of PEP to drive the incorporation of a carboxyl group in order to form
oxaloacetate. In C4 plants, PEP carboxylase is located in the mesophyll cells on the
external surfaces of the plant. Incorporation of CO2 in this manner helps to shuttle
CO2 to the bundle sheath cells, where the Calvin cycle enzymes are concentrated. It
also helps to avoid the wasteful photorespiration cycle due to the oxygenase activity
of the CO2 fixing enzyme, rubisco.
By catalyzing carbon dioxide fixation during photosynthesis, this enzyme is
responsible for virtually all of the reduced carbon found in living organisms.
Rubisco couples the inorganic and organic carbon pools on earth and is the most
significant route for linking these pools together by the synthesis of carbohydrate. It
13 Monosaccharide Oxygenase 335
Photosynthesis :
CO2 + H2O + ribulose-1,5-P2 2(3-phospho-D-glycerate) -----(Eq.1)
Photorespiration :
O2 + ribulose-1,5-P2 3-phospho-D-glycerate + 2- phosphoglycolate
-----(Eq.2)
is, however, a poor catalyst, having both a low affinity for carbon dioxide and a
small turnover number. Autotrophic organisms must devote a major part of their
synthetic capacity to produce sufficient enzyme to sustain life.
Rubisco catalyzes the first reaction in the pathways of both photosynthesis and
photorespiration, which are, respectively, its carboxylase and oxygenase activities
(Fig. 13.8).
Photosynthesis and photorespiration are interlocking metabolic cycles, with
Rubisco determining the relative rates of carbon flow through the two pathways
(Fig. 13.9). In the dark, plants are carrying out mitochondrial respiration by the
oxidation of substrates to CO2 and the conversion of O2 to H2O. On top of that,
there is another process in plants that, like mitochondrial respiration, consumes O2
and produces CO2 and, like photosynthesis, is driven by light. This process is called
photorespiration and is a costly side reaction of photosynthesis. Rubisco requires
both CO2 and Mg as obligatory cofactors, resulting in carbamylation of an active
site lysine residue by CO2 and coordination of Mg2+.
In the Oxygenase Reaction (Eq. 2) of Photorespiration However, the enediol
intermediate reacts with molecular oxygen to form a hydroperoxy derivative that
breaks down to 2-phosphoglycolate and 3-phosphoglycerate. Photorespiration
consumes ribulose bisphosphate, which results in no net gain of carbon, and also
requires the consumption of energy to recycle the lost carbon. Partitioning of
ribulose bisphosphate between the two reactions of photosynthesis and photores-
piration can vary significantly between different photosynthetic organisms, which
has stimulated interest in improving the carboxylation efficiency of crop plants.
Rubisco activity is modulated in plants by an inhibitor and an activator. The
inhibitor 2′-carboxy arabinitol 1-phosphate (2CAIP) accumulates in some plants
during darkness and binds to the active site of Rubisco. 2CAIP is degraded by a
specific phosphatase, which presumably allows Rubisco to function during pho-
tosynthesis in the light. Rubisco can be severely inhibited by a range of sugar
bisphosphates, including substrate analogs. The enzyme Rubisco activase has the
ability to relieve the inhibition caused by sugar bisphosphates, possibly by inter-
acting with Rubisco and altering the affinity of the enzyme for bisphosphates.
336 A. Kumar et al.
Pyruvate
+
Pi – elimination
3P D-Glycerate
2P-Glycolate
Addition &
Hydration
H2O
abstraction
Fig. 13.9 The reactions catalyzed by Rubisco. The first intermediate of catalysis is the C2, C3 cis-
enediol form of ribulose bisphosphate (ribulose 1,5-P2) after abstraction of the C3 proton. The
enediol can partition a number of ways, the majority into the products of carboxylation (upper
reactions) or oxygenation (lower reactions). However, a number of misprotonated isomers of
ribulose 1,5-P2, for example, xylulose-bisphosphate, have been detected with the wild-type
enzyme that are produced in quantity by mutations of specific amino acid residues involved in
proton transfer. Phosphate elimination of the carbanion forms of intermediates are also produced
by some mutants. R,—CHOH—CH20P02 v 3P D-glycerate, 3-phospho glycerate; 2P glycolate,
2-phosphoglycolate. Five discrete partial reactions have been described for the carboxylation
reaction (Eq. 1) of ribulose bisphosphate. The initial formation of a C2, C3-enediol intermediate of
the ribulose bisphosphate substrate is followed by its carboxylation, where the enediol reacts with
CO2 at the C2 position. The resulting six-carbon intermediate is hydrolytically cleaved to two
molecules of 3-phosphoglycerate, resulting in an overall gain of carbon during photosynthesis
13 Monosaccharide Oxygenase 337
the presence of ADP, and, thus, activase activity depends on the ratio of these
compounds in the chloroplast stroma. Furthermore, in most plants, the sensitivity of
activase to the ratio of ATP/ADP is modified by the stromal reduction/oxidation
(redox) state through another small regulatory protein, thioredoxin. In this manner,
the activity of activase and the activation state of RuBisCO can be modulated in
response to light intensity and, thus, the rate of formation of the ribulose
1,5-bisphosphate substrate (Zhang et al. 2002).
Regulation by Phosphate
In cyanobacteria, inorganic phosphate (Pi) participates in the coordinated regulation
of photosynthesis. Pi binds to the RuBisCO active site and to another site on the
large chain where it can influence transitions between activated and less active
conformations of the enzyme. Activation of bacterial RuBisCO might be particu-
larly sensitive to Pi levels, which can act in the same way as RuBisCO activase in
higher plants (Marcus and Gurevitz 2000).
Regulation by Carbon Dioxide
Since carbon dioxide and oxygen compete at the active site of RuBisCO, carbon
fixation by RuBisCO can be enhanced by increasing the carbon dioxide level in the
compartment containing RuBisCO (chloroplast stroma). Several times during the
evolution of plants, mechanisms have evolved for increasing the level of carbon
dioxide in the stroma (see C4 carbon fixation). The use of oxygen as a substrate
appears to be a puzzling process, since it seems to throw away captured energy.
However, it may be a mechanism for preventing overload during periods of high light
flux. This weakness in the enzyme is the cause of photorespiration, such that healthy
leaves in bright light may have zero net carbon fixations when the ratio of O2 to CO2
reaches a threshold at which oxygen is fixed instead of carbon. This phenomenon is
primarily temperature-dependent. High temperature decreases the concentration of
CO2 dissolved in the moisture in the leaf tissues. This phenomenon is also related to
water stress. Since plant leaves are evaporatively cooled, limited water causes high
leaf temperatures. C4 plants use the enzyme PEP carboxylase initially, which has a
higher affinity for CO2. The process first makes a 4-carbon intermediate compound,
which is shuttled into a site of C3 photosynthesis then de-carboxylated, releasing
CO2 to boost the concentration of CO2, hence the name C4 plants.
Crassulacean acid metabolism (CAM) plants keep their stomata (on the under-
side of the leaf) closed during the day, which conserves water but prevents the
light-independent reactions (i.e. the Calvin Cycle) from taking place, since these
reactions require CO2 to pass by gas exchange through these openings. Evaporation
through the upper side of a leaf is prevented by a layer of wax.
Sialic Acid Monooxygenase
Sialic acids are a family of nine-carbon acidic monosaccharides that occur obvi-
ously at the cease of sugar chains attached to the surfaces of cells and soluble
proteins. Sialic acids are sugars usually discovered on the outer terminal role of
glycan chains that cover the surface of all vertebrate cells inside the human body,
the very best concentration of sialic acid (as N-acetylneuraminic acid) takes place
13 Monosaccharide Oxygenase 339
Neu5Ac
Neu5Ac
LYSOSOME
Neu5Ac
CMP
CYTOSOL
CMP-Neu5Ac
CMP GOLGI
ManNAc
+
Pyruvate
Glucose
ATP N-Acetylneuramic Acid
9
ADP (Neu5Ac)
Fructose -6- PO4 8 PO4
Glutamine H2 O CMP-N-Acetylneuramic Acid + PPi
1
NH3
Glucosamine -6- PO4 +Glutamic acid N-Acetylneuramic Acid -9- PO4 +Pi
Acetyl CoA
2 7 10
PEP NADPH
N-Acetylglucosamine -6- PO4 N-Acetylmannosamine -6- PO4
3
ADP
N-Acetylglucosamine -1- PO4
CMP-N-glycolylneuramic Acid
UTP ATP
4 O2 & H 2 O
6
UDP-N-Acetylglucosamine
N-glycolylneuraminic Acid
5*
(Neu5Gc)
ManNAc
GlcNAc
Fig. 13.10 Schema displaying the metabolism of sialic acids modified from Varki et al. (1999):
1 Glycosamine-6-phoshate synthase; 2 Glucosamine-6-phosphate N-acetyltransferase; 3
N-acetylglucosamine-!6-phosphate mutase; 4 UDP-N-acetylglucosaminepyrophos-phorylase; 5 *key
enzyme: UDP-N-acetylglucosamine-2-epimerase; activity is initially low in young rats; 6
N-acetylmannosamine kinase; 7 N-acetylneuraminate-9-phosphate synthase; 8 N-acetylneuraminate-9-
phosphatephatase; 9 CMP-N-acetylneuraminatesynthetase; 10 monooxygenase
1. Higher plants incorporate large amounts of D-GluA residues into cell wall
glucuronoarabinoxylans. D-GluA forms irreversibly from UDP-D-Glc by UDP-
D-glucose dehydrogenase (EC 1.1.1.22)
+ +
2NAD 2NADH + H
NADPH + H+ NADP+
GDP-6-deoxy-D-lyxo-hexos-4-ulose GDP-L-fucose
(EC 1.1.1.271)
NADH + H+ NAD+
UDP-N-acetyl-3-O- UDP-N-
(1-carboxyvinyl)-D- acetylmuramic
glucosamine (EC 1.1.1.158) acid
O2 H2O
myo-inositol D-glucuronic
acid
(EC 1.13.99.1)
13 Monosaccharide Oxygenase 343
Appendix 2
Table 13.2 List of oxedoreductases as per their EC number—the enzymes oxidizes a substrate by
transferring the oxygen from molecular oxygen O2 (as in air) to it are called as oxygenases (EC
1.13 or EC 1.14.)
EC 1.1 Acting on the CH–OH group of donors
EC 1.2 Acting on the aldehyde or oxo group of donors
EC 1.3 Acting on the CH–CH group of donors
EC 1.4 Acting on the CH–NH2 group of donors
EC 1.5 Acting on the CH–NH group of donors
EC 1.6 Acting on NADH or NADPH
EC 1.7 Acting on other nitrogenous compounds as donors
EC 1.8 Acting on a sulfur group of donors
EC 1.9 Acting on a heme group of donors
EC 1.10 Acting on diphenols and related substances as donors
EC 1.11 Acting on a peroxide as acceptor
EC 1.12 Acting on hydrogen as donor
EC 1.13 Acting on single donors with incorporation of molecular oxygen (oxygenases)
EC 1.14 Acting on paired donors, with incorporation or reduction of molecular oxygen
EC 1.15 Acting on superoxide radicals as acceptor
EC 1.16 Oxidizing metal ions
EC 1.17 Acting on CH or CH2 groups
EC 1.18 Acting on iron-sulfur proteins as donors
EC 1.19 Acting on reduced flavodoxin as donor
EC 1.20 Acting on phosphorus or arsenic in donors
EC 1.21 Acting on the reaction X–H + Y–H = X–Y
EC 1.22 Acting on halogen in donors
EC 1.23 Reducing C–O–C group as acceptor
EC 1.97 Other oxidoreductases
References
Aboelella NW, Kryatov SV, Gherman BF, Bernnessel WW, Young VG Jr, Sarangi R,
Rybak-Akimova EV, Hogdson KO, Hedman B, Solomon EI, Cramer CJ, Tolman WB (2004)
Dioxygen activation at a single copper site: structure, bonding and mechanism of formation of
1:1 Cu-O2 adducts. J Am Chem Soc 126:16896–16911
Adams MA, Jia Z (2005) Structural and biochemical evidence for an enzymatic quinone redox
cycle in Escherichia coli—identification of a novel quinol monooxygenase. J Biol Chem
280:8358–8363
Andralojc PJ, Dawson GW, Parry MA, Keys AJ (1994). Incorporation of carbon from
photosynthetic products into 2-carboxyarabinitol-1-phosphate and 2-carboxyarabinitol.
Biochem J 304(3):781–786. PMC 1137402. PMID 7818481
344 A. Kumar et al.
Axcell BC, Geary PJ (1975) Purification and some properties of a soluble benzene-oxidizing
system from a strain of Pseudomonas. Biochem J 146:173–183
Bauwe H, Hagemann M, Fernie AR (2010) Photorespiration: players, partners and origin. Trends
Plant Sci 15(6):330–336
Behrenfeld MJ et al (2001) Biospheric primary production during an ENSO transition. Science 291
(5513):2594–2597. Abstract/FREE Full Text
Brown R (1967) Pyrenoid: its structure distribution and function. J Phycol 3(Suppl 1):5–7
Burton SG (2003) Oxidising enzymes as biocatalysts. Trends Biotechnol 21:543–549
Chambers KJ, Tonkin LA, Chang E, Shelton DN, Linskens MH, Funk WD (1998) Identification
and cloning of a sequence homologue of dopamine betahydroxylase. Gene 218:111–120
Chollet R, Anderson LL, Hovsepian LC (1975) The absence of tightly bound copper, iron, and
flavin nucleotide in crystalline ribulose 1,5-bisphosphate carboxylase–oxygenase from tobacco.
Biochem Biophy Res Commun 64:97–107
Chou HH, Takematsu H, Diaz S, Iber J, Nickerson E, Wright KL, Muchmore EA, Nelson DL,
Warren ST, Varki A (1998) A mutation in human CMP-sialic acid hydroxylase occurred after
the Homo-Pan divergence. Proc Natl Acad Sci USA 95:11751–11756
Chou HH, Hayakawa T, Diaz S, Krings M, Indriati E, Leakey M, Paabo S, Satta Y, Takahata N,
Varki A (2002) Inactivation of CMP-N-acetylneuraminic acid hydroxylase occurred prior to
brain expansion during human evolution. Proc Natl Acad Sci USA 99:11736–11741
Crafts-Brandner SJ, Salvucci ME (2000) Rubisco activase constrains the photosynthetic potential
of leaves at high temperature and CO2. Proc Natl Acad Sci USA 97(24):13430–13435.
Bibcode:2000PNAS…9713430C. doi:10.1073/pnas.230451497. PMC 27241. PMID
11069297
Dixon M, Webb EC (1979) Enzymes, 3rd edn. pp 218–220, Academic Press: New York
Ellis RJ (1979) The most abundant protein in the world. Trends Biochem Sci 4(11):241–244
Engel BD et al (2015) Native architecture of the Chlamydomonas chloroplast revealed by in situ
cryo-electron tomography. eLife 4:e04889
Ensley BD, Gibson DT (1983) Naphthalene dioxygenase: purification and properties of a terminal
oxygenase component. J Bacteriol 155:505–511
Ensley BD, Gibson DT, Laborde AL (1982) Oxidation of naphthalene by a multicomponent
enzyme system from Pseudomonas sp. strain NCIB 9816. J Bacteriol 149:948–954
Entsch B, van Berkel WJH (1995) Flavoprotein structure and mechanism. 1. Structure and
mechanism of p-hydroxybenzoate hydroxylase. FASEB J 9:476–483
Field CB, Behrenfeld MJ, Randerson JT, Falkowski P (1998) Primary production of the biosphere:
integrating terrestrial and oceanic components. Science 281(5374):237–240
Fitzpatrick PF (1999) Tetrahydropterin-dependent amino acid hydroxylases. Ann Rev Biochem
68:355–381
Fitzpatrick PF (2003) Mechanism of aromatic amino acid hydroxylation. Biochemistry 42:14084–
14091
Gallagher SC, Cammark R, Dalton H (1997) Alkene monooxygenase from Nocardia corallina
B-276 is a member of the class of dinuclear iron proteins capable of stereospecific
epoxygenation reactions. Eur J Biochem 247:635–641
Ghisla S, Massey V (1989) Mechanisms of flavoprotein catalyzed reactions. Eur J Biochem 181:1–
17
Gibson DT, Koch JR, Kallio RE (1968) Oxidative degradation of aromatic hydrocarbons by
microorganisms. I. Enzymatic formation of catechol from benzene. Biochemistry 7:2653–2662
Giordano M, Beardall J, Raven JA (2005) CO2 concentrating mechanisms in algae: mechanisms,
environmental modulation, and evolution. Annu Rev Plant Biol 56:99–131
Gutteridge S, Gatenby AA (1995) RubisCO synthesis, meeting, mechanism, and regulation. Plant
Mobile 7:809–819
Hartman FC, Harpel MR (1994) Shape, function, law and assembly of D-ribulose-1,5-bisphosphate
carboxylase/oxygenase. Annu Rev Biochem 63:197–234
Irie A, Koyama S, Kozutsumi Y, Kawasaki T, Suzuki A (1998) The molecular basis for the
absence of N-glycolylneuraminic acid in humans. J Biol Chem 273:15866–15871
13 Monosaccharide Oxygenase 345
Omura T, Sato R (1964) Carbon monoxide-binding pigment of liver microsomes. 1. Evidence for
its hemoprotein nature. J Biol Chem 239:2370–2378
Panke S, Witholt B, Schmid A, Wubbolts MG (1998) Towards a biocatalyst for (S)-styrene oxide
production: characterization of the styrene degradation pathway of Pseudomonas sp. strain.
Appl Environ Microbiol 64:2032–2043
Prigge ST, Eipper BA, Mains RE, Amzel LM (2004) Dioxygen binds end-on to mononuclear
copper in a precatalytic enzyme complex. Science 304:864–867
Read BA, Tabita FR (1994) High substrate specificity element ribulosebisphosphate
carboxylase/oxygenase from eukaryotic marine algae and properties of recombinant cyanobac-
terial Rubisco containing “algal” residue adjustments. Arch Biochem Biophys 312:210–218
Rousseaux CS, Gregg WW (2013) Interannual variation in phytoplankton primary production at a
global scale. Remote Sens 6(1):1–19. CrossRefGoogle Scholar
Rutishauser U, Acheson A, Hall AK, Mann DM, Sunshine J (1988) Science 240:53–57
Sage RF, Sage TL, Kocacinar F (2012) Photorespiration and the evolution of C4 photosynthesis.
Annu Rev Plant Biol 63:19–47
Salvucci ME, Osteryoung KW, Crafts-Brandner SJ, Vierling E (2001) Exceptional sensitivity of
Rubisco activase to thermal denaturation in vitro and in vivo. Plant Physiol 127(3):1053–1064.
doi:10.1104/pp.010357
Schlaich NL (2007) Flavin-containing monooxygenases in plants: looking beyond detox. Trends
Plant Sci 12:412–418
Schmitz F (1882) Die Chromatophoren der Algen: Vergleichende Untersuchungen über Bau und
Entwicklung der Chlorophyllkörper und der analogen Farbstoffkörper der Algen. M. Cohen &
Sohn, Bonn
Shen B, Hutchinson CR (1993) Tetracenomycin F1-monooxygenase—oxidation of a naphtha-
cenone to a naphthacenequinone in the biosynthesis of tetracenomycin-C in Streptomyces
glaucescens. Biochemistry 32:6656–6663
Smit H, Gaastra W, Kamerling JP, Vliegenthart JF, de Graaf FK (1984) Infect Immun 46:578–584
Somerville CR, Ogren WL (1979) A phosphoglycolate phosphatase-deficient mutant of
Arabidopsis. Nature 280(5725):833–836
Suzuki Y, Nagao Y, Kato H, Suzuki T, Matsumoto M, Murayama J (1987) Biochim Biophys Acta
903:417–424
Tabita FR (1995) The biochemistry and metabolic regulation of carbon metabolism and CO2
fixation in red micro organism. In: Blankenship RE, Madigan MT, Bauer CE (eds) Anoxygenic
photosynthetic micro organism. Kluwer, Dordrecht, pp 885–914
Takabe T, Akazawa T (1975) Biochemistry 14:46–50
Thierstein HR, Young JR (eds) (2004) Coccolithophores: from molecular processes to global
impact. Springer, Heidelberg
Torres PDE, Winkler M, Glieder A, Fraaije MW (2010) Monooxygenases as biocatalysts:
classification, mechanistic aspects and biotechnological applications. J Biotechnol 146:9–24
Valentine AM, Stahl SS, Lippard SJ (1999) Mechanistic studies of the reaction of reduced
methane monooxygenase hydroxylase with dioxygen and substrates. J Am Chem Soc
121:3876–3887
van Beilen JB, Funhoff EG (2007) Alkane hydroxylases involved in microbial alkane degradation.
Appl Microbiol Biotechnol 74:13–21
van Berkel WJH, Kamerbeek NM, Fraaije MW (2006) Flavoprotein monooxygenases, a diverse
class of oxidative biocatalysts. J Biotechnol 124:670–689
van Hellemond EW, Janssen DB, Fraaije MW (2007) Discovery of a novel styrene monooxy-
genase originating from the metagenome. Appl Environ Microbiol 73:5832–5839
Varki A (1992) Diversity in the sialic acids. Glycobiology 2:25–40
Varki A, Cummings R, Kesko J, Freeze H, Hart G, Marth J (1999) Sialic acids. In: Varki A,
Cummings R, Kesko J et al. (eds) The Essentials of Glycobiology. Cold Spring Harbor, NY:
Cold Spring Harbor Laboratory Press, pp 195–209
13 Monosaccharide Oxygenase 347
Vaucher J-P (1803) Histoire des Conferves D’eau Douce: Contenant Leurs Différents Modes De
Reproduction, Et La Description De Leurs Principales Espèces, Suivie De L’histoire Des
Trémelles Et Des Ulves D’eau Douce. JJ Paschoud, Geneva
Wackett LP, Hershberger CD (2001) Biocatalysis and biodegradation: microbial transformation of
organic compounds, st edn. ASM Press, Washington
Wallar BJ, Lipscomb JD (1996) Dioxygen activation by enzymes containing binuclear non-heme
iron clusters. Chem Rev 96:2625–2657
Whited GM, Gibson DT (1991) Toluene-4-monooxygenase, a three-component enzyme system
that catalyzes the oxidation of toluene to p-cresol in Pseudomonas mendocina KR1. J Bacteriol
173:3010–3016
Xin X, Mains RE, Eipper BA (2004) Monooxygenase X, a member of the copper-dependent
monooxygenase family localized to the endoplasmic reticulum. J Biol Chem 279:48159–48167
Xu F (2005) Applications of oxidoreductases: recent progress. Ind Biotechnol 1:38–50
Yeh W-K, Gibson DT, Liu T-N (1977) Toluene dioxygenase: a multicomponent enzyme system.
Biochem Biophys Res Commun 78:401–410
Zhang N, Kallis RP, Ewy RG, Portis AR (2002) Light modulation of Rubisco in Arabidopsis
requires a capacity for redox regulation of the larger Rubisco activase isoform. Proc Natl Acad
Sci USA 99(5):3330–3334. Bibcode:2002PNAS…99.3330Z. doi:10.1073/pnas.042529999.
PMC 122518. PMID 11854454