Sei sulla pagina 1di 14

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/236459056

Lignocellulose-Degrading Enzymes from


Termites and Their Symbiotic Microbiota.

Article in Biotechnology advances · April 2013


DOI: 10.1016/j.biotechadv.2013.04.005 · Source: PubMed

CITATIONS READS

49 924

2 authors, including:

Gaku Tokuda
University of the Ryukyus
86 PUBLICATIONS 2,492 CITATIONS

SEE PROFILE

All content following this page was uploaded by Gaku Tokuda on 03 April 2017.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
Biotechnology Advances 31 (2013) 838–850

Contents lists available at ScienceDirect

Biotechnology Advances
journal homepage: www.elsevier.com/locate/biotechadv

Research review paper

Lignocellulose-degrading enzymes from termites and their symbiotic microbiota


Jinfeng Ni a,⁎, Gaku Tokuda b,⁎⁎
a
State Key Laboratory of Microbial Technology, Shandong University, 27 Shandanan Road, Jinan, Shandong 250100, China
b
Tropical Biosphere Research Center, COMB, University of the Ryukyus, Nishihara, Okinawa, Japan

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

Available online 23 April 2013 Lignocellulose—the dry matter of plants, or “plant biomass”—digestion is of increasing interest in organis-
mal metabolism research, specifically the conversion of biomass into biofuels. Termites efficiently decom-
Keywords: pose lignocelluloses, and studies on lignocellulolytic systems may elucidate mechanisms of efficient
Termites lignocellulose degradation in termites as well as offer novel enzyme sources, findings which have signifi-
Lignocellulose degradation cant potential industrial applications. Recent progress in metagenomic and metatranscriptomic research
Cellulase
has illuminated the diversity of lignocellulolytic enzymes within the termite gut. Here, we review
Xylanase
Laccase
state-of-the-art research on lignocellulose-degrading systems in termites, specifically cellulases, xylanases,
and lignin modification enzymes produced by termites and their symbiotic microbiota. We also discuss re-
cent investigations into heterologous overexpression of lignocellulolytic enzymes from termites and their
symbionts.
© 2013 Elsevier Inc. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 838
2. The termite intestinal tract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 839
3. Cellulose and cellulase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 839
3.1. Cellulolytic systems in lower termites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 839
3.2. Cellulolytic systems in higher termites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 841
4. Hemicellulose and hemicellulase. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 842
4.1. Xylanases from lower termites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 842
4.2. Xylanases from higher termites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 843
5. Lignin and its modification enzymes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 844
5.1. Structural changes of lignin during passage of the gut . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 844
5.2. Lignin-modification enzymes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 845
6. Heterologous production of lignocellulases from termites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 845
6.1. Heterologous production of EGs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 845
6.2. Heterologous production of BGs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 846
6.3. Heterologous production of xylanases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 847
7. Future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 848
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 848
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 848

1. Introduction (Himmel et al., 2007; Ragauskas et al., 2006). Termites, which are ef-
ficient lignocellulose decomposers, thrive on dead plant materials and
Lignocellulose, mainly comprised of cellulose, hemicelluloses, and contribute to carbon mineralization, especially in tropical and sub-
lignin, is the most abundant biomass on earth and is recognized as a tropical regions (Kudo, 2009; Ohkuma, 2003; Yamada et al., 2005).
potential sustainable source for biofuels and biomaterial production Studies to elucidate termite lignocellulose-degrading systems should
identify many cellulose hydrolysis enzymes and enhance our under-
standing of mechanisms of lignocellulose degradation in termites
⁎ Corresponding author. Tel.: +86 531 88363323; fax: +86 531 88362903.
⁎⁎ Corresponding author. Tel.: +81 98 895 8543; fax: +81 98 895 8944.
(Scharf and Tartar, 2008). Such investigations may also contribute
E-mail addresses: jinfgni@sdu.edu.cn (J. Ni), tokuda@comb.u-ryukyu.ac.jp to optimization of plant biomass bioconversion processes. Therefore,
(G. Tokuda). termite biomass degradation has received significant interest in the

0734-9750/$ – see front matter © 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.biotechadv.2013.04.005
J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850 839

last decade. Using “termite” and “enzyme” as key words to search the of ectodermal origin, is the largest organ. The hindgut can be further
National Center for Biotechnology Information's database of biomed- subdivided into P1, P2, P3, P4, and P5 segments (Watanabe and
ical literature (PubMed) we found 238 papers: 230 manuscripts dated Tokuda, 2010). Among these segments, P3 is typically enlarged to harbor
from 2000 to the present (November 27, 2012) and 38 papers pub- numerous microorganisms. The gut microbial community contains all
lished prior to 2000 (1978–1999). Such findings indicate the in- three domains of organisms: Archaea, Bacteria, and Eukaryotes (pro-
creased research emphasis in the biology of termite enzymes for the tists). Relatively little microbiota is found in the foregut and midgut,
last 10 years. whereas abundant microbiota is found in the hindgut (Hongoh, 2011;
Termites (Isoptera or Termitoidae) are well-studied social in- Köhler et al., 2012). The hindgut compartments of higher termites are
sects, considered to be an epifamily of the cockroach (Blattaria or more developed and complex than those of lower termites. Except for
Blattodea). The diversity and classification of termites are well docu- Macrotermitinae, Sphaerotermitinae, and Foraminitermitinae, higher
mented elsewhere (Eggleton, 2011; Lo and Eggleton, 2011), so we termites have developed a “mixed segment” (half of the gut wall consists
have briefly summarized important facts here. Termite group is com- of midgut tissue; the remaining is hindgut tissue) between the mid-
prised of more than 2600 described species, leaving perhaps 500– gut and the hindgut, but the precise function of the mixed segment
1000 species currently undescribed (Eggleton, 2011). Termites are gen- has not been elucidated. In addition to the intestinal tract, the sali-
erally classified into seven families: Mastotermitidae, Kalotermitidae, vary glands also significantly contribute to the digestive physiology
Termopsidae, Hodotermitidae, Rhinotermitidae Serritermitidae, and of termites. Detailed descriptions of termite gut structures are docu-
Termitidae (Lo and Eggleton, 2011). The Mastotermitidae are the most mented elsewhere (see Bignell, 1994, 2011; Lo and Eggleton, 2011
primitive family, containing only one wood-feeding species in Australia. and references therein).
Termites of the Kalotermitidae family consume material from dry wood,
and Termopsidae nest in and feed on wet dead logs. Hodotermitidae are 3. Cellulose and cellulase
grass-harvesters, and Rhinotermitidae consume wood, primarily in
temperate zones. Serritermitidae contains few species and only in South Lignocellulose consists of cellulose (20–50%), hemicellulose (15–35%),
America, suggesting that they may be better classified within the and lignin (18–35%). Cellulose, a linear polysaccharide consisting of
Rhinotermitidae family (Eggleton, 2011). Termitidae is the largest fam- β-1,4-linked D-glucopyranosyl units, is the major component of plant
ily, consisting of approximately 2000 species and accounting for material (20–40%) and the most abundant biomass on earth (Tomme et
almost 75% of all known termites. Termitidae consists of at least al., 1995). Cellulases, found in the gut of lower and higher termites, are
seven sub-families, namely Macrotermitinae, Sphaerotermitinae, produced by organisms that catalyze the cellulolysis (or hydrolysis)
Foraminitermitinae, Apicotermitinae, Termitinae, Syntermitinae, of cellulose, and three classes have been identified in cellulolysis.
and Nasutitermitinae (Eggleton, 2011; Lo and Eggleton, 2011). Endo-β-1,4-glucanases (EC 3.2.1.4) hydrolyze cellulose chains in a
Termitidae have diverse feeding preferences, the majority feeding non-processive (random) manner, whereas exoglucanases such as
on soil. A few subfamilies such as Termitinae, Nasutitermitinae, and cellodextrinases (EC 3.2.1.74) or cellobiohydrolases (EC 3.2.1.91)
Syntermitinae contain wood- or litter-feeding termites. Termites of depolymerize cellulose chains from their reducing or non-reducing
the sub-family Macrotermitinae cultivate basidiomycete fungi ends in a processive or ordered manner. β-Glucosidases (EC 3.2.1.21)
(Termitomyces sp.) in their nests, and are thus commonly known as cleave cello-oligosaccharides (especially cellobiose) to liberate glucose.
fungus-growing termites.
Based on the presence or absence of flagellated protistan symbionts 3.1. Cellulolytic systems in lower termites
in the hindgut of termites, they are conventionally grouped into lower
and higher termites. The first six families, all of which harbor protistan Distribution patterns of cellulolytic enzymes in the gut of termites
symbionts in the hindgut, are referred to as “lower termites”. The have been studied extensively, but relevant studies on soil-feeding
remaining family, Termitidae, which lack protistan symbionts in the termites are limited. The distribution patterns and expression of cel-
hindgut, are traditionally referred to as “higher termites” (Lo and lulolytic enzymes in the termite gut varied by termite caste and de-
Eggleton, 2011). velopmental stages (Fujita et al., 2008; Shimada and Maekawa,
Studies of cellulose digestion in termites and related insects and the 2010). Thus, this review focuses on mature worker-caste termites
historical importance of these findings have been well-documented that feed on lignocellulosic materials and have significant digestive
elsewhere (Hongoh, 2011; Lo et al., 2011; Matsui et al., 2009; roles among the castes. Generally, lower termites possess strong hy-
Watanabe and Tokuda, 2001, 2010). In this review, we have focused on drolytic activity (45–85% of total gut activity) against carboxymethyl-
the lignocellulolytic system in termites, with special reference to cellulose (CMC) (representing endo-β-1,4-glucanase [EG] activity) in
lignocellulose-degrading enzymes such as cellulases, xylanases, and the salivary glands (Tokuda et al., 2004), whereas these termites pos-
laccases produced by termites and their symbiotic microbiota. We also sess stronger cellulolytic activity (40–88%) in the hindgut than in the
describe enzyme gene resources, heterologous overexpression systems, salivary glands when microcrystalline cellulose is the substrate (pri-
and relevant enzyme properties that hold promise for industrial marily representing cellobiohydrolase [CBH] activity) (Tokuda et al.,
applications. 2005). Regarding β-glucosidase (BG), lower termites have stronger
enzymatic activities both in the salivary glands and the hindgut
2. The termite intestinal tract than higher termites (Slaytor, 2000; Tokuda et al., 2002). The relevant
genes encoding these enzymes have been identified (Lo et al., 2011;
The intestinal tracts of termites are axially structured microenviron- Watanabe and Tokuda, 2010). Based on peptide sequence similarities,
ments with differences in metabolic activities and microbial community glycoside hydrolases are classified into more than 100 families (see
structures (Köhler et al., 2012). The termite gut generally consists of the http://www.cazy.org). Members of the same family frequently have
foregut, midgut, and hindgut. The foregut is an esophageal tract of ecto- different substrate specificities but share structural similarities that,
dermal origin with an enlarged anterior segment (“the crop”) and a pos- in turn, reflect their evolutionary origins (Henrissat and Bairoch,
terior segment (“the gizzard”) that plays a role in mechanical grinding of 1993). According to this classification, all endogenous EGs are affiliat-
ingested wood fragments. The midgut, which in insects is chiefly a secre- ed with the glycoside hydrolase family (GHF) 9 (Leonardo et al.,
tion site for digestive enzymes and nutrient absorption, is columnar and 2011; Tartar et al., 2009; Watanabe and Tokuda, 2010; Zhang et al.,
uniform with an endodermal origin. The midgut is located posterior to 2012b), while all endogenous BGs belong to GHF1, except for one
the foregut. The Malpighian tubules are usually attached to the end of putative endogenous GHF3 BG which has been identified from the
the midgut to excrete nitrogen wastes to the gut lumen. The hindgut, salivary gland EST library of Hodotermopsis sjostedti (Termopsidae)
840 J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850

(Yuki et al., 2008). Detailed cellulase gene structures of termites are GHF1 enzymes (Scharf and Tartar, 2008; Tartar et al., 2009; Todaka
reviewed by Lo et al. (2011). et al., 2007; Xie et al., 2012).
Compared to host cellulases, symbiotic protistan communities in As shown in Fig. 1, lower termites produce both EGs and BGs primar-
lower termites produce more complex cellulolytic enzymes. Meta- ily in the salivary glands. These endogenous cellulases probably hydro-
transcriptomic sequencing techniques have provided comprehensive lyze amorphous regions of cellulose during its passage through the
information about cellulolytic enzymes from the protistan symbionts of midgut. A combination of endogenous EG and BG can release glucose
five genera of termites, Mastotermes darwiniensis (Mastotermitidae), from filter paper (Zhang et al., 2010) and pine lignocelluloses (Scharf
Neotermes koshunensis (Kalotermitidae), H. sjostedti, Reticulitermes et al., 2011). An essential contribution of an endogenous EG to termite
speratus (Rhinotermitidae), Reticulitermes flavipes, and Coptotermes survival and fitness has been verified with RNAi investigations (Zhou
formosanus (Rhinotermitidae) in addition to the closely-related wood et al., 2008b). Partially degraded cellulosic fragments being moved to
roaches Cryptocercus punctulatus (Scharf and Tartar, 2008; Sethi et al, the hindgut are endocytosed by the symbiotic protists. Because no (or
2013; Tartar et al., 2009; Todaka et al., 2007, 2010a; Xie et al., 2012). In only a trace of) endogenous cellulases enter the hindgut (Fujita et al.,
the protistan cellulolytic system, GHF5 EGs, GHF7 EGs, and GHF7 CBHs 2010; Nakashima et al., 2002; Watanabe et al., 2006), cellulolysis in
appear to function as a core enzyme set that might be conserved across food vacuoles is conceivably accomplished synergism of multiple cellu-
protistan generations during the history of symbiosis between termites lolytic enzymes that originated from protists. Because ingested cellulos-
and intestinal protists. On the other hand, the GHF45 EGs might supple- ic materials are masticated and fragmented (less than 20 μm in the
ment these core cellulases; this family is absent from N. koshunensis and foregut; Fujita et al., 2010), this micro-fragmentation of food could
forms multiple clades in the phylogenetic tree (Todaka et al., 2010a). have increased surface area, facilitating access of cellulolytic enzymes
The symbiotic protists primarily express GHF3 BGs, which are pre- and stimulating endocytosis by symbiotic protists. In addition, the pres-
dicted to perform similar glucose release functions as endogenous ence of CBHs in the protistan cellulolytic system apparently confers

Fig. 1. Cellulolytic systems in lower and higher termites. Lower termites secrete endogenous endo-β-1,4-glucanases (EGs) and β-glucosidases (BGs) in the salivary glands. These enzymes
are probably mixed with wood particles ingested by the termites and may act primarily against amorphous cellulose. Chewed wood particles are partially degraded during passage
through the midgut. Remaining wood particles enter the hindgut, where they are digested in the food vacuoles of symbiotic flagellates with protistan EGs, BGs, and cellobiohydrolases
(CBHs) that are highly active against crystalline cellulose. These protists also produce hemicellulolytic enzymes. Such cellulolytic systems of wood-feeding higher termites have been
chiefly studied in Nasutitermitinae, which secrete endogenous BGs both in the salivary glands and the midgut, whereas endogenous EGs are secreted only in the midgut. These enzymes
also act primarily against amorphous cellulose, but may have an undescribed mechanism that enhances digestibility of crystalline cellulose (Tokuda et al., 2012). Details of this endoge-
nous cellulolytic system are depicted in Fig. 3. The hindgut harbors symbiotic bacteria and archaea. Among them, Treponema and Fibrobacteres are primarily involved in cellulolysis with
EGs and BGs. Like these cellulases, cellobiose- and cellodextrin-phosphorylases (GHF94) appear to be involved in cellulolysis, but there is no evidence that CBHs are present in the hind-
guts of higher termites. Nevertheless, the hindgut precipitates contain activity against crystalline cellulose (Tokuda and Watanabe, 2007). Many different hemicellulolytic enzymes are
also produced by these bacterial symbionts.
J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850 841

efficient digestibility of crystalline cellulose in the hindgut. This dual termite Odontotermes formosanus, in which EG genes are still expressed
cellulose digestion system has been proposed as a model of efficient cel- in the salivary glands (Tokuda et al., 2004). Endogenous GHF1 BGs of
lulose hydrolysis in lower termites (Nakashima et al., 2002; Tokuda et N. takasagoensis were expressed and secreted in the midgut and the sal-
al., 2007; Watanabe and Tokuda, 2010). ivary glands (Tokuda et al., 2009, 2012), whereas the fungus-growing
termite Macrotermes barneyi expressed an endogenous GHF1 BG gene
3.2. Cellulolytic systems in higher termites primarily in the midgut (Wu et al., 2012).
Recent metagenomic analyses have clarified the presence of diverse
Distribution of cellulase activities in higher termites is more variable. cellulase genes in the hindgut microflora of higher termites. Among 45
The reason for this is somewhat obscure, but ecological events may GHFs discovered in the hindgut microbial communities of Nasutitermes
have affected cellulolytic systems during the evolution of higher termites sp., seven GHFs (i.e. GHF5, GHF8, GHF9, GHF44, GHF45, GHF51, and
(Fig. 2). In most wood-feeding termites of the subfamily Nasutitermitinae, GHF74) might be acting as EGs, whereas GHF1 and GHF3 might partici-
EG activities are confined to the midgut, but Nasutitermes lujae share an pate in cellulolysis as BGs (Warnecke et al., 2007). In contrast, many
almost equal amount of EG activity both in the midgut and the hindgut GHF94 genes that encode cellodextrin or cellobiose phosphorylases
(Slaytor, 2000). In addition, some species of Termitinae have most EG were present in the metagenomic library, suggesting that several bacteria
activity in the hindgut, which is presumably attributable to the symbiotic degrade cello-oligosaccharides via phosphorylation without depending
amoeba (Slaytor, 2000) whose nature has not been clarified yet. Similar to on BGs. The presence of some of these GHFs in the hindgut was further
the distributions of EG activities, distribution of BG activities is also confirmed by a proteomic analysis (Burnum et al., 2011). Most of these
variable. Some termites, including Nasutitermes spp., have BG activity GHF genes were assumed to be derived from Spirochetes (genus Trepone-
primarily in the salivary glands and the midgut (Slaytor, 2000; Tokuda ma) or Fibrobacteres. Metagenomic analysis did not identify potential
et al., 1997). Other termites possess up to 99.5% BG activity in the hind- CBHs or components of cellulosomes (complex structures of cellulases
gut (Slaytor, 2000). Distribution of cellulase activities in the fungus- and carbohydrate-binding proteins that adhere to the cell surface of an-
growing termites (subfamily Macrotermitinae) is more complicated aerobic cellulolytic bacteria via docking proteins, cohesins and dockerins
(Rouland-Lefèvre, 2000) partially due to varying dependence on the (Lynd et al., 2002)). However, these results do not exclude the possibility
symbiotic fungi among species as a nutritional carbon source (Hyodo of efficient degradation of crystalline cellulose by hindgut microflora. A
et al., 2003). Contributions of the hindgut microorganisms to cellulose third mechanism of cellulose digestion is emerging from studies of
digestion in higher termites have long been neglected, but wood- Fibrobacter spp., which can disrupt crystalline structure of cellulose with-
fiber associated cellulase activities were reported in the hindgut of out exoglucanases or processive EGs (Ransom-Jones et al., 2012). A pro-
Nasutitermes takasagoensis and Nasutitermes walkeri (Tokuda and tein complex in the outer membrane of the cell is predicted to bind to
Watanabe, 2007). These cellulase activities were significantly de- cellulose fibers, cleave individual cellulose chains, and transport them
creased if antibiotics were administrated to termites. These bacterial to the periplasmic space where EGs cleave these cellulose chains
cellulolytic activities in the hindgut accounted for ~ 50% compared to (Wilson, 2009). Indeed, interactions between the intestinal bacteria and
cellulase activity in the midgut if microcrystalline cellulose was the lignocellulosic fragments were observed under an electron microscope
substrate (Tokuda and Watanabe, 2007). Thus, there is strong evi- in the anterior paunch (P3) of N. takasagoensis (Tokuda et al., 2005).
dence that the endogenous cellulolytic system in the midgut of One GHF5, EG, has been found in the hindgut metagenomic library of an-
higher termites is more important than that of lower termites. As with other xylophagous termite, Microcerotermes sp. (subfamily Termitinae)
lower termites, endogenous EGs in higher termites are affiliated with (Nimchua et al., 2012), from which cellulolytic Bacillus subtilis was isolat-
GHF9 and are secreted from the midgut, except for the fungus-growing ed (Taechapoempol et al., 2011). Functional screening of a fosmid library
from microbial DNA extracted from the gut of the fungus-growing ter-
mite Macrotermes annandalei failed to identify CBHs and EGs. Rather,
Lower termites
BG genes were found, but their affiliations were not determined (Liu et
al., 2011). With EST analysis of the symbiotic fungi Termitomyces sp.
Macrotermitinae from Macrotermes gilvus, identification is confirmed for all necessary cel-
lulolytic enzymes, CBHs (GHF6 and GHF7), EGs (GHF5, 44, 61), and BGs
(GHF1 and GHF3) (Johjima et al., 2006).
Sphaerotermitinae As mentioned above, the endogenous cellulolytic system of wood-
Loss of protists/
Fungus-growing feeding higher termites is thought to contribute to cellulose digestion
habit acquired
more significantly than lower termites. Fig. 3 illustrates the endogenous
Bacterial comb- Foraminitermitidae cellulolytic system of Nasutitermes as a representative of wood-feeding
growing habit
acquired? higher termites based on Tokuda et al. (2012). Masticated cellulosic frag-
Transition to
ments would be first attacked by salivary BGs that do not enter the
soil/humus-feeding
Apicotermitinae midgut, but the actual role of salivary BGs is ambiguous. A salivary BG
habit
tends to more preferentially hydrolyze β-1,3-glucans (found in fungal
Syntermitinae cell walls) than β-1,4 linkages generally recognized in cellulose chains
(Uchima et al., in press). A recent study on labial gland secretion of
Wood/litter-feeding habit termites indicated the presence of p-arbutin that is hydrolyzed by BG to
re-acquired in some Termitinae produce hydroquinone, a precursor of p-benzoquinone, an irritating de-
species?
Acquisition of cellulolytic fensive secretion (Sillam-Dussès et al., 2012). In addition, a possible func-
amoeba in some soil- Nasutitermitinae tion of BG as a proteinaceous pheromone and reproductive suppressor
feeding species
has also been suggested (Korb et al., 2009; Matsuura et al., 2009). It is
Fig. 2. Possible ecological events that could have affected the cellulolytic systems in probable that salivary BGs in higher termites play roles other than
higher termites. Higher termites (i.e. family Termitidae) have currently been thought cellulolysis. Ingested cellulosic particles are fragmented into less than
to exist in seven subfamilies. Phylogenetic relationships shown here were based on In- 150 μm in the foregut. The midgut epithelium secretes both EGs and
ward et al. (2007) and Lo and Eggleton (2011). Although not illustrated here, BGs that make essential contributions to cellulose digestion. Luminal con-
Termitinae and Nasutitermitinae might be polyphyletic. Ecological events potentially
affecting the cellulolytic systems during the evolution of higher termites are indicated
centrations of these enzymes are extremely high (~1500 U/ml for EG and
to the left of the tree. These events were assumed from Slaytor (2000), Inward et al. ~80 U/ml for BG), which may be important for primarily hydrolyzing
(2007) and Garnier-Sillam et al. (1989). amorphous regions of cellulose. Although a purified or recombinant EG
842 J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850

Fig. 3. Endogenous digestive system in N. takasagoensis. Wood fragments ingested by the termites may first mix with the salivary BGs. The wood particles in the foregut are broken
down to 150 μm in diameter and enter the midgut. The midgut tissue secretes both EGs and BGs into the luminal space. Based on the midgut luminal volume (approx. 75 nl), the
concentrations of EG and BG activities are approximately 1500 units/ml and 80 units/ml, respectively (one unit is the amount of enzyme that produces 1 μmol of reducing sugar or
glucose/min). It is presumed that amorphous regions of cellulose exposed on the surface of the wood particles are digested primarily by these enzymes, and released glucose might
be absorbed across the midgut tissue. Finally, the partially digested wood particles are moved to the hindgut via the mixed segment. The role that the mixed segment plays in di-
gestion has yet to be clarified. PM, peritrophic membrane.

of this termite shows trace activity against microcrystalline cellulose chains of hemicellulose are composed of xylose, or glucose and mannose,
(Hirayama et al., 2010; Tokuda et al., 1997), the midgut crude enzyme which is often acetylated or shortly branched with arabinose, galactose, or
can hydrolyze highly crystalline cellulose to some extent (Tokuda et al., other acidic sugars. Chemical compositions of hemicelluloses vary across
2012). However, a detailed mechanism of this phenomenon has yet to plant species: large differences are documented to exist between gymno-
be elucidated. Short cello-oligosaccharides released by these enzymes sperms and angiosperms. Usually, hemicelluloses of gymnosperms con-
would be degraded in the ectoperitrophic space where BGs accumulate sist primarily of acetylated glucomannan, arabinoglucronoxylan, and
and glucose products would be absorbed across the midgut wall. As galactoglucomannan. In contrast, hemicelluloses of angiosperms primari-
shown in Fig. 1, remaining cellulosic fragments enter the mixed segment ly contain non-acetylated glucuronoxylan and glucomannan. Thus, the di-
and then the hindgut, but as mentioned before, the role of the mixed seg- gestion of hemicelluloses requires a large set of poly- and di-saccharidases
ment in cellulose digestion is unclear. Cellulosic fragments would be fur- which include xylanases (EC 3.2.1.8), β-xylosidases (EC 3.2.1.37),
ther hydrolyzed by prokaryotic cellulases in the hindgut. However, the α-glucuronidase (EC 3.2.1.131), α-arabinofuranosidase (EC 3.2.1.39)
detailed mechanism underlying the cellulolytic system requires further and acetylxylan esterase (EC 3.1.1.72) (Juturu and Wu, 2012; Liu et al.,
study. In addition, it is likely that other wood-feeding higher termites, es- 2011; Saha, 2003). Xylan, a linear polymer of β-xylopyrannosyl units
pecially those belonging to Termitinae, have different cellulolytic systems linked by β-1,4-glycosidic bonds, is the major component of hemicellu-
than Nasutitermes (Slaytor, 2000). Hence, further explorations of xyloph- lose. Xylanases and β-xylosidases cleave the backbone of xylan to release
agous species are required to fully understand cellulolytic systems in xylose (Ahmed et al., 2009). Addition of xylanase to a cellulase mixture
higher termites. In fungus-growing termites, clear conclusions are not enhances enzyme accessibility to cellulose in a lignocellulosic substrate
available regarding the mechanism of cellulose digestion. For example, via increases in fiber swelling and fiber porosity, significantly improving
in the case of O. formosanus, intestinal cellulase activity may be insuffi- hydrolytic efficiency (Hu et al., 2011). Glucose residues present in hemi-
cient to produce the necessary amount of glucose required for respiration cellulose are also susceptible to cellulases (Scharf et al., 2011). Although
(Tokuda et al., 2005) and the symbiotic fungus is likely a sole carbon multiple hemicellulolytic enzymes could collaboratively act against ligno-
source based on the stable carbon isotope ratio (Hyodo et al., cellulose with cellulases, studies on hemicellulose degradation in termites
2003). In spite of these studies, its cellulolytic mechanism has been are limited compared to what is known about cellulose digestion. Thus,
investigated recently, suggesting an “acquired enzyme hypothesis”— we have primarily focused on xylanases.
that lignocellulolytic enzymes acquired from the symbiotic fungi act co-
operatively with endogenous digestive enzymes in termites (Deng et 4.1. Xylanases from lower termites
al., 2008a,b).
In lower termites examined to date, the majority of xylanase activities
4. Hemicellulose and hemicellulase are confined to the hindgut (e.g. Arakawa et al., 2009; Smith et al., 2009;
Zhou et al., 2007), although a report suggests that (1) host GHF9 + GHF1
Hemicellulose is a general term for major noncellulosic polysaccha- enzymes together can effectively release glucose from beechwood xylan,
rides in plant cell walls which are the second most common polysaccha- and that (2) a host laccase could enhance this (Scharf et al., 2011). The
rides on earth. Cellulose is composed of only glucose, whereas main presence of numerous xylanolytic bacteria and yeasts is reported from
J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850 843

the gut of M. darwiniensis, Zootermopsis angusticollis (Termopsidae), and fungus-growing termites often contains abundant xylanase activity and
Neotermes castaneus (Slaytor, 2000). Purification and molecular cloning significant differences in these distribution patterns occur among species
of major xylanases from C. formosanus revealed that these xylanases (Rouland-Lefèvre, 2000). There are two hypotheses to explain the origin
were affiliated with GHF11 and their origin was attributed to one of of midgut xylanase activity: 1) the enzyme is acquired from symbiotic
three intestinal protistan species, Holomastigotoides mirabile (Table 1) fungi, 2) the enzyme is endogenous (Rouland-Lefèvre, 2000; Slaytor,
(Arakawa et al., 2009). Recent transcriptomic analyses indicated the ab- 2000). Accordingly, some xylanases with different molecular weight
sence of any hemicellulases in host tissues of H. sjostedti (Yuki et al., (36, 56, 22.5 kDa) have been purified from the fungus-growing termite
2008), R. flavipes (Tartar et al., 2009), and C. formosanus (Zhang et al., Macrotermes bellicosus and its symbiotic fungus (Table 1) (Matoub and
2012b) except for GHF35 β-galactosidases in C. formosanus and a GHF2 Rouland, 1995) a xylanase with big molecular weight over 80 kDa was
β-mannosidase from H. sjostedti. In contrast, meta-transcriptomic analy- also purified from symbiotic fungus of M. subhyalinus (Faulet et al.,
ses of hindgut prokaryotic communities in lower termites provided a 2006). A clear picture of origin of the midgut xylanases has yet to be elu-
number of hemicellulolytic enzymes including GHF10 and GHF11 cidated, but a recent study indicated the presence of a GHF11 xylanase in
xylanases (Table 1) (Tartar et al., 2009; Todaka et al., 2007, 2010a; the gut microbiota of the fungus-growing termite, M. annandalei
Cairo et al., 2011; Xie et al., 2012). Todaka et al. (2010a) suggest that (Table 1) (Liu et al., 2011). GH11 xylanase was also existed in the gut
GHF10 xylanases are members of the core enzyme set and GHF11 microbiota of higher termite Nasutitermitidae (Brennan et al., 2004).
xylanases assist these core enzymes. Primitive termites M. darwiniensis, According to recent metagenomic analyses of the hindgut microbiota
N. koshunensis as well as the wood-roach C. punctulatus lack GHF11 in wood-feeding higher termites Nasutitermes sp. and Microcerotermes
xylanases. However, GH10 xylanase is much less abundant in sp., these metagenomic libraries seem to contain hemicellulolytic enzyme
Rhinotermitidae, the most apical lineage of lower termites, such as R. genes more abundantly than cellulase genes (Nimchua et al., 2012;
speratus, R. flavipes, and C. formosanus (Tartar et al., 2009; Todaka et al., Warnecke et al., 2007). In these termites, GHF8, GHF10, GHF11, and
2010a; Xie et al., 2012) and GHF11 might play a central role in xylan GHF43 xylanases were obtained from the metagenomic libraries of the
lysis in these termites. This is consistent with reports suggesting that hindgut microflora, and the presence of an apparent xylanosome operon
GHF11 xylanase purified from C. formosanus accounted for 87% of total was suggested in Microcerotermes sp. (Nimchua et al., 2012). On the
xylanase activity (Arakawa et al., 2009). In addition, GHF11 xylanase other hand, GHF10 xylanases were detected with proteomic analysis
and GHF1 β-xylosidase/β-glucosidase were reported from Reticulitermes of Nasutitermes sp. GHF8 was also detected, but substrate specificity
santoensis gut bacterial isolates (Matteotti et al., 2011, 2012). Contribu- (xylanase or cellulose) has not been defined at this time (Burnum et
tion of bacterial xylanases to hemicellulose degradation in lower termites al., 2011). A thermostable microbial GHF10 xylanase was found in
has yet to be verified. bacteria Saccharopolyspora pathumthaniensis S582 isolated from a grass-
feeding termite of another subfamily Apicotermitinae, Speculitermes sp.
4.2. Xylanases from higher termites (Sinma et al., 2011). An EST analysis of the fungal symbiont Termitomyces
sp. of the fungus-growing termite M. gilvus indicated that the fungus also
Except for the possible presence of bifunctional cellulase/xylanase in produces hemicellulolytic enzymes including xylanases (Johjima et al.,
the midgut of Trinervitermes trinervoides (subfamily Nasutitermitinae) 2006). Thus, symbiotic microbiota in higher termites has a potential
(Potts and Hewitt, 1974), there are no reports of endogenous xylanases role in degrading hemicelluloses and hence may be rich reservoirs of
from litter- or wood-feeding higher termites. However, the midgut of hemicellulolytic genes.

Table 1
Xylanases purified or sequenced from termites or symbiotic microorganisms.

Enzyme Origin MW (kDa) Glycoside hydrolase family Reference

Endoxylanase Purified from symbiotic fungus of M. bellicosus 36 n.d. Matoub and Rouland (1995)
Exoxylanase Purified from M. bellicosus 56 n.d. Matoub and Rouland (1995)
Exoxylanase Purified from symbiotic fungus of M. bellicosus 22.5 n.d. Matoub and Rouland (1995)
Xylanase From termite Nasutitermes sp. gut library (XYL6419) n.d. GH11 Brennan et al. (2004)
Xylanase Purified from symbiotic fungus of M. subhyalinus 80–87 n.d. Faulet et al. (2006)
Endo-1,4-beta-xylanase Purified from symbiotic flagellatesof Coptotermes formosanus 19, 17, 18 GH11 Arakawa et al. (2009)
Endo-1,4-beta-xylanase Cloned from a gut microbiota fosmid library of M. annandalei 64.5 GH11 Liu et al. (2011)
Xylanase Purified from bacteria from higher termite 36 GH10 Sinma et al. (2011)
Xylanase Purified from bacteria from higher termite 205 GH10 Dheeran et al. (2012)
Endo-1,4-beta-xylanase Cloned from genomic library of bacteria from R. santoensis 29 GH11 Matteotti et al. (2012)
Xylanase Cloned from hindgut fosmid library from Microcerotermes sp. n.d. Nimchua et al. (2012)
GH8, 10, 11

Endo-1,4-beta-xylanase From protist of C.formosanus through 454 pyrosequencing 34 GH10 Xie et al. (2012)
Endo-1,4-beta-xylanase From EST analysis of symbiotic fungus of fungus-growing termite n.d. GH10, 11 Johjima et al. (2006)
*Xylanase From hindgut microbiota metagenomic analysis of Nasutitermes n.d. GH8, 10, 11, 43 Warnecke et al. (2007)
*Xylanase From cDNA library of protist of R. speratus n.d. GH10, 11, 43 Todaka et al. (2007)
*Xylanase From EST analysis of several lower termites n.d. GH8, 10, 11, 43 Todaka et al. (2010)
*Xylanase From metatranscriptomic analysis of R. flavipes n.d. GH10, 11 Tartar et al. (2009)
*Xylanase From metaproteomic analysis of C. gestroi n.d. GH10, 11, 43 Cairo et al. (2011)
*Xylanase From transcriptome analysis of C. formosanus n.d. GH8, 10, 11 Zhang et al. (2012)
n.d.: not determined.
a
Obtained from “omics” study; xylanase activity has not been confirmed.
Xylanases found from lower termites are highlighted in gray.
844 J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850

5. Lignin and its modification enzymes incapable of mineralizing larger lignin moieties. Recently, investigations
into the structural details of lignin changes during passage through the
5.1. Structural changes of lignin during passage of the gut termite gut have been initiated and are described below.
A current scheme of the termite lignolytic system is illustrated in
Lignin has a complex amorphous and recalcitrant structure, Fig. 4. In a series of recent publications, Ke and colleagues analyzed
consisting of p-hydroxyphenylpropane monomer units joined by structural changes of lignin in the gut of C. formosanus via a combination
ether or C\C bonds (Cornwell et al., 2009). Lignins comprise 20–30% of spectroscopic techniques and NMR, with a special emphasis on pyrol-
of wood's dry weight, and gymnosperms primarily contain guaiacyl ysis gas chromatography–mass spectrometry analysis. Because pyroly-
lignin, whereas angiosperms mainly contain syringyl lignin. Here we sis is conducted at 210–610 °C, the structure of large lignin molecules
summarize the structural changes of lignin which have been well is difficult to differentiate from resulted pyrolytic fragments. Neverthe-
documented within the Coptotermes species. Early studies on lignin less, such data are useful for ascertaining structural modifications of lig-
degradation in termites have been elegantly described in a review by nin based on pyrolytic patterns and semi-quantitative, high-resolution
Breznak and Brune (1994). In such studies, 14C-[lignin]-lignocellulose structures. With these methods, initial structural changes already
was administrated to Coptotermes acinaciformis but no significant occurred at the termite mastication step (Ke et al., 2012a). Further mod-
14
release of respired CO2 was detected. A preliminary study with C. ifications appeared to continuously occur in the termite foregut and the
formosanus by Kyou et al. (1996) suggested that milled-wood-lignin midgut (Ke et al., 2010). These modifications were characterized by lig-
was specifically endocytosed by one of three symbiotic protists, nin ring decarboxylation and side chain oxidation. Cleavage of some
Holomastigotoides hartmanni and their microscopic observation aromatic rings is presumed to occur in the hindgut based on the pyro-
suggested that its surface was likely degraded (Kyou et al., 1996). lytic analysis of short aromatic model compounds fed to termites (Ke
Gel permeation chromatography of termite fecal lignin indicated a et al., 2011b). No potential lignin degradation enzymes have been
modification of the lignin–carbohydrate complex, but side chain and found from metatranscriptomes of the symbiotic protistan community
aromatic ring cleavages of ingested lignin molecules were not detected in the hindgut of C. formosanus (Xie et al., 2012), whereas in general, a
with infrared and UV spectrometry. A simple β-O-4′ type lignin role has been suggested for bacteria in lignin degradation (Bugg et al.,
model compound (1-(4-O-benzylguaiacyl)-3-hydroxy-2-(4-methyl- 2011). Apparently, termites harbor intestinal bacteria with aromatic
umberlliferyl)-propanone) was fed to C. formosanus for one week and degrading capabilities (Bugg et al., 2011; Le Roes-Hill et al., 2011). Clas-
4-methyl-umbelliferone measurements indicated that ~45% of the sification of these bacteria was coincident with classes of previously
compound was degraded by the termites (Hirai et al., 2000). In contrast, identified bacteria with in vitro lignin-degradation abilities. The meta-
degradation of oligomers (3- to 5-mer) synthesized from coniferyl alco- bolic pathway of bacterial lignin degradation is still unclear, but reports
hol, which was fed to C. formosanus for 20 days, was not detected with of bacteria that produce peroxidases suggest that such bacteria may be
HPLC (Hirai et al., 2000), although possible oxidations of the lignin olig- involved in these lignin modifications (Bugg et al., 2011; Le Roes-Hill et
omers in another termite H. sjostedti were detected by the same exper- al., 2011). Analysis of termite feces revealed modifications on lignin
iment. Moreover, a comparative analysis of lignin between a dietary alkyl side groups, ring hydroxyl groups, and ring methoxyl groups
wood pieces and feces of C. formosanus using solid-state CP/MAS 13C with conservation of the abundant β-O-4′ interunit lignin linkage and
NMR revealed that the host and the gut symbionts had little or no ability retention of the original aromatic properties (Ke et al., 2011a,c).
to degrade natural lignin: no alterations in aromatic ring carbons and aryl Therefore, it is conceivable that the termite digestive system does not
methoxyl carbons were noted although wood polysaccharides were reduce aromatic moieties in large lignin molecules but only modifies
significantly degraded (Hyodo et al., 1999). Termites are thought to be some functional groups (Ke et al., 2011c). Indeed, other research
capable of degrading phenylpropanoid lignin monomers or dimers but indicates that total lignin increased from 28.9% in wood to 59.2% in

Lignocellulose
(Lignin + cellulose + hemicellulose)

Laccase Lignin ring decarboxylation and side


Aldo-keto reductases chain oxidation etc., at the chewing step,
Catalases and in the foregut and the midgut

Lignocellulose with reduced glucan


content: conservation of the abundant
Exposure of cellulose (and β-O-4’ interunit lignin linkage and
hemicellulose) fibers retention of the original aromatic
properties

Some presumed aromatic ring


Cellulases and hemicellulases cleavage by the hindgut bacteria
(no enzymes involved in this
process have been isolated)

Feces
(Total lignin proportion increased without reducing aromatic
moieties)

Fig. 4. A scheme of lignin modification in the gut of termites. Modifications of lignin and hemicellulose at the chewing steps and in the gut are thought to result in unlocking of
cellulose fibers from lignocellulose and enabling efficient cellulose digestion. However, lignin is thought to retain the original aromatic properties throughout the gut passage.
J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850 845

feces of C. formosanus with observations of similar chemical changes of (Sethi et al., 2013). However, such collaborative actions have yet to be
lignin (Hyodo et al., 1999). Glucan content was significantly reduced established to occur in vivo.
during gut passage (Li et al., 2012). Such lignin modifications likely Details of possible lignin modifications have yet to be elucidated in
facilitate the enzymatic hydrolysis of wood ingested by termites (Li et higher termites. No genes encoding lignolytic enzymes were identified
al., 2012), but genes encoding lignolytic enzymes or pure enzymes in the metagenomic library from the hindgut microflora of wood-
have not been isolated from termite gut bacteria. feeding Nasutitermes (Warnecke et al., 2007). Still, some knowledge ex-
ists regarding the putative lignocellulytic enzymes from fungus-growing
5.2. Lignin-modification enzymes termites, in which lignin degradation is thought to be accomplished
through symbiosis with the white-rot fungi, Termitomyces sp. Little effect
Lignin mineralization is extensively studied in white-rot and of delignification without the aid of Termitomyces was suggested in
brown-rot fungi and the following enzymes are known to participate in O. formosanus (Li et al., 2012). Predominant laccase activities against
lignin degradation: lignin peroxidase (LiP) (EC 1.11.1.14), manganese DMP and ABTS were detected in the fungus combs and fungi isolated
peroxidase (MnP) (EC 1.11.1.13), versatile peroxidase (VP) (EC from the nests of three genera of fungus-growing termites (i.e.
1.11.1.16), and laccase (EC 1.10.3.2) (Bugg et al., 2011). Lignin is Macrotermes, Odontotermes, and Microtermes) (Taprab et al., 2005). The
depolymerized and mineralized by radical reactions that are triggered relevant genes involved in plant cell wall degradation were identified
by oxidation with these enzymes. In recent years, cellobiose dehydroge- from the symbiotic fungi of fungus-growing termites (Johjima et al.,
nase (CDH) (EC 1.1.99.18) has been shown to participate in the 2006). A small amount of laccase activity was also detected in the gut
lignocellulolytic process, possibly via the production of free radicals of O. formosanus (the laccase activities were in this order: symbiotic
(Harreither et al., 2009). On the other hand, LiP, MnP, VP, and CDH have fungus > fungus comb > worker alimentary tracts) (Pan et al., 2009).
not been identified in insects, whereas laccases are commonly involved The laccase purified from the fungus combs of O. formosanus has been
in melanization and sclerotization of the cuticular layers in insects. characterized (Zhou et al., 2010b): the molecular mass of the laccase
Recently, laccases have been documented as candidates for lignin- was 65 kDa, and the optimum pH and temperature were pH 4.0 and
modification enzymes (“lignases”) in termites. Although “lignases” gen- 10 °C with ABTS as the substrate, respectively (Vmax and Km were
erally represent enzymes that efficiently degrade or mineralize lignin, 3.62 μmol min−1 mg−1 and 119.52 μM, respectively). These character-
this does not appear to be true in termites, according to the aforemen- istics are not significantly different from other fungal laccases (Zhou
tioned studies. From R. flavipes, cDNAs encoding putative laccases et al., 2010b). With these fungi, lignin is progressively degraded in
were found in a host EST library (Scharf and Tartar, 2008; Tartar et al., the combs and degrees of such degradation differ among various
2009), while strong oxidative activity against pyrogallol, a monomeric Termitomyces species or strains (Hyodo et al., 2003). In contrast,
lignin model compound, was detected in the salivary glands and foregut the gut pH of Macrotermitinae ranges from 5.0 to 8.5 (Bignell and
extract of this termite (Tartar et al., 2009). Further studies on these Eggleton, 1995), which may not be a suitable physical condition for
genes indicated that these laccases were primarily produced in the the fungal laccases and may account for the small amount of activity.
salivary glands and that recombinant laccases, which were expressed
in Trichoplusia ni larva using a Baculovirus-expression vector, showed 6. Heterologous production of lignocellulases from termites
trace or no activity against widely-used laccase substrates such as
2,2′-azino-bis(3-ethylbenzothiazoline-6-sulphonic acid) (ABTS) and For industrial applications of termite lignocellulolytic enzymes, re-
syringaldazine as well as melanin precursors but preferentially oxidized combinant enzyme productions (overexpressions) in heterologous
lignin monomers such as 2,6-dimethylphenol (DMP) and pyrogallol in hosts are useful and necessary. In the last several years, increasing in-
the presence of H2O2 (Coy et al., 2010). These enzymes also act against vestigations have been reported regarding the heterologous expres-
soluble lignin derivatives (Coy et al., 2010), but a detailed mode of ac- sions of lignocellulolytic enzymes obtained from termites and their
tion is unclear. Addition of the salivary laccase reduces glucose produc- symbiotic microbiota. However, an expression of CBH has yet to be
tion from an enzymatic reaction by endogenous EG and BG of R. flavipes achieved; although, this is apparently an essential enzyme for crystal-
when pine wood sawdust is used as a substrate, whereas it enhances line cellulose digestion. Next, we summarize what is known about re-
the activity of a protistan GHF11 xylanase when using beechwood combinant enzymes and compare enzymatic properties of EGs, BGs,
xylan as a substrate (which contains lignin) (Scharf et al., 2011; Sethi and xylanases from termites or their symbionts.
et al., 2013). Of note, the simultaneous actions of endogenous (salivary
glands and/or midgut) and symbiotic cellulases (hindgut) are unlikely 6.1. Heterologous production of EGs
in vivo (Tokuda et al., 2007). Previous studies suggested that lignin
modifications similar to those in the gut of C. formosanus were observed Comparisons of enzymatic properties of recombinant EGs from ter-
in the gut of R. flavipes (Ke et al., 2010) but degradation activity against mites are shown in Table 2. Recombinant production (overexpression)
the β-O-4′ type lignin model compound in the Rhinotermitid R. speratus of termite GHF9 enzymes was previously difficult, although the function-
was reduced, compared to activity observed C. formosanus (~13%) al expression of NtEG has been detected using the Congo-red plate assay
(Hirai et al., 2000). Moreover, an early study using R. flavipes also (Tokuda et al, 1999). Successful overexpression of termite GHF9 EG was
indicated no significant change of lignin content in wood particles first reported in 2005 with Escherichia coli (Ni et al., 2005). The
during passage through the gut, but suggested the possibility of overexpressed mutant EG was obtained by family shuffling of EG
chemical lignin modifications (Breznak and Brune, 1994). Therefore, genes from four different termite species (R. speratus, N. takasagoensis,
putative laccases in R. flavipes might be unable to degrade large C. formosanus, and C. acinaciformis). The mutant enzyme had 20–30-fold
insoluble lignin molecules and contribute to functional group higher activity against CMC and had properties similar to those of the na-
modifications of lignin or detoxification of lignin monomers that tive enzymes (Ni et al., 2005). Using the adapted mutant cDNAs as paren-
have antimicrobial activity (Borneman et al., 1986). These activities tal genes combined with native-form cDNAs, further family shuffling was
may also be related to a potential ability of termites to degrade performed and a mutant EG with improved thermostability was
toxic polycyclic aromatic hydrocarbons, a finding that was recently obtained. This mutant EG had high specific activity (889 units/mg) and
been described (Ke et al., 2012b). In addition, aldo-ketoreductases and the thermostability was 10 °C higher than parental enzymes (Ni et al.,
catalases have been shown to enhance lignocellulose degradation in 2007a). Furthermore, activity related amino acid mutations of a GHF9
vitro (Sethi et al., 2013). These two enzymes alone, or together can signif- EG were identified (Ni et al., 2010).
icantly enhance glucose liberation from pine-wood lignocelluloses by Currently, a number of termite EGs have been reported to be
combining to termite endogenous GHF9, GHF1, and protistan GHF7 expressed in different hosts. EG (CfEG3a) of C. formosanus, in both
846 J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850

Table 2
Enzymatic properties of purified recombinant EGs from termites and their symbionts.

Source organism Specific activity Optimal temperature Optimal pH Km Vmax Expression host Reference
(U/mg) (°C) (CMC) (CMC)
(mg/ml) (U/mg)

Chimeric termite EGs


A18 544 55 7 10.7 1074 E.coli Ni et al. (2005)
PA68 560 55 6.9 12.7 889 E.coli Ni et al. (2007a)

Reticulitermes speratus
RsEG (GHF9) 1200 45 5.5 2 1429 A.orizae Hirayama et al. (2010)

Nasutitermes takasagoensis
NtEG (GHF9) 1392 65 6 4.67 1667 A.orizae Hirayama et al. (2010)

R.flavipes
Cell-1 (GHF9) n.d. n.d. n.d. 14.66 0.838 E.coli Zhou et al. (2010a)
Cell-1 (GHF9) 1.3 50–60 6.5–7.5 9.93 1.056 T.ni⁎ Zhou et al. (2010a)

Coptotermes formosanus
CfEG3a (His-tagged) (GHF9) 328 37 5 2.2. 590 E.coli Zhang et al. (2009, 2011)
CfEG5 (GHF9) 325 43 5.6 5.6 548 E.coli Zhang et al. (2011)

Protists from C.formosanus


CFP-EG1 (GHF5) 105 70 6 1.9 148.2 E.coli Inoue et al. (2005)

Protists from R.speratus


RsSymEG1 (GHF7) 603 45 6.5 1.97 769.6 A.orizae Todaka et al. (2010b)

One unit (U) is defined as the amount of enzyme that releases 1 μmol of reducing sugar (glucose equivalent) per minute. Activities against carboxymethylcellulose (CMC) are
shown for comparisons. n.d, not determined. In addition to above-mentioned cellulases, 26 GHF5, 7 GHF9, and 2 GH45 bacterial EG genes from the metagenomic library in the hind-
gut of Nasutitermes sp. and one GHF5 EG from the metagenomic library in the hindgut of Microcerotermes sp. were expressed in E.coli and P.pastoris and their activity was confirmed
(Nimchua et al., 2012; Warnecke et al., 2007).
⁎ Baculovirus was used as an expression vector.

native form (nCfEG) and C-terminal His-tagged form (tCfEG) was Km and Vmax for CMC were 1.97 mg/ml and 769.6 units/mg protein,
expressed in E. coli using pET28 vector (Zhang et al., 2009). The re- respectively.
combinant nCfEG and tCfEG showed cellulolytic activity against Recombinant EGs are useful for studying cellulose metabolism in ter-
CMC with specific activities of 328 and 385 units/mg respectively. mites and have the potential for being engineered as new biocatalysts for
Another EG (CfEG5) from C. formosanus was also expressed in E. coli cellulose-based biofuel production. The recombinant cellulase should
with a specific activity of 325 units/mg (Zhang et al., 2011). The re- also be useful for designing and screening inhibitors to develop target-
combinant CfEG5 was active against filter-paper cellulose, resulting specific and environment-friendly bio-termiticides (Zhang et al., 2011;
in mostly cellobiose and cellotriose. A native form of EGs from R. Zhou et al., 2008a,b).
speratus and N. takasagoensis (RsEG and NtEG) was also produced in
the filamentous fungus Aspergillus oryzae, which is a more suitable
host for mass production than E. coli (Hirayama et al., 2010). The 6.2. Heterologous production of BGs
activities of purified recombinant EGs against CMC were significantly
higher than that of any other EGs reported (1200 and 1392 units/mg BG is critical for cellulose degradation and glucose production in the
for RsEG and NtEG, respectively). These specific activities of recombi- termite. GHF1 BGs from termites and their symbionts have been success-
nant endogenous termite EGs were approximately 100 times higher fully overexpressed in prokaryotic and eukaryotic hosts (Table 3). NkBG,
than commercially available Trichoderma EGs (Hirayama et al., obtained from the salivary glands of N. koshunensis (Tokuda et al., 2002),
2010). These EGs were also expressed in the yeast Pichia pastoris was first heterologously expressed in E. coli (Ni et al., 2007b). The enzy-
using vector pBGP3, which allowed one-step purification of heterolo- matic properties of the recombinant BG are mostly consistent with the
gous proteins with N-terminal 6 × His and myc-epitope tags (Uchima native enzyme. Km and Vmax against cellobiose were 3.8 mM and
and Arioka, 2012). An endogenous salivary EG from R. flavipes was 220 units/mg, respectively. The optimum pH and thermostability were
also expressed using both Baculovirus and E. coli expression systems 5.0 and 45 °C respectively. However, the specific activity of the recombi-
to compare their properties (Zhou et al., 2010a). As a result, the nant enzyme was 156.7 U/mg, which is almost 3-fold higher than that of
Baculovirus-mediated expression of EG was more readily obtained in the partially purified BG of N. koshunensis (Tokuda et al., 2002), and it
solubilized form and was more active than the enzyme expressed in preferentially hydrolyzed laminaribiose (consisting of β-1,3-linked glu-
E. coli. Recombinant Baculovirus/T. ni-expressed enzyme activity was cose) over cellobiose (consisting of β-1,4-linked glucose). To construct
comparable to the native enzyme, and the optimal activity was a more efficient mass production system, NkBG was expressed in A.
obtained at pH 6.5–7.5 and 50–60 °C. In addition, enhanced activity oryzae (Uchima et al., 2011). Specificity against p-nitrophenyl-β-D-
was documented up to 70 °C in the presence of CaCl. Km and Vmax glucoside (pNPG) was 12.4 U/mg. Km and Vmax against pNPG were
for the Baculovirus-mediated expression of EG against CMC were 0.77 mM and 16 units/mg, respectively. The other enzyme proper-
0.993% w/v and 1.056 μmol/min/mg, respectively. ties were similar to the enzyme expressed in E. coli. Interestingly,
As for protistan EGs, a GHF5 EG gene was overexpressed in E.coli, glucose stimulated NkBG activity; in contrast, other BGs are often
and its optimal pH and temperature were 5.8–6.0 and 70 °C, respective- inhibited by glucose, the end product of enzymatic hydrolysis. For
ly (Inoue et al., 2005). Km and Vmax for CMC were 1.90 mg/ml and example, activity of the commercially available BG, Novozyme188,
148.2 units/mg protein, respectively. Also, a GHF7 EG from a symbiotic is inhibited by 0.6 M glucose (Uchima et al., 2012), but NkBG activity
protist of R. speratus was expressed in A. oryzae (Todaka et al., 2010b). was 100% at this glucose concentration, compared to activity measured
The recombinant EG had relatively high specificity (603 units/mg) without glucose. Moreover, incubation of NkBG with 200 mM glucose
and its optimal pH and temperature were 6.5 and 45 °C, respectively. delayed enzyme inactivation using heat (55 °C) (Uchima et al., 2011).
J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850 847

Table 3
Enzymatic properties of purified recombinant BGs (GHF1) from termites and their symbionts.

Source organism Specific Optimal Optimal Km Vmax Relative act. (%) at Expression Reference
activity temperature pH (mM) (U/mg) glucose conc. 0.6 Ma host
(U/mg) (°C)

Neotermes koshunensis
NkBG (cellobiose) 156. 7 45 5 3.8 220 n.d. E.coli Ni et al. (2007b)
NkBG (pNPG) 12.4 40 5 0.77 16 110 A.oryzae Uchima et al. (2011)

Reticulitermes favipes
RfBGluc-1 (cellobiose) n.d. 40 6.5–7 1.44 638 n.d. T.nib Scharf et al. (2010)
RfBGluc-1 (pNPG) n.d. 40 6.5–7 1.66 22.92 n.d. T.nib Scharf et al. (2010)

Coptotermes formosanus
β-glucosidase (cellobiose) n.d. 49 5.6–6.2 2.3 462.6 85 E.coli Zhang et al. (2012a)

Macrotermes barneyi
MbmgBG1 206 (cellobiose) 45 (pNPG) 5 (pNPG) n.d. n.d. n.d. E.coli Wu et al. (2012)

Nasutitermes takasagoensis
G1sgNtBG1 (pNPG) n.d. 50 6 n.d. n.d. 70 P.pastris Uchima et al. (2013)
G1mgNtBG1 (pNPG) 5.83 65 5.5 0.67 8 50 P.pastris Uchima et al. (2012)

Bacerial isolate from R.santoensis


BglB (pNPG) 0.441 40 6 n.d. n.d. n.d. E.coli Matteotti et al. (2011)
BglB (pNPX) 2.112 n.d n.d n.d. n.d. n.d. E.coli Matteotti et al. (2011)

Symbiont of Globitermes sulphureus


bgl-gs1 (pNPG) 110 90 6 0.18 61.6 n.d. E.coli Wang et al. (2012)

One unit (U) is defined as the amount of enzyme that releases 1 μmol of glucose (from cellobiose as substrate) or p-nitrophenol (from p-nitrophenyl-β-D-glucopyranoside; pNPG or
p-nitrophenyl-β-D-xylopyranoside; pNPX as substrate) per minute. n.d, not determined.
a
Activity of Novozyme 188, a commercially β-glucosidase, was found to be completely inhibited in the presence of 0.6 M glucose (Uchima et al., 2012).
b
Baculovirus was used as an expression vector.

GHF1 BG genes from other lower termites, R. flavipes and 2012). This BG (bgl-gs1) was overexpressed in E. coli and recombinant
C. formosanus, have also been heterologously expressed. RfBGluc-1 BG had an optimal temperature with pNPG at 90 °C which retained
was produced in T. ni with a baculovirus-insect expression system more than 60% of its activity for 120 min at 75 °C. The specific activities
(Scharf et al., 2010). Km and Vmax against pNPG were 1.66 mM and of bgl-gs1 on pNPG and salicin were 110 units/mg and 14 units/mg;
22.92 μmol/min/mg, respectively, while Km and Vmax against cellobiose Km and Vmax values were 0.18 mM and 61.6 units/mg for pNPG and
were 1.44 mM and 638 μmol/min/mg, respectively. The temperature 2.59 mM and 71.9 units/mg, respectively.
stability of the recombinant RfBGluc-1 increased about 20% in the pres-
ence of 30 mM CaCl2 (Scharf et al., 2010). Recombinant RfBGluc-1, 6.3. Heterologous production of xylanases
when combined with a host GHF9 EG, had 100–300-fold higher glucose
production from pine sawdust, compared to EG or BG alone (Scharf et Recombinant xylanases and those purified from termite gut bacteria
al., 2011). The rate of glucose production was further enhanced by an are listed in Table 4. Recently, a xylanase (XylB8) from Gram-positive
addition of a protistan GHF7 (Sethi et al., 2013). bacteria of Reticulitermes santonensis was cloned by functional activity
BG from C. formosanus was functionally expressed in E. coli (Zhang screening (Matteotti et al., 2012). XylB8 was affiliated with GHF11
et al., 2012a). Km and Vmax against cellobiose were 2.33 mM and and expressed in E. coli. Recombinant xylanase had maximal activity
462.6 μmol/min/mg, respectively. Glucose had no significant effect against beechwood xylan at a pH of 5.0 and 55 °C. Surprisingly, this
on BG activity when the glucose concentration was increased to up xylanase retained 28% activity after preincubation at 100 °C. Km and
to 0.3 M. An approximately 6% reduction occurred when glucose Vmax were 9 mg/ml and 3333 units/mg, respectively. Compared to the
reached 0.4 M. Nevertheless, 71% activity remained when 1.0 M glucose activity against beechwood xylan, it also hydrolyzed birchwood xylan
was present in the reaction mixture. This BG was irreversibly inactivated (96%) and weakly degraded CMC (6%).
by an inhibitor of conduritol B epoxide, a finding which might be useful A GHF11 xylanase gene was also obtained from a fosmid library of the
in the development of enzyme-specific bio-termiticides. gut symbionts of the fungus-growing termite M. annandalei (Liu et al.,
G1NtsgBG1 and G1NtmgBG1, obtained from the salivary glands and 2011). This xylanase (Xyl6E7) was expressed in E. coli. Activity against
the midgut of the higher termite N. takasagoensis (Tokuda et al., 2009), beechwood xylan was 771 units/mg; activity against birchwood xylan
were expressed in the yeast Pichia pastoris (Uchima et al., in press). In con- was 733 units/mg. The optimal pH was 7.5 and the activity was temper-
trast to monomeric forms of BGs in lower termites, the purified recombi- ature sensitive. Maximal activity was observed at 55 °C, a preincubation
nant G1NtsgBG1 and G1NtmgBG1 were assumed to be homotrimeric at 30 °C to 55 °C for 5 min reduced activity to less than 50%. However,
with molecular masses of 163 kDa and 169.5 kDa, respectively. These L-cysteine restored activity to more than 60% after preincubation at
BGs were relatively thermostable and highly glucose-tolerant. Further- 55 °C for 30 min.
more, G1NtmgBG1 acted synergistically with Celluclast 1.5 L, releasing In addition, xylanases (GHF10 and GHF11) from the metagenomic
more reducing sugars from Avicel than Novozym 188 combined with library of Microcerotermes sp. were partially characterized (Nimchua
Celluclast 1.5 L, suggesting that this termite BG holds promise for supple- et al., 2012). These xylanases performed optimally at pH 8.0 and
mental use in cellulose hydrolysis (Uchima et al., 2012). Recently, 55 °C. They were stable at alkaline pHs and had greater than 70% ac-
MbmgBG1 obtained from the midgut of the fungus-growing termite M. tivity at pH 9.0.
barneyi was functionally expressed in E. coli with relatively high specific- Recently, reports of xylanases purified from termite gut bacteria sug-
ity (206 units/mg against cellobiose) (Wu et al., 2012). Apart from endog- gest that because the bacteria are culturable, these xylanases could be
enous BGs from termites, a highly thermostable GHF1 BG has been attractive for industrial applications. A thermostable xylanase was iden-
identified from a metagenomic library of the gut of the higher termite tified from a gut bacterium (S. pathumthaniensis S582) of Speculitermes
Globitermes sulphureus (Amitermes-group Termitinae) (Wang et al., sp. (Sinma et al., 2011). The optimal temperature and pH of the xylanase
848 J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850

Table 4
Enzymatic properties of recombinant xylanases and purified xylanase from termite gut bacteria.

Source organism Substrate Specific activity Optimal temperature Optimal pH Km Vmax Expression host Reference
(U/mg) (°C) (mg/ml) (U/mg)

Gram-positive bacteria from Reticulitermes santoensis


mXylB8 (GHF11) Beechwood xylan 1837 55 5 9 3333 E.coli Matteotti et al. (2012)
Gut symbionts from Macrotermes annandalei
Xyl6E7 (GHF11) Beechwood xylan 771 n.d. n.d. n.d. n.d. E.coli Liu et al. (2011)
Birchwood xylan 733 50–55 7.5 6.96 1057.8 E.coli Liu et al. (2011)
Gut symbionts from Microcerotermes sp.
X0012 (GHF11) Birchwood xylan n.d. 55 8 n.d. n.d. E.coli Nimchua et al. (2012)
X1098 (GHF10) Birchwood xylan n.d. 55 8 n.d. n.d. E.coli Nimchua et al. (2012)

Purified bacterial xylanase


Saccharopolyspora pathumthaniensis S582 isolated from Speculitermes sp.
Extracellular xylanase (GHF10) Beechwood xylan 965 70 6.5 3.92 256 Non-recombinant Sinma et al. (2011)
Paenibacillus macerans IIPSP3 isolated from wood-feeding higher termites
Xylanase (possibly GHF10) Beechwood xylan 4170 60 4.5 6 7407 Non-recombinant Dheeran et al. (2012)
Birchwood xylan n.d. n.d. n.d. 6.1 7575 Non-recombinant Dheeran et al. (2012)

One unit (U) is defined as the amount of enzyme that releases 1 μmol of reducing sugar (xylose equivalent) per minute; n.d., not determined.

were 70 °C and 6.5. Km and Vmax were 3.92 mg/ml and 256 units/mg. Acknowledgments
The ORF of this xylanase encoded 368 amino acid residues belonging
to GHF10. We thank Prof. Y. Shen at Shandong University for reading the man-
Another thermostable xylanase was purified from a gut bacterium uscript, and for his helpful comments. J.N is grateful to Prof. B.G. Kim
(Paenibacillus macerans IIPSP3) of wood-feeding higher termites (species and the Korea Foundation for Advanced Studies for support to conduct
not identified) (Dheeran et al., 2012). The optimal temperature was research at Seoul National University. This work was supported by
60 °C, but it was active over a broad range of temperatures (40–90 °C) grants from the National Basic Research Program of China (973 pro-
and pHs (3.5–9.5). The half-life of xylanase was 6 h at 60 °C and 2 h at gram: 2011CB707402) and the National Natural Science Foundation of
90 °C. This xylanase has a molecular mass of 205 kDa, and the internal China (31272370, 30870085). Part of this work was supported by the
amino acid sequence was not significantly homologous with known Program for Promotion of Basic and Applied Researches for Innovations
xylanases. A typical feature of this xylanase was not only thermostability in Bio-oriented Industry (PROBRAIN) and by a grant for the 21st Centu-
but its high specific activity (Vmax = 7407 units/mg). ry COE program of University of the Ryukyus and Grants-in-aid for Sci-
Although uncharacterized at present, the protistan GHF11 xylanase entific Research No. 20380037, 17405025, 16780037, and 13760043
gene was expressed in T. ni larva using an insect Baculovirus-expression from the Japan Society for the Promotion of Science.
vector system (Sethi et al., 2013), and significantly more xylose was lib-
erated from pine lignocellulose when the enzyme was combined with References
host laccase than when the xylanase was tested alone, indicating that
the combination of hemicellulose degradation and lignin modification Adav SS, Ravindran A, Sze SK. Quantitative proteomic analysis of lignocellulolytic en-
zymes by Phanerochaete chrysosporium on different lignocellulosic biomass. J Pro-
is inducible in vitro. teomics 2012;75:1493–504.
Ahmed S, Riaz S, Jamil A. Molecular cloning of fungal xylanases: an overview. Appl
Microbiol Biotechnol 2009;84:19–35.
Arakawa G, Watanabe H, Yamasaki H, Maekawa H, Tokuda G. Purification and molecu-
7. Future perspectives lar cloning of xylanases from the wood-feeding termite, Coptotermes formosanus
Shiraki. Biosci Biotechnol Biochem 2009;73:710–8.
Termites can digest 74–99% lignocelluloses (Prins and Kreulen, Bignell DE. Soil-feeding and gut morphology in higher termites. In: Hunt JH, Nalepa CA,
editors. Nourishment and Evolution in Insect Societies. Boulder, CO: Westview
1991). Hence, lignocellulose-degrading enzymes produced by ter-
Press; 1994. p. 131–58.
mites and their symbiotic microorganisms should permit efficient uti- Bignell DE. Morphology, physiology, biochemistry and functional design of the termite
lization of lignocellulosic materials. Multiple genes explored through gut: an evolutionary wonderland. In: Bignell DE, Roisin Y, Lo N, editors. Biology of
“omics” studies and successful recombinant techniques with such Termites: A Modern Synthesis. Dordrecht: Springer; 2011. p. 375–412.
Bignell DE, Eggleton P. On the elevated intestinal pH of higher termites (Isoptera:
enzyme genes have provided initial tools for further research. How- Termitidae). Insect Soc 1995;42:57–69.
ever, our knowledge is limited with respect to the potential for indus- Borneman WS, Akin DE, VanEseltine WP. Effect of phenolic monomers on ruminal bac-
trial applications of termite lignocellulolytic enzymes. Studies on teria. Appl Environ Microbiol 1986;52:1331–9.
Brennan Y, Callen WN, Christoffersen L, Dupree P, Goubet F, Healey S, et al. Unusual mi-
hemicellulose degradation are few and effects of lignin modifications crobial xylanases from insects guts. Appl Environ Microbiol 2004;70:3609–17.
on cellulose and hemicellulose digestion are not fully understood. Be- Breznak JA, Brune A. Role of microorganisms in the digestion of lignocellulose by ter-
cause little heterologous production of termite-derived enzymes has mites. Annu Rev Entomol 1994;39:453–87.
Bugg TDH, Ahmad M, Hardiman EM, Singh R. The emerging role for bacteria in lignin
been reported, further studies on more diverse enzyme overexpression degradation and bio-product formation. Curr Opin Biotechnol 2011;22:394–400.
techniques using different systems, especially overexpression of CBHs, Burnum KE, Callister SJ, Nicora CD, Purvine SO, Hugenholtz P, Warnecke F, et al. Prote-
are needed. Also, novel enzymes with high lignocellulose degradation ome insights into the symbiotic relationship between a captive colony of
Nasutitermes corniger and its hindgut microbiome. ISME J 2011;5:161–4.
activity may reside in termite genes, but the difficulty of identifying Cairo JPLF, Leonardo FC, Alvarez TM, Ribeiro D, Büchli F, Costa-Leonardo AM, et al.
functions of “hypothetical proteins” may have caused investigators to Functional characterization and target discovery of glycoside hydrolases from the
overlook new enzymes. Thus, a comprehensive quantitative proteomic digestome of the lower termite Coptotermes gestroi. Biotechnol Biofuels 2011;4:50.
Cornwell WK, Cornelissen JHC, Allison SD, Bauhus J, Eggleton P, Preson CM, et al. Plant
analyses combined with enzyme activity screenings are needed. Com-
traits and wood fates across the globe: rotted, burned, or consumed? Global Change
parative secretome analysis may also identify novel enzymes involved Biol 2009;15:2431–49.
in lignocellulosic biomass decomposition (Adav et al., 2012). Addition- Coy MR, Salem TZ, Denton JS, Kovaleva ES, Liu Z, Barber DS, et al. Phenol-oxidizing
ally, optimal enzyme “cocktails” for efficient lignocellulose degradation laccases from the termite gut. Insect Biochem Mol Biol 2010;40:723–32.
Deng TF, Chen CR, Cheng ML, Pan CY, Zhou Y, Mo JC. Differences in cellulase activity
(based on the termite enzymatic system) could provide novel ap- among different castes of Odontotermes formosanus (Isoptera: Termitidae) and
proaches for industrial applications. the symbiotic fungus Termitomyces albuminosus. Sociobiology 2008a;51:697–704.
J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850 849

Deng TF, Zhou Y, Cheng ML, Pan CY, Chen CR, Mo JC. Synergistic activities of the symbiotic Liu N, Yan X, Zhang ML, Xie L, Wang QA, Huang YP, et al. Microbiome of fungus-growing
fungus Termitomyces albuminosus on the cellulase of Odontotermes formosanus termites: a new reservoir for lignocellulase genes. Appl Environ Microbiol 2011;77:
(Isoptera: Termitidae). Sociobiology 2008b;51:733–40. 48–56.
Dheeran P, Nandhagopal N, Kumar S, Jaiswal YK, Adhikari DK. A novel thermostable Lo N, Eggleton P. Termite phylogenetics and co-cladogensis with symbionts. In: Bignell
xylanase of Paenibacillus macerans IIPSP3 isolated from the termite gut. J Ind DE, Roisin Y, Lo N, editors. Biology of Termites: A Modern Synthesis. Dordrecht:
Microbiol Biotechnol 2012;39:851–60. Springer; 2011. p. 51–67.
Eggleton P. An introduction to termites: biology, taxonomy and functional morpholo- Lo N, Tokuda G, Watanabe H. Evolution and function of endogenous termite cellulases.
gy. In: Bignell DE, Roisin Y, Lo N, editors. Biology of Termites: A Modern Synthesis. In: Bignell DE, Roisin Y, Lo N, editors. Biology of Termites: A Modern Synthesis.
Dordrecht: Springer; 2011. p. 1-26. Dordrecht: Springer; 2011. p. 51–67.
Faulet BM, Niamke S, Gonnety JT, Kouame LP. Purification and biochemical properties Lynd LR, Weimer PJ, van Zyl WH, Pretorius IS. Microbial cellulose utilization: funda-
of a new thermostable xylanase from symbiotic fungus, Termitomyces sp. Afr J mentals and biotechnology. Microbiol Mol Biol Rev 2002;66:506–77.
Biotechnol 2006;5:273–82. Matoub M, Rouland C. Purification and properties of the xylanases from the termite
Fujita A, Miura T, Matsumoto T. Differences in cellulose digestive systems among castes Macrotermes bellicosus and its symbiotic fungus Termitomyces sp. Comp Biochem
in two termite lineages. Physiol Entomol 2008;33:73–82. Physiol B Biochem Mol Biol 1995;112:629–35.
Fujita A, Hojo M, Aoyagi T, Hayashi Y, Arakawa G, Tokuda G, et al. Details of the digestive Matsui T, Tokuda G, Shinzato N. Termites as functional gene resources. Recent Pat
system in the midgut of Coptotermes formosanus Shiraki. J Wood Sci 2010;56:222–6. Biotechnol 2009;3:10–8.
Garnier-Sillam E, Toutain F, Villemin G, Renoux J. Études préliminaires des meulesoriginales Matsuura K, Yashiro T, Shimizu K, Tatsumi S, Tamura T. Cuckoo fungus mimics termite
du termite xylophage Sphaerotermes sphaerothorax (Sjostedt). Insect Soc 1989;36: eggs by producing the cellulose-digesting enzyme β-glucosidase. Curr Biol 2009;19:
293–312. 30–6.
Harreither W, Sygmund C, Dunhofen E, Vicuna R, Haltrich D, Ludwig R. Cellobiose de- Matteotti C, Haubruge E, Thonart P, Francis F, De Pauw E, Portetelle D, et al. Character-
hydrogenase from lignolytic basidiomycete Ceriporiopsis subvermispora. Appl Envi- ization of a new β-glucosidase/beta-xylosidase from the gut microbiota of the ter-
ron Microbiol 2009;75:2750–7. mite (Reticulitermes santonensis). FEMS Microbiol Lett 2011;314:147–57.
Henrissat B, Bairoch A. New families in the classification of glycosyl hydrolases based Matteotti C, Bauwens J, Brasseur C, Tarayre C, Thonart P, Destain J, et al. Identification
on amino acid sequence similarities. Biochem J 1993;293(Pt 3):781–8. and characterization of a new xylanase from gram-positive bacteria isolated from
Himmel ME, Ding SY, Johnson DK, Adney WS, Nimlos MR, Brady JW, et al. Biomass recal- termite gut (Reticulitermes santonensis). Protein Expr Purif 2012;83:117–27.
citrance: engineering plants and enzymes for biofuels production. Science 2007;315: Nakashima K, Watanabe H, Saitoh H, Tokuda G, Azuma JI. Dual cellulose-digesting sys-
804–7. tem of the wood-feeding termite, Coptotermes formosanus Shiraki. Insect Biochem
Hirai H, Shinzato N, Nakagawa A, Watanabe Y, Kurane R. Degradation of lignin model Mol Biol 2002;32:777–84.
compounds by various termites. Mokuzai Gakkaishi 2000;46:63–7. Ni J, Takehara M, Watanabe H. Heterologous overexpression of a mutant termite cellu-
Hirayama K, Watanabe H, Tokuda G, Kitamoto K, Arioka M. Purification and characteriza- lase gene in Escherichia coli by DNA shuffling of four orthologous parental cDNAs.
tion of termite endogenous β-1,4-endoglucanases produced in Aspergillus oryzae. Biosci Biotechnol Biochem 2005;69:1711–20.
Biosci Biotechnol Biochem 2010;74:1680–6. Ni J, Takehara M, Miyazawa M, Watanabe H. Random exchanges of non-conserved
Hongoh Y. Toward the functional analysis of uncultivable, symbiotic microorganisms in amino acid residues among four parental termite cellulases by family shuffling im-
the termite gut. Cell Mol Life Sci 2011;68:1311–25. proved thermostability. Protein Eng Des Sel 2007a;20:535–42.
Hu J, Arantes V, Saddler JN. The enhancement of enzymatic hydrolysis of lignocellulosic Ni J, Tokuda G, Takehara M, Watanabe H. Heterologous expression and enzymatic char-
substrates by the addition of accessory enzymes such as xylanase: is it an additive acterization of beta-glucosidase from the drywood-eating termite, Neotermes
or synergistic effect? Biotechnol Biofuels 2011;4:36. koshunensis. Appl Entomol Zool 2007b;42:457–63.
Hyodo F, Azuma J-I, Abe T. Estimation of effect of passage through the gut of a lower ter- Ni J, Takehara M, Watanabe H. Identification of activity related amino acid mutations of
mite, Coptotermes formosanus Shiraki, on lingin by solid-state CP/MAS 13C NMR. a GH9 termite cellulase. Bioresour Technol 2010;101:6438–43.
Holzforschung 1999;53:244–6. Nimchua T, Thongaram T, Uengwetwanit T, Pongpattanakitshote S, Eurwilaichitr L.
Hyodo F, Tayasu I, Inoue T, Azuma J-I, Kudo T, Abe T. Differential role of symbiotic fungi in lig- Metagenomic analysis of novel lignocellulose-degrading enzymes from higher ter-
nin degradation and food provision for fungus-growing termites (Macrotermitinae: mite guts inhabiting microbes. J Microbiol Biotechnol 2012;22:462–9.
Isoptera). Funct Ecol 2003;17:186–93. Ohkuma M. Termite symbiotic systems: efficient bio-recycling of lignocellulose. Appl
Inoue T, Moriya S, Ohkuma M, Kudo T. Molecular cloning and characterization of a cel- Microbiol Biotechnol 2003;61:1–9.
lulase gene from a symbiotic protist of the lower termite, Coptotermes formosanus. Pan CY, Zhou Y, Deng TF, Mo JC. Activities of ligninase in Odontotermes formosanus
Gene 2005;349:67–75. (Isoptera: Termitidae) and its symbiotic fungus. Sociobiology 2009;53:177–87.
Inward DJG, Vogler AP, Eggleton PA. Comprehensive phylogenetic analysis of termites Potts RC, Hewitt PH. Some properties and reaction characteristics of the partially puri-
(Isoptera) illustrates key aspects of their evolutionary biology. Mol Phylogenet Evol fied cellulase from the termite Trinervitermes trinervoides (Nasutitermitinae).
2007;44:953–67. Comp Biochem Physiol 1974;47B:327–37.
Johjima T, Taprab Y, Noparatnaraporn N, Kudo T, Ohkuma M. Large-scale identification Prins RA, Kreulen DA. Comparative aspects of plant cell wall degradation in insects. An-
of transcripts expressed in a symbiotic fungus (Termitomyces) during plant bio- imal Feed Sci Technol 1991;32:101–18.
mass degradation. Appl Microbiol Biotechnol 2006;73:195–203. Ragauskas AJ, Williams CK, Davison BH, Britovsek G, Cairney J, Eckert CA, et al. The path
Juturu V, Wu JC. Microbial xylanases: engineering, production and industrial applica- forward for biofuels and biomaterials. Science 2006;311:484–9.
tions. Biotechnol Adv 2012;30:1219–27. Ransom-Jones E, Jones DL, McCarthy AJ, McDonald JE. The fibrobacteres: an important
Ke J, Sun JZ, Nguyen HD, Singh D, Lee KC, Beyenal H, et al. In-situ oxygen profiling phylum of cellulose-degrading bacteria. Microb Ecol 2012;63:267–81.
and lignin modification in guts of wood-feeding termites. Insect Sci 2010;17: Rouland-Lefèvre C. Symbiosis with fungi. In: Abe T, Bignell DE, Higashi M, editors. Termites:
277–90. Evolution, Sociality, Symbioses, Ecology. Dordrecht, The Netherlands: Kluwer Academic
Ke J, Laskar DD, Singh D, Chen SL. In situ lignocellulosic unlocking mechanism for car- Publishers; 2000. p. 289–306.
bohydrate hydrolysis in termites: crucial lignin modification. Biotechnol Biofuels Saha BC. Hemicellulose bioconversion. J Ind Microbiol Biotechnol 2003;30:279–91.
2011a;4:17. Scharf ME, Tartar A. Termite digestomes as sources for novel lignocellulases. Biofuel
Ke J, Singh D, Chen SL. Aromatic compound degradation by the wood-feeding termite Bioprod Bior 2008;2:540–52.
Coptotermes formosanus (Shiraki). Int Biodeterior Biodegrad 2011b;65:744–56. Scharf ME, Kovaleva ES, Jadhao S, Campbell JH, Buchman GW, Boucias DG. Functional
Ke J, Singh D, Yang XSC. Thermal characterization of softwood lignin modification by and translational analyses of a beta-glucosidase gene (glycosyl hydrolase family
termite Coptotermes formosanus (Shiraki). Biomass Bioener 2011c;35:3617–26. 1) isolated from the gut of the lower termite Reticulitermes flavipes. Insect Biochem
Ke J, Laskar DD, Gao DF, Chen SL. Advanced biorefinery in lower termite-effect of com- Mol Biol 2010;40:611–20.
bined pretreatment during the chewing process. Biotechnol Biofuels 2012a;5:11. Scharf ME, Karl ZJ, Sethi A, Boucias DG. Multiple levels of synergistic collaboration in
Ke J, Singh D, Chen SL. Metabolism of polycyclic aromatic hydrocarbons by the wood-feeding termite lignocellulose digestion. PLoS One 2011;6:e21709.
termite Coptotermes formosanus (Shiraki). J Agric Food Chem 2012b;60:1788–97. Sethi A, Slack J, Kovaleva ES, Buchman GW, Scharf ME. Lignin-associated metagene expres-
Köhler T, Dietrich C, Scheffrahn RH, Brune A. High-resolution analysis of gut environment and sion in a lignocellulose-digesting termite. Insect Biochem Mol Biol 2013;43:91-101.
bacterial microbiota reveals functional compartmentation of the gut in wood-feeding Shimada K, Maekawa K. Changes in endogenous cellulase gene expression levels and
higher termites (Nasutitermes spp.). Appl Environ Microbiol 2012;78:4691–701. reproductive characteristics of primary and secondary reproductives with colony
Korb J, Weil T, Hoffmann K, Foster KR, Rehli M. A gene necessary for reproductive sup- development of the termite Reticulitermes speratus (Isoptera: Rhinotermitidae). J
pression in termites. Science 2009;324:758. Insect Physiol 2010;56:1118–24.
Kudo T. Termite-microbe symbiotic system and its efficient degradation of lignocellu- Sillam-Dussès D, Krasulová J, Vrkoslav V, Pytelková J, Cvačka J, Kutalová K, et al. Comparative
lose. Biosci Biotechnol Biochem 2009;73:2561–7. study of the labial gland secretion in termites (Isoptera). PLoS One 2012;7:e46431.
Kyou K, Watanabe T, Yoshimura T, Takahashi M. Lignin modification by termite and its Sinma K, Khucharoenphaisan K, Kitpreechavanich V, Tokuyama S. Purification and charac-
symbiotic protozoa. Wood Res 1996;83:50–4. terization of a thermostable xylanase from Saccharopolyspora pathumthaniensis S582
Le Roes-Hill M, Rohland J, Burton S. Actinobacteria isolated from termite guts as a isolated from the gut of a termite. Biosci Biotechnol Biochem 2011;75:1957–63.
source of novel oxidative enzymes. Antonie Van Leeuwenhoek InterJ General Mol Slaytor M. Energy metabolism in the termite and its gut microbiota. In: Abe T, Bignell
Microbiol 2011;100:589–605. DE, Higashi M, editors. Termites: Evolution, Sociality, Symbioses, Ecology. Dor-
Leonardo FC, da Cunha AF, da Silva MJ, Carazzolle MF, Costa-Leonardo AM, Costa FF, drecht, The Netherlands: Kluwer Academic Publishers; 2000. p. 307–32.
et al. Analysis of the workers head transcriptome of the Asian subterranean ter- Smith JA, Scharf ME, Pereira RM, Koehler PG. Comparisons of gut carbohydrolase ac-
mite, Coptotermes gestroi. Bull Entomol Res 2011;101:383–91. tivity patterns in Reticulitermes flavipes and Coptotermes formosanus (Isoptera:
Li HJ, Lu JR, Mo JC. Physiochemical lignocellulose modification by the Formosan subter- Rhinotermitidae) workers and soldiers. Sociobiology 2009;53:113–24.
ranean termite Coptotermes formosanus Shiraki (Isoptera: Rhinotermitidae) and its Taechapoempol K, Sreethawong T, Rangsunvigit P, Namprohm W, Thamprajamchit B,
potential uses in the production of biofuels. Bioresources 2012;7:675–85. Rengpipat S, et al. Cellulase-producing bacteria from Thai higher termites,
850 J. Ni, G. Tokuda / Biotechnology Advances 31 (2013) 838–850

Microcerotermes sp.: enzymatic activities and ionic liquid tolerance. Appl Biochem Uchima CA, Tokuda G, Watanabe H, Kitamoto K, Arioka M. A novel glucose-tolerant
Biotechnol 2011;164:204–19. β-glucosidase from the salivary gland of the termite Nasutitermes takasagoensis. J
Taprab Y, Johjima T, Maeda Y, Moriya S, Trakulnaleamsai S, Noparatnaraporn N, et al. Gen Appl Microbiol 2013. [in press].
Symbiotic fungi produce laccases potentially involved in phenol degradation in Wang Q, Qian C, Zhang XZ, Liu N, Yan X, Zhou Z. Characterization of a novel
fungus combs of fungus-growing termites in Thailand. Appl Environ Microbiol thermostable β-glucosidase from a metagenomic library of termite gut. Enzyme
2005;71:7696–704. Microb Technol 2012;51:319–24.
Tartar A, Wheeler MM, Zhou X, Coy MR, Boucias DG, Scharf ME. Parallel metatranscriptome Warnecke F, Luginbuhl P, Ivanova N, Ghassemian M, Richardson TH, Stege JT, et al.
analyses of host and symbiotic gene expression in the gut of the termite Reticulitermes Metagenomic and functional analysis of hindgut microbiota of a wood-feeding
flavipes. Biotechnol Biofuels 2009;2:25. higher termite. Nature 2007;450:560–5.
Todaka N, Moriya S, Saita K, Hondo T, Kiuchi I, Takasu H, et al. Environmental cDNA Watanabe H, Tokuda G. Animal cellulases. Cell Mol Life Sci 2001;58:1167–78.
analysis of the genes involved in lignocellulose digestion in the symbiotic protist Watanabe H, Tokuda G. Cellulolytic systems in insects. Annu Rev Entomol 2010;55:
community of Reticulitermes speratus. FEMS Microbiol Ecol 2007;59:592–9. 609–32.
Todaka N, Inoue T, Saita K, Ohkuma M, Nalepa CA, Lenz M, et al. Phylogenetic analysis Watanabe H, Takase A, Tokuda G, Yamada A, Lo N. Symbiotic “Archaezoa” of the prim-
of cellulolytic enzyme genes from representative lineages of termites and a related itive termite Mastotermes darwiniensis still play a role in cellulase production.
cockroach. PLoS One 2010a;5:e8636. Eukaryot Cell 2006;5:1571–6.
Todaka N, Lopez CM, Inoue T, Saita K, Maruyama J, Arioka M, et al. Heterologous ex- Wilson DB. Evidence for a novel mechanism of microbial cellulose degradation. Cellu-
pression and characterization of an endoglucanase from a symbiotic protist of lose 2009;16:723–7.
the lower termite, Reticulitermes speratus. Appl Biochem Biotechnol 2010b;160: Wu Y, Chi S, Yun C, Shen Y, Tokuda G, Ni J. Molecular cloning and characterization of an
1168–78. endogenous digestive β-glucosidase from the midgut of the fungus-growing ter-
Tokuda G, Watanabe H. Hidden cellulases in termites: revision of an old hypothesis. mite Macrotermes barneyi. Insect Mol Biol 2012;21:604–14.
Biol Lett 2007;3:336–9. Xie L, Zhang L, Zhong Y, Liu N, Long YH, Wang SY, et al. Profiling the metatranscriptome
Tokuda G, Watanabe H, Matsumoto T, Noda H. Cellulose digestion in the wood-eating of the protistan community in Coptotermes formosanus with emphasis on the
higher termite, Nasutitermes takasagoensis (Shiraki): distribution of cellulases and lignocellulolytic system. Genomics 2012;99:246–55.
properties of endo-β-1,4-glucanase. Zool Sci 1997;14:83–93. Yamada A, Inoue T, Wiwatwitaya D, Ohkuma M, Kudo T, Abe T, et al. Carbon mineral-
Tokuda G, Lo N, Watanabe H, Slaytor M, Matsumoto T, Noda H. Metazoan cellulase ization by termites in tropical forests, with emphasis on fungus combs. Ecol Res
genes from termites: intron/exon structures and sites of expression. Biochim 2005;20:453–60.
Biophys Acta 1999;1447:146–59. Yuki M, Moriya S, Inoue T, Kudo T. Transcriptome analysis of the digestive organs of
Tokuda G, Saito H, Watanabe H. A digestive β-glucosidase from the salivary glands of Hodotermopsis sjostedti, a lower termite that hosts mutualistic microorganisms in
the termite, Neotermes koshunensis (Shiraki): distribution, characterization and its hindgut. Zool Sci. 2008;25:401–6.
isolation of its precursor cDNA by 5′- and 3′-RACE amplifications with degenerate Zhang DH, Lax AR, Raina AK, Bland JM. Differential cellulolytic activity of native-form
primers. Insect Biochem Mol Biol 2002;32:1681–9. and C-terminal tagged-form cellulase derived from Coptotermes formosanus and
Tokuda G, Lo N, Watanabe H, Arakawa G, Matsumoto T, Noda H. Major alteration of the expressed in E. coli. Insect Biochem Mol Biol 2009;39:516–22.
expression site of endogenous cellulases in members of an apical termite lineage. Zhang DH, Lax AR, Bland JM, Yu JJ, Fedorova N, Nierman WC. Hydrolysis of filter-paper
Mol Ecol 2004;13:3219–28. cellulose to glucose by two recombinant endogenous glycosyl hydrolases of
Tokuda G, Lo N, Watanabe H. Marked variations in patterns of cellulase activity against Coptotermes formosanus. Insect Sci 2010;17:245–52.
crystalline- vs. carboxymethyl-cellulose in the digestive systems of diverse, Zhang DH, Lax AR, Bland JM, Allen AB. Characterization of a new endogenous endo-β-1,4-
wood-feeding termites. Physiol Entomol 2005;30:372–80. glucanase of Formosan subterranean termite (Coptotermes formosanus). Insect Biochem
Tokuda G, Watanabe H, Lo N. Does correlation of cellulase gene expression and cellulo- Mol Biol 2011;41:211–8.
lytic activity in the gut of termite suggest synergistic collaboration of cellulases? Zhang D, Allen AB, Lax AR. Functional analyses of the digestive β-glucosidase of Formo-
Gene 2007;401:131–4. san subterranean termites (Coptotermes formosanus). J Insect Physiol 2012a;58:
Tokuda G, Miyagi M, Makiya H, Watanabe H, Arakawa G. Digestive β-gluosidases from 205–10.
the wood-feeding higher termite, Nastitermes takasagoensis: intestinal distribution, Zhang D, Lax AR, Henrissat B, Coutinho P, Katiya N, Nierman WC, et al. Carbohydrate-active
molecular characterization, and alteration in sites of expression. Insect Biochem enzymes revealed in Coptotermes formosanus (Isoptera: Rhinotermitidae) transcriptome.
Mol Biol 2009;39:931–7. Insect Mol Biol 2012b;21:235–45.
Tokuda G, Watanabe H, Hojo M, Fujita A, Makiya H, Miyagi M, et al. Cellulolytic environ- Zhou X, Smith JA, Oi FM, Koehler PG, Bennett GW, Scharf ME. Correlation of cellulase gene
ment in the midgut of the wood-feeding higher termite Nasutitermes takasagoensis. expression and cellulolytic activity throughout the gut of the termite Reticulitermes
J Insect Physiol 2012;58:147–54. flavipes. Gene 2007;395:29–39.
Tomme P, Warren RAJ, Gilkes NR. Cellulose hydrolysis by bacteria and fungi. Adv Microb Zhou X, Wheeler MM, Oi FM, Scharf ME. Inhibition of termite cellulases by carbohydrate-
Physiol 1995;37:1-81. based cellulase inhibitors: evidence from in vitro biochemistry and in vivo feeding
Uchima CA, Arioka M. Expression and one-step purification of recombinant proteins studies. Pesticide Biochem Physiol 2008a;90:31–41.
using an alternative episomal vector for the expression of N-tagged heterologous Zhou X, Wheeler MM, Oi FM, Scharf ME. RNA interference in the termite Reticulitermes
proteins in Pichia pastoris. Biosci Biotechnol Biochem 2012;76:368–71. flavipes through ingestion of double-stranded RNA. Insect Biochem Mol Biol 2008b;38:
Uchima CA, Tokuda G, Watanabe H, Kitamoto K, Arioka M. Heterologous expression and 805–15.
characterization of a glucose-stimulatedβ-glucosidase from the termite Neotermes Zhou X, Kovaleva ES, Wu-Scharf D, Campbell JH, Buchman GW, Boucias DG, et al. Produc-
koshunensis in Aspergillus oryzae. Appl Microbiol Biotechnol 2011;89:1761–71. tion and characterization of a recombinant β-1,4-endoglucanase (glycohydrolase fam-
Uchima CA, Tokuda G, Watanabe H, Kitamoto K, Arioka M. Heterologous expression in ily 9) from the termite Reticulitermes flavipes. Arch Insect Biochem Physiol 2010a;74:
Pichia pastoris and characterization of an endogenous thermostable and high 147–62.
glucose-tolerant β-glucosidase from the termite Nasutitermes takasagoensis. Appl Zhou Y, Deng TF, Pan CY, Chen CR, Mo JC. Purification of a laccase from fungus combs in
Environ Microbiol 2012;78:4288–93. the nest of Odontotermes formosanus. Process Biochem 2010b;45:1052–6.

View publication stats

Potrebbero piacerti anche