You are on page 1of 24

REVIEW ARTICLE

Lignin-degrading enzymes
Loredano Pollegioni1,2, Fabio Tonin1 and Elena Rosini1,2
 degli studi dell’Insubria, Varese, Italy
1 Dipartimento di Biotecnologie e Scienze della Vita, Universita
2 The Protein Factory, Centro Interuniversitario di Biotecnologie Proteiche, Politecnico di Milano, ICRM CNR Milano, Universit
a degli Studi
dell’Insubria, Italy

Keywords A main goal of green biotechnology is to reduce our dependence on fossil


biorefinery; green biotechnology; lignin reserves and to increase the use of renewable materials. For this, lignocellu-
valorization; ligninolytic enzymes;
lose, which is composed of cellulose, hemicellulose and lignin, represents
lignocellulosic biomass
the most promising feedstock. The latter is a complex aromatic heteropoly-
Correspondence mer formed by radical polymerization of guaiacyl, syringyl, and p-hydroxy-
L. Pollegioni, Dipartimento di Biotecnologie phenyl units linked by b-aryl ether linkages, biphenyl bonds and
e Scienze della Vita, Universita degli studi heterocyclic linkages. Accordingly, lignin appears to be a potentially valu-
dell’Insubria, via J. H. Dunant 3, able renewable aromatic chemical, thus representing a main pillar in future
21100 Varese, Italy biorefinery. The resistance of lignin to breakdown is the main bottleneck in
Fax: +39 0332421500
this process, although a variety of white-rot fungi, as well as bacteria, have
Tel: +39 0332421506
been reported to degrade lignin by employing different enzymes and cata-
E-mail: loredano.pollegioni@uninsubria.it
bolic pathways. Here, recent investigations have expanded the range of nat-
(Received 16 September 2014, revised 29 ural biocatalysts involved in lignin degradation/modification and significant
December 2014, accepted 30 January 2015) progress related to enzyme engineering and recombinant expression has
been made. The present review is focused primarily on recent trends in lig-
doi:10.1111/febs.13224 ninolytic green biotechnology to suggest the potential (industrial) applica-
tion of ligninolytic enzymes. Future perspectives could include synergy
between natural enzymes from different sources (as well as those obtained
by protein engineering) and other pretreatment methods that may be
required for optimal results in enzyme-based, environmentally friendly,
technologies.

Introduction
The term lignin derives from the Latin word ‘lignum’, Chemically, lignin is a cross-linked macromolecular
which means wood. It represents one of the three com- material derived from oxidative coupling of monolig-
ponents of lignocellulosic biomass, along with cellulose nols, mainly hydroxycinnamyl alcohols: the three main
and hemicellulose. Lignin is the second most abundant components are p-coumaryl, coniferyl and synapyl
natural substance in nature after cellulose and, annu- alcohols. Lignins show plant-specific composition, with
ally, approximately 5 9 106 metric tons of lignin is molecular weight and linkage motifs (Fig. 1) varying
produced industrially [1]. Indeed, lignin is the most according to plant species and environmental factors.
abundant renewable source of aromatic polymers on Lignification is obtained by cross-linking reactions
earth: its degradation is mandatory for carbon recy- of the lignin monomers or by polymer–polymer cou-
cling. pling via radicals produced by oxidases: the resonance

Abbreviations
ABTS, 2,20 -azino-bis(3-ethylbenzothiazoline-6-sulphonic acid); DyP, dye-decolorizing peroxidases; HBT, 1-hydroxybenzo-triazole; LiP, lignin
peroxidase; LRET, long-range electron transfer; MnP, manganese peroxidase; PDB, Protein Data Bank; VA, veratryl alcohol; VP, versatile
peroxidase.

1190 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

Fig. 1. (A) Structure of primary monomers


(top) and corresponding structural units in
lignin (bottom) indicated as p-hydrophenyl
(H), guaiacyl (G) and syringyl (S) units. (B)
Schematic representation of a lignin
structure: the most frequent bonds are
indicated. Softwood lignin contains mostly
G units and very low levels of H units;
hardwood lignin contains similar levels of
G and S units (and traces of H units); and
grasses contain all three units (G/S/H ratio
is 96 : traces : 4, 50 : 50 : traces,
70 : 25 : 5, respectively).

delocalization of radicals to couple at different sites lenge is to develop sustainable technologies for con-
yields an array of units linked by carbon–carbon and verting waste biomass to biofuels, chemicals and new
carbon–oxygen (ether) bonds. The most frequent biobased materials (such as bioplastics).
bonds include b-O-4, b-5, b-b, 5-5, 5-O-4 and b-1 cou- Most biorefineries focus on cellulose and hemicellu-
plings (Fig. 1B). lose valorization, the so-called ‘sugar-based platform’
As a result of its complex structure, almost 98% of [2], whereas lignin is usually considered to be a low-
this renewable polymer derived from biomass is value product. A major aim of modern biorefineries is
burned as a source of energy in the pulp and paper to (efficiently) convert lignocellulosic materials into
industry. Accordingly, a great biotechnological chal- valuable products by unraveling the limitations as a

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1191


Enzymatic degradation of lignin L. Pollegioni et al.

result of a non-uniform structure, unique chemical ral resources) and available at low cost [9–12]. Indeed,
reactivity and the presence of various organic and the process of lignin degradation is enhanced by sev-
inorganic impurities in lignin samples. The controlled eral additional and accessory enzymes, and some of
breaking down of C-C and C-O bonds in lignin by a them are now available. A recent review reporting on
selective depolymerization process could produce a immobilization of LiP, MnP and laccase is also avail-
whole series of monomeric, aromatic species. Thus, lig- able [13].
nin could represent a renewable source of aromatic
fine chemicals.
Peroxidases
The main strategies for depolymerizing lignin to
produce valuable chemicals are cracking or hydro- The extracellular fungal LiP, MnP and VP are all
lysis reactions, catalytic reduction, catalytic oxidation members of class II peroxidases within the superfamily
and enzymatic reactions [3–6]. A detailed description of heme peroxidases [14].
of available lignin preparations is provided elsewhere
[7].
LiP
The present review summarizes the state-of-art in
enzymatic systems active on the main lignin bonds, Reaction and properties
with a particular emphasis on heterologous enzyme
LiP (EC 1.11.1.14), a glycoprotein of 38–46 kDa (and
expression and protein engineering studies that aim to
pI of 3.2–4.0) containing 1 mol of iron protoporphyrin
improve their practical use. In the 1980s, reactive oxy-
IX per 1 mol of protein, catalyzes the H2O2-dependent
gen species were reported to be involved in cleaving
oxidative depolymerization of lignin. The reaction can
lignin chains in basidiomyceteous white-rot fungal cul-
be represented as:
tures: the ligninolytic activity in Phanerochaete chry-
sosporium cell extracts was correlated with production 1; 2  bisð3; 4  dimethoxyphenylÞpropane  1; 3  diol
of H2O2 [8]. Over the years, four enzymatic activities þ H2 O2
3; 4  dimethoxybenzaldehyde þ 1  ð3; 4
have been reported to depolymerize lignin in decaying  dimethoxyphenylÞethane  1; 2  diol þ H2 O
plant cell walls: lignin peroxidases (LiPs), manganese
ð1Þ
peroxidases (MnPs), versatile peroxidases (VPs) and
laccases. These enzymes have gained attention as LiP shows a very high redox potential (E00  1.2 V
potential biological catalysts for lignin biodegradation. versus SHE at pH 3.0) compared to laccases ( 0.8 V
Of note, lignocellulolytic enzymes are too large to at pH 5.5), horseradish peroxidases (0.95 V at pH 6.3)
penetrate into undecayed wood cell walls; therefore, and MnP (0.8 V at pH 4.5) [15]: this property enables
reactive oxygen species could be the agents responsible LiP to catalyze the oxidation of nonphenolic aromatic
for local lignin decay. Phenolic compounds related to compounds, even in the absence of a mediator. The
recurring lignin phenolic moieties are known to medi- reaction catalyzed by LiP is reported in Fig. 2A [7]. In
ate the oxidation of nonphenolic compounds by a more detail: 2-electron oxidation of ferric [Fe(III)] LiP
radical mechanism, once oxidized by laccases to phen- produces compound I intermediate, a oxoferryl iron
oxyl radicals. Laccases possess relative low redox porphyrin radical cation [Fe(IV)=O+] with the reduc-
potentials (≤ 0.8 V) that restrict their action to the tion of H2O2. Next, two consecutive one-electron
oxidation of the phenolic lignin components. The oxi- reduction steps of compound I [from an electron
dation of nonphenolic lignin moieties is performed via donor substrate, such as veratryl alcohol (VA)], yield-
redox mediators: compounds of small molecular ing at first compound II [Fe(IV)=O] and VA radical
weight act as electron shuttles performing the oxida- cation [VA+] and then returning LiP to the ferric oxi-
tion of complex polymers (e.g. lignin) that do not dation state, completed the catalytic cycle (central loop
have access to the enzyme’s active site. After the enzy- in Fig. 2B). Indeed, a single 2-electron reduction can
matic oxidation and the conversion into stable radi- return compound I to the resting state in some cases
cals, mediators diffuse far away, thus enabling the [7]. Oxidation of phenolic compounds is associated
oxidation of target compounds not considered to be with inactivation of LiP as a result of the accumula-
substrates of laccase, as a result of their high redox tion of compound III (Fig. 2B); VA+ appears to
potential and high volume. An ideal redox mediator protect LiP from inactivation as a result of the pro-
would be a small-size molecule capable of generating duction of compound III [16]. Notably, in the presence
stable radicals (in the oxidized form) that do not inac- of dioxygen, additional oxidative steps take place to
tivate the enzyme, and that is also easily recyclable, form the quinine form and promote aromatic ring
environmental-friendly (specially obtained from natu- cleavage.

1192 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

Fig. 2. Reaction catalyzed by LiP. (A)


Mechanism of the reaction of LiP on a B
nonphenolic arylglycerol b-aryl ether lignin
model compound [31, 124]. (B) Details of
the steps in the catalytic cycle of LiP in
the presence of VA as substrate:
compounds I, II and III are intermediates
with differing heme oxidation states [7].
The lower loop represents the reaction of
LiP compound II with H2O2 at pH 3.0 in
the presence of excess H2O2 and the
absence of a reducing substrate, yielding a
catalytic inactive form of the enzyme,
known as compound III. Compound III is
then converted to the resting enzyme by
spontaneous autoxidation or by oxidation
with a VA radical cation.

LiPs are active on a wide range of aromatic com- the distal H146. The peroxide binding site is located
pounds, such as VA, methoxybenzenes and nonphenol- on the distal side of the heme; a channel connects this
ic b-O-4 linked arylglycerol b-aryl ethers with a redox pocket to the exterior of the protein. The negative
potential up to 1.4 V in the presence of H2O2 [16] charge(s) that develops in the cleavage of peroxides is
(Table 1). Using a lignin model dimer as the substrate, stabilized by R43, which also stabilizes the ferryl oxy-
the cation radical decays with the spontaneous Ca-Cb gen of compound I; the distal H47 acts as a H+ accep-
fission of the alkyl side chain, with the products resem- tor for the bound peroxide. Notably, W171 on the
bling those found when fungi degrade lignin [17]. LiPs protein surface shows hydroxylation at Cb (post-trans-
can oxidize phenolic lignin model compounds more lational modification): this residue plays a main role in
readily than nonphenolic ones [18]. Unfortunately, the the binding and oxidation of VA through long-range
size of the oligomeric lignin model substrates signifi- electron transfer (LRET), as also demonstrated by
cantly affects the catalytic efficiency of LiP (i.e. the site-directed mutagenesis studies [24]. Characterization
activity of LiP on a b-O-4-linked lignin model trimer of W171A (which is unable to oxidize VA) and F267L
is 25-fold lower than that observed on VA) [19]. (showing an increased Km) variants of LiP isoenzyme
The crystal structure of P. chrysosporium LiP is H8 from Phanerochaete crysosporium demonstrated
composed of eight main and eight minor a-helices and that the site including these two residues, rather than
a limited b structure, and it is organized in a proximal the heme access channel, is the site of VA oxidation
and a distal domain (Fig. 3) [20–23]. The heme is [25].
embedded in a crevice between the domains: two small
channels connect the prosthetic group to the solvent
Expression of LiP
(see below). The LiP contains four disulfide bridges,
two Ca2+ binding sites (one per each domain, puta- Very recently, a strain of P. crysosporium was geneti-
tively involved in maintaining the topology of the cally modified to co-express two endogenous genes
active site) and two glycosylation sites in the proximal encoding for a LiP (lipH8) and a MnP (mnp1) yielding
domain. The heme iron is coordinated with H176 at 4.5 kUL1 and 9.1 kUL1, respectively, and a spe-
the proximal side, and W339 is hydrogen-bonded to cific activity of 29.9 U(mg LiP)1 (on VA as sub-

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1193


Enzymatic degradation of lignin L. Pollegioni et al.

Table 1. Structures of most common substrates for the different Table 1. (Continued).
lignin-degrading enzymes reported in the present review.
Substrate Structure
Substrate Structure
Sinapyl alcohol
2,6-Dimeth-
oxyphenol

4-Methoxyphenol Syringyl alcohol

ABTS

Veratryl alcohol

Cathecol

strate) and 11.7 U(mg MnP)1 (as Mn3+-malonate


Coniferyl alcohol complex formation) [26].
Concerning recombinant expression, an engineered
cDNA from P. crysosporium encoding the LiP isoen-
zyme H2 has been successfully overexpressed in Esc-
herichia coli [27]. It was produced in an inactive and
Guaiacol insoluble form in inclusion bodies: the active enzyme
was produced by refolding with a buffer containing
glutathione, urea, Ca2+ and heme (final yield of
3.4 mg enzymeL1 of culture). The turnover number
of purified recombinant LiP (at pH 2.5) was 39 s1,
Methoxybenzene measured as the oxidation of VA; this was similar to
the value obtained when the enzyme was isolated from
the same source (30 s1). Subsequently, the recombi-
nant enzyme was reported to possess a MnP activity,
with a turnover number of 14 s1.
The LiP isoenzyme H8 from P. chrysosporium was
Methylhydr-
also overexpressed in E. coli: the recombinant pro-
oquinone
tein was recovered from inclusion bodies with a
folding efficiency of 1% in terms of active enzyme
[24]. Indeed, this enzyme was homologously overex-
pressed under the control of the glyceraldehyde 3-
phosphate dehydrogenase promoter, reaching approx-
Phenol red
imately 2 mgL1 of pure enzyme in 4 days of
growth [28].
P. chrysosporium LiP was also expressed in Pichi-
a methanolica cells under alcohol oxidase promoter
control and contained the LiP a-factor signal pep-
tide from P. chrysosporium or Saccharomyces cerevisi-
ae; the extracellular activity was 932 UL1 and
1933 UL1, respectively [29].

1194 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

A B

Fig. 3. Crystal structure of


P. chrysosporium LiP [Protein Data Bank
(PDB) code: 1LGA]. (A) Tertiary structure
of LiP: blue, proximal, C-terminal domain;
green, distal, N-terminal domain; heme is
indicated as yellow ball sticks. (B) Details
of the heme environment. The proximal
H176 forms hydrogen-bonds with D238,
which facilitates stabilization of Fe(IV)-O
intermediate in compound I (the Asp-His-
Fe triad is a conserved feature of plant
peroxidase superfamily).

In more recent work, the cDNA encoding a MnP also reported) [33,34]. Details concerning lignin char-
(mnp1) and that encoding for a LiP (clg5) were fused and acterization using direct (based on unmodified lignin
the chimeric protein was expressed in E. coli cells [30]. samples) and indirect (via chemical modification of
Following protein refolding and activation (in the pres- the samples) methods have been reported previously
ence of Ca2+ and heme) the MnP activity reached [12].
27 UL1, with a specific activity of 0.3 U(mg protein)1,
whereas the LiP activity was below detection levels.
MnP
Reaction and properties
Drawbacks in using LiP for oxidative depolymerization
of lignin MnP (EC 1.11.1.13) is a glycosylated heme protein
with molecular mass of ~ 40–50 kDa; it represents
A purified isoenzyme from the fungus P. chrysospori-
the most common lignin-modifying peroxidase pro-
um in aqueous/organic solvent mixtures at low lignin
duced and secreted by almost all wood-colonizing
concentrations (~ 4 mgL1) was able to cleave the
basidiomycetes, including white-rot and various soil-
Ca-Cb bond of the propyl side chain of synthetic
colonizing fungi. This enzymatic activity was detected
lignins to give benzylic aldehydes at the Ca position
in P. chrysosporium almost 30 years ago [35,36]: the
(i.e. soluble lower molecular weight products similar
46-kDa acidic (pI 4.5) glycoprotein was demonstrated
to those obtained from P. chrysosporium cultures
to depolymerize and (re)polymerize the synthetic lig-
incubated with lignin for 6–24 h) [31]. The main
nin molecules produced starting from coniferyl and/or
drawback was the polymerization reaction that
sinapyl alcohol in the presence of 0.2 mM Mn2+, at
occurred because the small phenolic products were
pH 4.5 and at 37 °C [37]. MnP catalyzes the reac-
not removed. The preference of LiP for phenolic lig-
tion:
nin units explains the propensity to polymerize rather
than depolymerize lignin samples under in vitro condi-
2Mn(II) þ 2Hþ þ H2 O2
2Mn(III) þ 2H2 O ð2Þ
tions [32]. To minimize phenoxy radical coupling, lim-
iting low concentrations of the substrate enabled the Indeed, MnP is the only heme peroxidase showing a
depolymerization and re-polymerization of a synthetic one-electron Mn2+ oxidation reaction mechanism
lignin obtained from coniferyl alcohol to a similar (Fig. 4). The Mn3+ dissociates from the enzyme and is
extent [31]. complexed by carboxylic acids, in particular oxalate
Lignin isolated samples are very polydisperse: the and malate: the Mn3+-chelator complex works as a
measurement of the molecular size range of lignin, as diffusible oxidant of phenolic compounds, giving a
well as the changes that can be caused by the process phenoxy radical intermediate that yields various degra-
of lignin degradation, is a challenging task. For deter- dation products. The Mn3+ form oxidizes phenolic
mination of the molecular size of lignin samples, the compounds (such as 2,6-dimethoxyphenol, guaiacol, 4-
most advantageous methods are gel permeation chro- methoxyphenol and phenolic lignin residues), whereas
matography, light scattering and vapor pressure it is inactive on VA or nonphenolic substrates
osmometry (the use of ultrafiltration and MS was (Table 1) [38].

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1195


Enzymatic degradation of lignin L. Pollegioni et al.

Recombinant expression of MnP from P. chrysosporium


P. chrysosporium possesses at least six different MnP.
The recombinant MnP from P. chrysosporium (isoen-
zyme H4) has been expressed in E. coli [43]: the inac-
tive and insoluble protein was refolded in the presence
of 2 M urea, CaCl2, heme and oxidized glutathione.
The recombinant, active MnP showed the same spec-
tral properties as the native fungal enzyme, a kcat of
450 s1 on Mn2+, and Km values for H2O2 and Mn2+
of 96 lM and 52 lM, respectively (Table 2).
P. chrysosporium MnP was also successfully
expressed in Pichia pastoris cells [44]: active MnP
(containing the a-factor secretion signal) was secreted
in broth culture, reaching a maximum of 120 UL1.
The kcat value for H2O2 was lower (168 s1), whereas
the Km value for Mn2+ was similar to that of native
Fig. 4. The catalytic cycle of the MnP resembles that of LiP
enzyme (Table 2). The expression level of the active
(Fig. 2B). The conversion of compound I into compound II can be MnP was increased by adding heme to the fermenta-
achieved by using alternative electron donors (i.e. ferrocyanide and tion broth: a logarithmic relationship between heme
a variety of phenolic compounds) [125]. concentration and MnP activity was observed
[45]. The scale-up to a 2-L fed-batch bioreactor, in
the presence of 0.1 gL1 of heme, increased the
MnP activity to 2500 UL1, corresponding to
LiP and MnP from P. chrysosporium show a similar
15.6 mgL1.
tertiary structure based on a helical topography: they
are globular proteins formed by 11–12 a-helices pre-
dominantly in two domains, with a central cavity con- Protein engineering of MnP
taining the heme group (Fig. 5A) [20,39,40]. The
The interconversion of structure/function between
structural similarity points to divergent evolution. Pro-
MnP and LiP was made apparent by site-directed
tein stability is strongly based on disulfide bridges: LiP
mutagenesis studies. The S168W variant of P. chrysog-
contains four disulfide bridges, whereas MnP contains
enum MnP was demonstrated to be a hybrid enzyme
an additional bridge (in its C-terminal region) and two
because it retained the manganese oxidase activity
Ca2+ binding sites that are located in a cation-binding
(kcat  260 s1 at pH 4.3) (Table 2) and acquired the
site on the protein surface; the site can accommodate a
ability to oxidize a multitude of LiP substrates (i.e. a
number of different metal ions. Concerning the heme
kcat value of 11 s1 was observed in the presence of
prosthetic group, it is stabilized by the two Ca2+ ions:
VA) [46,47]. S168 in MnP corresponds to the W171
one tightly bound on the proximal side and the other
residue located on the surface of LiP (i.e. the VA oxi-
bound on the distal side. Mn2+ binding is largely a
dation site).
result of R177 whose orientation is dependent on a
As noted above, MnP purified from P. chrysospori-
salt-bridge with E35 [41]. The extracellular MnP from
um is susceptible to thermal inactivation as a result of
P. chrysosporium, for which the crystal structure was
 resolution, showed the presence of a loss of the Ca2+ ion [48], and the recombinant
solved at 1.45 A
enzyme, lacking glycosylation, was also more suscepti-
divalent Cd2+ bound at the Mn-binding site: this rep-
ble [49]. The double substitution A48C and A63C
resented the first report of MnP-inhibitor complex
introduced a disulfide bond in proximity to the distal
structure [42]. Cd2+ also binds to a putative second
calcium-binding site of MnP [50]. This MnP variant
weak metal-binding site (constituted by the carboxy-
showed activity and spectral properties resembling
terminal oxygen, the side chain of D84 and two sol-
those of wild-type, glycosylated MnP but proved to be
vent molecules) displaying tetrahedral geometry. In the
more stable: the half-life of inactivation at 42 °C
open conformation of E39, a network of hydrogen
increased from 2 min for the wild-type to > 1 h for
bonds connect E39 and D84 (i.e. the ligands in the
the enzyme variant. Indeed, the enzyme variant was
two metal sites) through D85 and water. This second
also more stable at higher pH values (i.e. at pH 6.0).
site is exposed to the solvent and could accommodate
Furthermore, the M273L and N81S substitutions
a Mn2+ ion with a bound chelator.

1196 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

A B

Fig. 5. Structure of MnP and VP. (A)


Schematic representation of the 3D
structure of MnP from P. chrysosporium
(PDB code: 1YYD) [40]: blue) proximal,
C-terminal domain; green) distal, N-
terminal domain; heme is indicated as
yellow ball sticks. (B) 3D structure of
P. eryngii VP (PDB code: 2BOQ) [55]. (C)
Axial view of the heme region, including
the putative Mn2+ binding site in the
crystal structure of P. eryngii VP [54]. The
putative LRET pathways from H232 and
from W164 (pathways I and II) are
indicated as well as pathway III from P76,
absent in vpl and envisaged from the
homology model of isoenzyme VSP1 [55].

Table 2. Kinetic parameters of different MnPs.

Km (lM)
Activity
Enzyme Host (s1) H2O2 Mn2+ References

P. chrysosporium 410 107 42 [43]


Wild-type E. coli 450 96 52 [43]
S168W variant E. coli 260 110 110 [47]
Wild-type P. pastoris 168 58 [44]
P. eryngii MnPL2 118 (165 Umg1) 4 19 [128]
B. adusta MnP1 132 (180 Umg1) 10 17 [128]

yielded MnP variants highly resistant to 1 mM H2O2 containing peroxidases in the extracellular broth,
[51]. including three isoenzymes of MnP. Engineering of the
After MnP activity was identified in the white-rot fun- MnP3 isoenzyme produced five enzyme variants that
gus P. chrysosporium, the enzyme was also discovered in were overexpressed in E. coli [52]; after the inclusion
most of the wood lignin and litter-decaying basidiomy- bodies refolded, 1.4 mgL1 fermentation broth of
cetes. Phlebia radiata secretes several lignin heme- active MnP3 was purified (~ 7% of refolding efficiency).

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1197


Enzymatic degradation of lignin L. Pollegioni et al.

VP LRET from W164 in VP is similar to that observed in


P. chrysosporium LiP, with the main exception that
Reaction and properties
W164 exists in the form of a neutral radical [55]. The
VP (EC 1.11.1.16) is characterized by a broad sub- basic VP catalytic cycle proposed previously [56] was
strate preference and by sharing typical features of the then implemented to account for VA oxidation. This
MnP and LiP fungal peroxidase families. For example, model implies the presence of two enzyme forms,
both VP and LiP oxidize VA (the latter showing termed IB and IIB (in equilibrium with the classical
higher affinity and efficiency) but VP only oxidizes compounds I and II), characterized because one oxida-
Reactive Black 5 and other dyes. The VP activity has tion equivalent would be in the form of a tryptophan
been identified in the white-rot fungal genera Pleurotus radical, whose presence is required during oxidation of
(i.e. Pleurotus eryngii and Pleurotus ostreatus) and high-redox potential substrates.
Bjerkandera (i.e. Bjerkandera adusta) [53]. The catalytic efficiency of VP from P. eryngii on
The crystal structure of VP from P. eryngii (heterol- Mn2+ is similar to that of MnP: in contrast, VP is 10-
ogously expressed in E. coli W311O) has been solved fold less efficient than LiP in oxidizing VA [55,57] and
to 1.3 A and includes 11–12 helices, four disulfide is active on both phenolic and nonphenolic b-O-4 lig-
bridges, two structural Ca2+ sites, a heme pocket con- nin model dimers. P. eryngii VP degraded a 0.25 mM
taining the distal His47 and proximal His169 residues, guaiacylglycerol-b-guaiacyl ether and veratrylglycerol-
and a Mn2+ binding site (Fig. 5B) [54,55]. The Mn2+ b-guaiacyl ether solution with a 65% and 20% yield,
binding site is formed by one of the heme propionates respectively, in the presence of 0.1 mM H2O2 [58].
and the carboxylates of Glu36, Glu40 and Asp175 resi- Indeed, VP from P. ostreatus induced the oxidative
dues (Fig. 5C) and resembles that observed in the degradation of a nonphenolic b-O-4 lignin model
MnP. The substrate promiscuity of VP is the result of dimer [59] and a significant depolymerization of a syn-
a high redox potential (E00 > +1.4 V versus SHE) and thetic lignin was obtained in the presence of VA only.
the presence of different catalytic sites connected to
the heme pocket for low- and high-redox potential
Expression of P. eryngii VP
substrate oxidation [53]. Three putative long-range
electron-transfer pathways for oxidating high redox A genetically modified P. ostreatus strain with high VP
potential aromatic substrates have been identified, productivity, reaching at most 230 UL1 by using a
comprising the residues: (a) H232–D231 and the proxi- synthetic medium, was isolated [60]. By optimizing the
mal H169; (b) W164, L165 and the methyl-group C of culture conditions, 21 mgL1 of VP (7300 UL1)
the heme (the only pathway involved in VA oxida- could be produced (Table 3): the kinetic parameters
tion); and (c) P76, A77, D78 and the distal H47 for Mn2+, H2O2 and VA of P. ostreatus VP are
(Fig. 5C). Site-directed mutagenesis studies demon- reported in Table 4 [61]. Co-expression in P. crysospo-
strated that only the W164 pathway is involved in oxi- rium of a MnP and LiP (from the same microorgan-
dation of VA and Reactive Black 5 dye [55]. ism) and of VP vpl2 from P. eryngii enabled a high
The catalytic cycle of VP resembles the one reported level of extracellular production of this VP: 21 kUL1
for LiP (Fig. 2B): the enzyme catalyzes electron trans- culture on 2,6-dimethoxy-phenol as substrate [26].
fer from the substrate to form compound I and II The proper folding and expression of an active form
intermediates. The oxidation of Mn2+ and aromatic of recombinant VP is strongly hampered by its com-
substrates by direct electron transfer to the heme and plex structure (i.e. by the presence of four disulfide

Table 3. Expression yield of VP from different sources.

Expression level

Enzyme Host (UL1) (mgL1) References

P. ostreatus 230 0.6 [60]


P. ostreatus (modified strain) 7300 21 [61]
P. eryngii wild-type P. crysosporium 21 000 342 [26]
P. eryngii wild-type E. nidulans 0.4 [56]
P. eryngii wild-type S. cerevisiae 21 [64]
R4 variant S. cerevisiae 57 000 [64]
P. eryngii vpl2 E. coli 12.5 [63]

1198 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

Table 4. Kinetic parameters of different VPs.

VA Mn2+ ABTS H2O2

1 1 1
Enzyme Host Activity (s ) Km (lM) Activity (s ) Km (lM) Activity (s ) Km (lM) Activity (s1) Km (lM) References

P. eryngii 12.9 3000 118 19 [55]


Wild-type E. coli 8.0 2750 298 189 [55]
Wild-type E. coli (0.25 mUmg1)a (27.5 mUmg1)a [62]
Wild-type S. cerevisiae 54 45 220 540 135 51 [64]
R4 variant S. cerevisiae 75 120 365 56 490 200 [64]
2-1B variant S. cerevisiae 98 4300 850 34 1720 800 [64]
vpl2b E. coli 6.4 138 78 5.4 0.7 [63]
P. ostreatus 3.2 (4.4 Umg1) 31 (43.1 Umg1) 36 (50.8 Umg1) [69]
39 (54 Umg1) 207 (289 Umg1) 411 (574 Umg1) [59]
E. coli 13 (18 Umg1) 186 (259 Umg1) 142 (198 Umg1) [59]
Amycolatopsis E. coli 24 210 [69]
sp. 75iv2
(DyP2)

a
Specific activity.
b
After removing thioredoxin tag.

bridges, two Ca2+ ions and the heme prosthetic (yielding a 94-fold enhancement of the Km value for
group). An enzyme expression level resembling that Mn2+). Indeed, these evolved VP variants can operate
achieved for the homologous VP from P. eryngii was at higher H2O2 concentrations than the wild-type
obtained by employing Emericella nidulans or Aspergil- enzyme, with a significant improvement in activity
lus niger as hosts (~ 0.4 mgL1) (Table 3) [56]. Unfor- (from 135 s1 to 1720 s1 for the wild-type enzyme
tunately, the heterologous expression in E. coli cells and the 2-1B variant, respectively) (Table 4). At satu-
resulted in the formation of inclusion bodies and low rating H2O2 concentrations, the R4 and 2-1B variants
enzymatic activity (Table 4) [62]. Most recently, vpl2 (containing four and seven point substitutions, respec-
gene from P. eryngii was successfully overexpressed in tively) showed specific activities of 3530 Umg1 and
E. coli BL21(DE3) cells as a chimeric protein fused to 11.3 kUmg1, respectively, on ABTS. Under these
thioredoxin and by inducing protein expression by conditions, the expression of R4 VP reached
IPTG in the presence of 0.1 gL1 of hemin [63]. Over- ~ 57 kUL1 of culture broth (ABTS oxidation activ-
expressed VP-thioredoxin was soluble and was pro- ity) (Table 3), the highest value reported so far for a
duced with an overall yield of 12.5 mg pure peroxidase [64].
proteinL1 of fermentation broth. Specific activity To improve resistance to H2O2, VP from P. eryngii
increased > 10-fold after removing the tag: the main was recently the object of a rational design investiga-
kinetic parameters are reported in Table 4. VP from tion based on site-directed mutagenesis of residues
P. eryngii was more efficiently produced as a soluble which are easily oxidized (e.g. methionine) and close
and active enzyme in S. cerevisiae, reaching a secretion to heme and the H2O2-binding site. A number of sin-
level of ~ 21 mgL1 (Table 3) [64]. gle point variants, as well as double and triple point
variants, with improved stability over time in the pres-
ence of 4 mM H2O2 have been isolated [65].
Protein engineering of P. eryngii VP
To obtain a highly stable enzyme against temperature,
Dye-decolorizing peroxidases (DyPs)
alkaline pH and H2O2, P. eryngii VP was subjected to
six rounds of directed evolution using the yeast S. ce- Among the machinery involved in degrading lignin in
revisiae as expression host [64]. Among the selected the soil bacterium Rhodococcus jostii RHA1, three
enzyme variants, the R4 VP showed an improved DyPs have been identified recently [66]. DyPs, existing
affinity for 2,20 -azino-bis(3-ethylbenzothiazoline-6-sul- in a wide range of fungi and bacteria, contain a heme
phonic acid) (ABTS) (10-fold lower Km value than the prosthetic group, possess a ferredoxin-like fold (resem-
wild-type VP) and activity (i.e. kcat = 365 s1), thus bling that of chlorite dismutase) and have been classi-
enhancing its kcat/Km ratio (16-fold compared to the fied into four subfamilies (A–D). DyPs exhibit
wild-type VP) (Table 4). A similar behavior was peroxidase activity toward anthraquinones, ABTS, car-
observed for the thermostable 2-1B variant on ABTS; otenoids, methoxylated aromatics (i.e. VA) and lignin
however, Mn2+ binding was negatively affected model compounds: indeed, RHA1 degrades lignin and

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1199


Enzymatic degradation of lignin L. Pollegioni et al.

lignocellulose, producing a number of monocyclic phe- color; one type 2 and two type 3 copper ions form a
nolic compounds [67]. R. jostii RHA1 contains two trinuclear cluster (in the T2 and T3 sites) coordinated
DyPs: DypA, which is predicted to be in the periplasm by a highly conserved pattern of four histidine resi-
and is putatively involved in iron uptake, and DypB, dues. In the resting enzyme form, the four copper ions
which is responsible for lignin degradation. DypB are in the +2 oxidation state [24].
showed the greatest kinetic efficiency for ABTS (kcat/ The structure of laccase from Trametes versicolor
Km = 2000 M1s1) and also oxidized Mn2+ (kcat/ containing a full complement of coppers, the complete
Km = 25.1 M1s1), a reactivity that could enlarge its polypeptide chain and seven carbohydrate moieties
ability to degrade lignin-based compounds. Engineer- was solved previously [72] (Fig. 6A). All known fungal
ing of the Mn-binding pocket enabled a 14-fold laccases show a similar structure, consisting of three b-
improvement in the MnP activity of DyPB [68]. barrel sequential domains related to small blue copper
The strain Amylocolatopsis sp. 75iv2 also demon- proteins. The T1 site is located in the third domain
strated high capacity to modify lignin to produce acid- and the substrate binding site is located in a small,
precipitable polymeric lignin (as well as extracellular negatively-charged cavity close to the T1 site, whereas
phenol oxidation activity): DyP2, a member of C-type the T2–T3 site is placed between the first and the third
subfamily, was isolated and characterized [69]. DyP2 domain (Fig. 6B). A detailed observation shows the
has novel function for this family, with high activity presence of two channels: one for dioxygen diffusion
both as a peroxidase and MnP (kcat/Km in the range to reach the two type 3 copper ions, and the second
105 to 106 M1s1): MnP activity is close the same for release of water from the type 2 copper ion
order of magnitude as canonical MnPs and VPs which (Fig. 6B). Of note, the coordination distances in the
are involved in fungal lignin breakdown. As a peroxi- T1 site were proposed to affect the redox potentials.
dase, it demonstrates high activity against a broad The structure of T. versicolor laccase was solved at
range of peroxidase substrates, including RB5. DyP2 2.4 A resolution in complex with 2,5-xylidine, a weak
also demonstrates an oxidase reactivity, dependent on reducing substrate and the compound used as inducer
Mn2+, with a Km value similar to that for MnP activ- in the fungus culture [73]. The cavity accommodating
ity, and broadens its substrate scope to include the 2,5-xylidine is wide (10 9 10 9 20 A)  and can host
degradation of nonphenolic lignin model compounds substrates of various sizes. Several residues (F162,
[69]. Crystallographic studies show that a distinct Mn- L164, F265, F332 and F337) make hydrophobic inter-
binding site exists, located 15 A from the heme site. actions with the aromatic ring of the ligand and two
DyP2 was reported to rapidly break down lignin charged/polar residues interact with its amino group:
model dimers containing the b-O-4 linkage in the the conserved aspartate at position 206 and H458.
absence of redox mediators: the presence of a low The latter also coordinates the copper Cu1 that func-
redox potential phenolic site (DE = 0.6–0.8 V) in the tions as the primary electron acceptor: H458 repre-
substrate is mandatory for this activity [69]. sents the entrance door of this electron during its
transfer from the substrate to Cu1. The structure of
laccase–2,5-xylidine complex has been considered a
Laccase
suitable model for engineering the efficiency and spec-
Reaction and properties ificity of laccases.
On the other hand, bacterial laccase from Strepto-
Laccase (EC 1.10.3.2, benzenediol:oxygen oxidoreduc-
myces coelicolor A3(2) significantly differs from all lac-
tase) is a polyphenol oxidase that catalyzes the overall
cases studied so far: it consists of two domains and
reaction:
forms trimers, thus resembling the quaternary struc-
4 benzenediol þ O2
4 benzosemiquinone þ 2H2 O ture of nitrite reductases more than that of large lac-
cases (Fig. 6C) [74]. To produce a geometry of the
ð3Þ
active site similar to that of large laccases in S. coeli-
A recent and comprehensive review of laccases is color laccase, three trinuclear copper clusters are
provided elsewhere [70]. Fungal laccases are produced placed between domains 1 and 2 of each pair of neigh-
as intracellular and extracellular isoenzymes differing bor chains.
in oligomeric state (monomeric or dimeric) and glyco- Laccase possesses moderately low redox potentials
sylation level (10–45%) [71]. Laccase contains four (0.42–0.79 V) and can oxidize phenolic compounds
copper ions of three different types: type 1 copper in (i.e. polyphenols, methoxy-substituted phenols and
the T1 site shows a strong absorption band at around aromatic diamines, with Km values in the range 1–
600 nm, thus giving laccase its characteristic blue 10 mM) to the corresponding radicals [70]. The log

1200 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

A B

Fig. 6. (A) Ribbon diagram of laccase from


T. versicolor showing two channels
leading to the T2/T3 cluster (PDB code:
1GYC) [72]. (B) Structural details of the
laccase active site: the T1, T2 and T3 sites
are indicated. The T1 copper is
coordinated with Nd1 of an His and a
thiolate sulfur of a Cys as equatorial
ligands; the T2 copper is coordinated to
two His-Ne2 and one OH– yielding a
trigonal coplanar configuration; each T3 C
copper coordinates three His residues and
a bridging oxygen atom and shows a
distorted tetrahedral geometry. The arrows
mark the flow of substrates, electrons and
O2. Further details are provided elsewhere
[70]. (C) Comparison of the ribbon diagram
of laccase from T. versicolor organized in
three sequentially arranged domains (left)
and of trimeric bacterial laccase from
S. coelicolor A3(2) organized in two
domains A1 and B2 (PDB code: 3CG8)
(right) [74].

(kcat/Km) value increases with the redox potentials phenoxy radical formation, which leads to Ca oxida-
difference between the T1 site and the substrate donor tion, alkyl-aryl cleavage and Ca-Cb cleavage (Fig. 8A)
[75]. The substrate acceptance of laccases can be [77,78]. Indeed, laccases also oxidize nonphenolic
enlarged using mediator compounds: laccase slowly model compounds and b-1 lignin dimers in the
oxidizes ABTS, thus allowing the degradation of non- presence of a mediator such as ABTS, 1-hydroxybenzo-
phenolic lignin substrates (see below) [9] and the indi- triazole (HBT) and the natural compound 3-hydroxy-
rect oxidation of large molecules. Indeed, laccase does anthranilic acid [9]. The oxidation of a nonphenolic
not use hydrogen peroxide. b-O-4 model compound by employing laccase coupled
Regarding the catalytic cycle, the enzyme catalyzes a to HBT is reported in Fig. 8B, promoting aromatic ring
four single-electron substrate oxidation reaction and cleavage, Ca-Cb cleavage and Ca oxidation.
the enzyme is fully reduced by O2 through two consec- Interestingly, laccases are widely distributed in
utive two-electron steps (Fig. 7) [7,70]. An important higher plants (i.e. Rhus vernicifera, where laccase is
issue for laccase activity is the pH-dependence of the involved in the xylem tissue lignifications) [79], in fungi
activity. For phenolic substrates, the pH-dependence is (i.e. Pleurotus), and in a few bacteria (including Bacil-
bell-shaped: the activity increase with the pH is attrib- lus). Atypical laccases include those possessing two
uted to the pH-dependence decrease in redox poten- domains (i.e. from Botrytis cinerea and Tricholo-
tials of the phenol groups, whereas the decrease in ma giganteum), those showing a homodimeric (i.e.
activity with pH has been ascribed to hydroxide inhibi- Thapsia villosa, Phellinus ribis, Rhizoctonia solani, etc.)
tion of the trinuclear cluster [76]. The maximum of this or a tetrameric structure (i.e. Agaricus bisporus D621,
bell-shaped curve occurs at acidic pH values, although Podospora anserina, Aspergillus nidulans, etc.) and
a high activity at alkaline pH is a desired feature for those with unusual spectral properties (i.e. white or
industrial applications. yellow laccases; see below) [70].
Importantly, laccases subtract one electron from The constitutive, extracellular production of lac-
phenolic-OH groups of phenolic lignin model com- cases in basidiomycete fungi is low [80], although it
pounds forming phenoxy radicals, which often is significantly stimulated by specific inducers: for
undergo polymerization via radical coupling. This example, metals (Cu2+, Ca2+, Ag+ and Mn2+); aro-
reaction is also accompanied by demethylation, qui- matic or phenolic compounds related to lignin or lig-
none formation and ring cleavage [77]. Phenolic b-1 nin derivatives; ethanol (laccase production from
lignin models are degraded by laccase as a result of T. versicolor employing a medium containing

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1201


Enzymatic degradation of lignin L. Pollegioni et al.

Fig. 7. Proposed mechanism for the


reduction and reoxidation of the copper
sites in the catalytic cycle of laccase
[7,126]. The cycle starts with oxidation of
the substrate by the T1 copper
(Cu2+ ? Cu+), the primary electron
acceptor (the T1 site shows a relatively
high redox potential, ~ 700–800 mV). This
electron is transferred to the T2/T3
trinuclear site through a conserved His-
Cys-His motif (the rate-limiting step in
catalysis): a successive four-electron
oxidation (from different substrate
molecules) produced the fully reduced
enzyme form (step 1). The molecular
oxygen binds to the T2/T3 site and
converts to a peroxide intermediate with
transfer of two electrons from the T3
coppers (step 2). The peroxide
intermediate decays to an oxy radical
facilitated by electron transfer from the T2
copper and accelerated by pH decrease:
by 2-electron reductive cleavage of the
O-O bond a water molecule is released
(steps 3 and 4).

40 gL1 of ethanol was 2.6 kUL1, a 20-fold of 20 UL1 (8 mgL1) and the enzyme was hypergly-
increase compared to the basal level); and copper (it cosylated [81].
is important for the proper enzyme folding). Notably, The cDNA encoding for an extracellular laccase
when growing T. pubescens using glucose as the main from Streptomyces ipomoea CECT 3341 was cloned
carbon source, a significant laccase production and the protein overexpressed in E. coli [82]: it pos-
started when the sugar was consumed from the med- sessed a good stability at high pH (the enzyme
ium; in a medium containing 40 gL1 of glucose, retained its initial activity after 36 h of incubation at
10 gL1 of peptone and 2 mM Cu2+, 330 kU of lac- pH values between 5.0 and 9.0) and at high tempera-
caseL1 of fermentation broth was obtained. A ture, as well as resistance to high NaCl concentrations
review of laccase gene transcriptional regulation is and to several laccase inhibitors. The substrate prefer-
provided elsewhere [70]. ence of this laccase was pH-dependent: at alkaline pH,
a wide range of phenolic compounds were oxidized,
including the syringyl and guaiacil moieties derived
Recombinant expression
from lignin. At optimal pH, a Km of 0.4 mM and kcat
The cDNA coding for the lac1 gene isolated from Pyc- of 10 s1 were determined in the presence of ABTS.
noporus cinnabarinus was expressed in the methylo- T. versicolor, a typical white-rot fungus, secretes sev-
trophic yeast P. pastoris: after 10 days of cultivation, eral laccase isoenzymes. The LccB and LccC isoen-
laccase activity in the culture medium reached a value zymes were recently expressed in P. pastoris GS115,

1202 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

Fig. 8. (A) Three different pathways of


oxidation of phenolic b-1 lignin model
compounds by laccase [77,78]. (B) Three
different pathways of oxidation of
nonphenolic b-O-4 lignin model
compounds by laccase in the presence of
the HBT mediator [127].

reaching an expression level of 34 kUL1 [83]. The levels for P. pastoris X33 expressing the LccA isozyme
two isoenzymes show similar kinetic properties: the were lower (18 kUL1) [84]. Indeed, the lcc1 gene was
specific activity was 51 Umg1 and 63 Umg1 and cloned into the genome of Yarrowia lipolytica, reach-
the Km for ABTS was 0.43 mM and 0.29 mM for LccB ing an expression level of 250 UL1 of culture med-
and LccC, respectively. On the other hand, production ium [85].

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1203


Enzymatic degradation of lignin L. Pollegioni et al.

fungi Panus tigrinus and P. radiata catalyze the oxida-


Protein engineering of laccases from different sources
tion of nonphenolic b-1 lignin model compounds in
A directed evolution approach enabled the isolation of the absence of a mediator [88,89]. The first white lac-
three variants of Lcc1 laccase that yielded a four-fold case was purified from the basidiomycete P. ostreatus
increase in enzyme productivity (i.e. 1000 UL1) [85]. by Palmieri et al. [90]: this enzyme contains only one
The better variant (L185P/Q214K, named rM-4A) copper, two zinc atoms, and one iron atom in each
showed a 2.4-fold increase in the catalytic efficiency protein molecule. The laccase from Trametes hirsuta
toward ABTS (with a kcat and Km value of 210 s1 contains copper and manganese in a 3 : 1 ratio.
and 120 lM, respectively). The recombinant Lcc1 lac- A white laccase was recently purified from the deu-
case was used in dye decolorization. teromycete fungus Myrothecium verrucaria NF-05, a
By employing a combined approach of directed evo- strain producing up to 40 kU of laccaseL1 after
lution and rational design, the thermostable laccase 13 days of cultivation [91]. The highest activity was
from the basidiomycete PM1 (showing a moderate apparent at pH 4.0 and 30 °C and was significantly
high redox potential, i.e. > 0.7 V) was engineered by enhanced by adding Na+, Mn2+, Cu2+ and Zn2+ (it
Mate et al. [86]. The original signal sequence was was instead inhibited by dithiothreitol, NaN3 and hal-
replaced by the a-factor preproleader, and the protein ogen anions). The enzyme showed a higher affinity for
was expressed in S. cerevisiae. The OB-1 variant iso- ABTS (Km = 86 lM) than for other tested aromatic
lated after eight rounds of random mutagenesis substrates (i.e. cathechol) and a turnover rate of
showed high activity and stability in terms of tempera- 270 s1. This enzyme could be effectively used for dye
ture, pH range and organic solvents; a 34 000-fold decolorization.
enhancement in expression of the OB-1 activity
resulted from improvement of maximal activity and of
Bacterial laccases
the expression level (~ 8 mgL1). A turnover rate of
200 s1 and a Km value of 6.3 lM was apparent with Currently, only fungal laccases are used in biotechno-
the substrate ABTS. logical processes and, as stated above, most of the
Subsequently, an additional high redox potential lac- characterized enzymes have been isolated from fungi.
case isolated from P. cinnabarinus (named PcL), and To date, few bacterial laccases have been studied,
whose original signal sequence was replaced by the a- although rapid progress in genome analysis suggests
factor preproleader for expression in S. cerevisiae, was that these enzymes are widespread in bacteria. Here,
the object of protein engineering studies by the same the most representative enzyme is the CotA laccase
research group [87]. After six rounds of directed evolu- from Bacillus subtilis, an endospore coat protein show-
tion (by error-prone PCR and DNA shuffling), the ing a high thermostability [92,93] and for which the
best variant obtained (a-3PO) showed a 13.7-fold crystal structure has been determined [94].
higher kcat for ABTS than the wild-type (483 s1 ver- The 65-kDa CotA protein was purified to homoge-
sus 38 s1, respectively), improved kinetic parameters neity from an overproducing E. coli strain [93]. The
for various substrates (i.e. the catalytic efficiency for optimal enzymatic activity was found at pH 3.0 for
sinapic acid and 2,6-dimethoxyphenol increased 5.4- ABTS and at pH 7.0 for syringaldazine. Remarkably,
and 1.6-fold, respectively), an enzyme production of the purified enzyme showed an inactivation half-life of
300 UL1 and a secretion level of ~ 2 mgL1. More- ~ 2 h at 80 °C. The copper content of recombinant
over, the 7A9 enzyme variant was overexpressed in CotA from B. subtilis when produced in E. coli cells
A. niger: to facilitate the secretion process, the a-factor was strongly dependent on the presence of copper and
signal was replaced by the glucoamylase sequence from oxygen in the culture medium: in a copper-supple-
A. niger. A production level of 23 mgL1 was mented medium, the switch from aerobic to microaero-
achieved and the recombinant enzyme displayed bio- bic conditions produced an 80-fold increase in copper
chemical features similar to those of the variant protein incorporation and 20-fold enhancement in the
expressed in S. cerevisiae. kcat value (Table 5) [95].
A CotA laccase from Bacillus licheniformis was also
expressed in E. coli [96]. As shown in Table 5, the
White laccases
kinetic constants Km and kcat for ABTS were 6.5 lM
In recent years, it has been reported that atypical ‘yel- and 83 s1. The highest activity was observed at
low or white’ laccases (i.e. those lacking the peak at 85 °C; however, it is as thermostable as CotA from
around 600 nm in the absorption spectrum and in B. subtilis. Because the application of this laccase has
which the A280/A610 ratio is ~ 20) from the white-rot been hampered by expression levels (~ 26 mgL1), the

1204 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

Table 5. Kinetic parameters of different CotA bacterial laccases on ABTS as substrate.

Source Host kcat (s1) Km (lM) kcat/Km Reference

B. subtilis E. coli 17 106 0.16 [93]


B. subtilis (microaerobic) E. coli 322 124 2.60 [95]
B. licheniformis E. coli 83 6.5 12.80 [96]
B. pumilus E. coli 290 80 3.63 [98]

CotA double variant K316N/D500G was produced Cleaving b-O-4 ether bond
that reached an 11.4-fold higher level of expression
An extracellular b-etherase capable of cleaving the
than the wild-type enzyme [97]. Indeed, this laccase
phenolic b-O-4 lignin model dimer (i.e. the guaiacyl-
variant is also more efficient than the wild-type
glycerol b-guaiacyl ether) has been isolated from the
enzyme in converting ferulic acid (21% conversion ver-
fermentation broth of an ascomycete of the genus
sus 14%) and in decolorizing a range of industrial
Chaetomium, in particular, by the newly isolated fun-
dyes.
gus 2BW-1 (approximately 250 UL1) [102]. This 65-
Recently, a CotA laccase from Bacillus pumilus was
kDa enzyme catalyzes the addition of two molecules
also expressed and purified from E. coli cells [98]. The
of H2O at Ca and Cb positions and the cleavage of
recombinant enzyme showed a high activity (kcat for
the b-aryl ether bond of the model compound. The
ABTS was 290 s1) (Table 5), a pH optimum of 4.0, a
proposed reaction mechanism might proceed through
temperature optimum at around 70 °C and a wide
an intermediate quinine methide in which b-aryl ether
substrate specificity.
bond scission occurs, as shown in Fig. 9A. This is con-
The laccase from S. coelicolor shows also interesting
sistent with the observation that the b-aryl ether cleav-
features since it is relatively small in size (~ 32 kDa)
age enzyme requires a hydroxyl group and a Ca
and possesses a high activity at alkaline pH values
alcohol structure. A water molecule attacks the Ca
[99]: it is representative of a new enzyme family, the
position in the quinonemethide, the b-aryl ether link-
two-domain laccases, lacking the second domain
age is cleaved, and an additional water molecule
involved in the substrate binding (Fig. 6C). The
attacks the Ca position. As a result, the guaiacylglyc-
recombinant enzyme expressed in E. coli was highly
erol b-O-guaiacyl is converted to guaiacylglycerol and
stable (it retained activity as dimer in denaturating gels
guaiacol.
after boiling and SDS treatment), showed an optimal
Subsequently, an additional b-O-4 aryl-ether cleaving
kinetic efficiency on 2,6-dimethoxyphenol at pH 9.4
enzyme system from the protobacterium Sphingobium
(Km of ~ 0.5 mM and kcat of 10 s1), and was quite
sp. SYK-6 was identified and characterized [103]. This
resistant under various conditions (i.e. in the presence
reaction is part of a three-step process (shown in
of 1 mM NaN3). Laccase from S. ipomoea was also
Fig. 9B): the three proteins involved, LigD (a Ca-dehy-
overexpressed in E. coli BL21(DE3) as host and used
drogenase), LigF (a b-etherase) and LigG (a glutathione
for bleaching of kraft pulps [100]. Interestingly, an effi-
lyase belonging to the omega class of glutathione trans-
cient expression system was set up in Streptomyces livi-
ferase class and whose structure was solved) [104], were
dans (homologous expression host of this bacterial
successfully expressed and purified in E. coli [105]. This
laccase), reaching an expression level of 350 mgL1,
enzymatic system was demonstrated to specifically
the highest production yield reported so far for a bac-
cleave the b-aryl ethers of a model lignin substrate, the 1-
terial laccase [101]. The secreted enzyme showed an
(4-hydroxy-3-methoxy-phenyl)-2-(2-methoxyphenolxy)-
oxidative activity on a wide pH range, depending on
1,3-propanediol: LigD catalyzes the NAD+-dependent
the tested substrate.
oxidation of Ca of lignin substrate, LigF cleaves the
intermediate with attachment of glutathione at the Cb
Additional enzymatic delignification activities position; glutathione is then oxidized by LigG and the
product is released. Altogether, the enzyme cascade Lig-
In the past decade, several new enzymatic approaches
DFG catalyzes the reductive cleavage of the ether bond
for delignification have been reported, based on the
and the oxidation of the secondary alcohol via reduction
dermination of a number of novel biochemical path-
of NAD+ and oxidation of two molecules of reduced
ways. It has been suggested that these accessory
glutathione. Here, the co-substrates NAD+ and
enzymes enhance lignin degradation as a result of the
glutathione can be regenerated by introducing the
well-known lignolytic enzymes (see above).

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1205


Enzymatic degradation of lignin L. Pollegioni et al.

Fig. 9. Cleavage of b-O-4 ether bond. (A) The mechanism of catalysis of the b-aryl ether cleavage enzyme from the fungus Chaetomium
2BW-1 as proposed previously [102]. An asterisk indicates 18O labeling in water and reaction products. (B) Reaction pathway of the LigDFG
enzyme system [105].

NAD+-dependent glutathione reductase from Allochro- subsequently cleaved by the C-C hydrolase LigY, to
matium vinosum, allowing for a self-sufficient balanced form 5-carboxyvanillic acid and 4-carboxy-2-hydroxy-
enzymatic process. Approximately 35% of a model pentadienoic acid. Two decarboxylase enzymes (named
compound was cleaved after 7 days of reaction [103]. LigW and LigW2) have been identified that subse-
The ability of this enzymatic cascade to release lignin quently convert 5-carboxyvanillic acid into the valu-
monomers from complex lignin structures (from differ- able compound vanillic acid.
ently prepared real lignin substrates) was also reported
[105]. Notably, LigE and LigP glutathione reductase
O-Demethylation enzyme systems
catalyzes a reaction similar to that of LigF on a different
isomeric guaiacylglycerol-b-guaiacyl ether intermediate S. paucimobilis SYK-6 was reported to grow using the
produced by LigD in the same degradation pathway lignin-derived compounds vanillate and syringate as
[106]. the sole source of carbon and energy [109]. In detail,
vanillate and syringate are converted to protocatechu-
ate and 3-O-methylgallate by the O-demethylases
Biphenyl bond cleavage enzyme systems
LigM and DesA, respectively [110]. These tetrahydrof-
In addition to b-O-4 aryl-ether linkage, the biphenyl olate-dependent enzymes were expressed in E. coli,
linkage is also a main component of lignin structure, showing that vanillate and 3-O-methylgallate were also
commonly occurring between two guaiacyl units substrates for DesA (but with a very low relative activ-
(Fig. 1). Oxidative cleavage of the C3 side chain ity compared to syringate) and that vanillate and 3-O-
attached to the biphenyl group would generate 2,20 -di- methylgallate were converted to protocatechuate and
hydroxy-3,30 -dimethoxy-5,50 -dicarboxy-biphenyl, which gallate, respectively, by LigM enzyme with similar,
can be utilized for the growth of Sphingomonas pauc- specific activities (whereas syringate was not a sub-
imobilis SYK-6 [107,108]. As shown in Fig. 10, deme- strate for LigM) [111].
thylation of one methoxy group was reported to be
catalyzed by the non-heme, iron-dependent demethy-
General oxidative activities
lase enzyme LigX; the catecholic product of LigX is
the substrate for oxidative cleavage by the extradiol Among the fungal extracellular enzymes proposed to
dioxygenase LigZ, and the ring fission product is be involved in lignin degradation, we can include

1206 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

Fig. 10. The pathway for the degradation of the lignin biphenyl component in S. paucimobilis by LigX, LigZ and LigY [108].

general oxidases generating H2O2, which provide the lized enzyme was separated and the remnant wood
hydrogen peroxide required by peroxidases, and myce- was obtained as a yellowish oil after dimethylcarbon-
lium-associated dehydrogenases [112]. The H2O2-gener- ate evaporation, suggesting a successful delignification
ating oxidases include aryl-alcohol oxidase process. ESI-MS analysis demonstrated the production
(EC 1.1.3.7) (found in various fungi; e.g. P. eryngii) of small oligomers from both beech wood and conife-
and glyoxal oxidase (identified in P. chrysosporium). In ryl alcohol-derived synthetic lignin.
addition, quinone reductases (EC 1.6.5.5) are also Treating unbleached kraft pulp with xylanase
involved in lignin degradation by fungi [113]. More- (EC 3.2.1.8) did not release significant amounts of lig-
over, cellobiose dehydrogenase (EC 1.1.99.18), an nin; however, the enzyme facilitated extraction of the
enzyme produced by many different fungi under cellu- lignin fraction because it removed the linkage between
lolytic conditions, is also involved in lignin degrada- xylan and lignin [122].
tion in the presence of H2O2 and chelated iron ions
[114]. It was proposed to act on the reduction of qui-
Conclusions
nones, which can be used by ligninolytic enzymes or in
support of a MnP reaction [115]. Lignin represents a feedstock of enormous potential
The I.2.A extradiol dioxygenase subfamily appears for producing low-molecular-weight chemicals. Bec-
to be of particular importance for breaking down ause of its complex structure, considerable effort has
monocyclic aromatic compounds [116]. Regarding the been devoted to understanding the major natural path-
catechol 2,3-dioxygenase (EC 1.13.11.2) subfamily, dif- ways and enzymes involved in lignin degradation. It is
ferences in kinetic properties have been reported for now clear that, in nature, lignin degradation is a
variants differing in a few residues [117] (e.g. in the multi-enzymatic process involving an array of acces-
residue at position 218). The catechol 2,3-dioxygenase sory enzymes in addition to the four major ligninolytic
variant possessing His218 showed the highest activity oxidases. To reach such an ambitious goal, further
on catechol, whereas the one possessing Tyr218 investigations will be aimed at: (a) defining the role
showed the highest activity on 4-methylcatechol (and and interaction of enzyme combinations in lignin deg-
3-methylcatechol) and a low Km for all catecholic sub- radation; (b) determining the role of small molecule
strates [118]. mediators (both natural and synthetic) in the catalytic
A recently identified type of hydrolase, the so-called mechanism of the ligninolytic enzymes, as well as in
perhydrolase, has been proposed as useful biocatalyst the overall process of lignin degradation; (c) reconsti-
to produce peroxycarboxylic acids in situ (by using tuting in vitro fungal/bacterial ligninolytic systems by
H2O2 and carboxylic acids in aqueous media) and sub- means of purified enzyme components; and (d) com-
sequently employed for delignification [119,120]. Other bining enzyme treatment with other pretreatment
hydrolases, mostly lipases (EC 3.1.1.3), can be used as methods (the ‘combined’ approach).
excellent biocatalysts for in situ peracid formation in Of note, ligninolytic enzymes find applications in a
nonaqueous media, starting from carboxylic acids and number of industrial processes other than biorefinery
diluted H2O2. Indeed, an immobilized lipase B from (e.g. laccase is used for clarifying musts and wines,
Candida antartica was successfully employed in a per- waste water treatment, biostaining of denim jeans,
acid-mediated lignocellulosic delignification process in assaying the total antioxidant amount in blood sam-
a nonaqueous solvent: here, dimethylcarbonate was ples, and as biosensors, etc.). A recent review discuss-
used both as a solvent and as an acyl-donor reagent ing the application of laccase as a potential
for the lipase-mediated oxidation [121]. After 9 h of pretreatment strategy for biofuels production is pro-
incubation of beech wood lignin at 80 °C, the immobi- vided elsewhere [123]. Similarly, the development of

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1207


Enzymatic degradation of lignin L. Pollegioni et al.

novel ligninolytic enzymes will pave the way for addi- mediators with lignin model compounds. Biochim
tional, unrelated industrial uses. Biophys Acta 1379, 381–390.
10 Crestini C, Jurasek L & Argyropoulos DS (2003) On
the mechanism of the laccase – mediator system in the
Acknowledgements oxidation of lignin. Chem Eur J 9, 5371–5378.
This work was supported by grants from FP7 project 11 Ca~nas AI & Camarero S (2010) Laccases and their
ValorPlus (Project number 613802) and Fondazione natural mediators: biotechnological tools for
Cariplo – Regione Lombardia project Biorefill (Project sustainable eco-friendly processes. Biotechnol Adv 28,
694–705.
ID 42611813) to LP. FT is a student of the PhD
12 Lange H, Decina S & Crestini C (2013) Oxidative
school in ‘Biotechnology, Biosciences and Surgical
upgrade of lignin – recent routes reviewed. Eur Polym
Technologies’ at Universit
a degli studi dell’Insubria.
J 49, 1151–1173.
13 Asgher M, Shahid M, Kamal S & Iqbal HMN (2014)
Author contributions Recent trends and valorization of immobilization
strategies and ligninolytic enzymes by industrial
ER and FT researched bibliography; ER and FT drew biotechnology. J Mol Catal B-Enzym 101, 56–66.
figures; ER and LP wrote the paper; LP reviewed the 14 Welinder KG, Mauro JM & Norskov-Lauritsen L
manuscript. All the authors gave final approval. (1992) Structure of plant and fungal peroxidases.
Biochem Soc Trans 20, 337–340.
References 15 Wang W & Wen X (2009) Expression of lignin
peroxidase H2 from Phanerochaete chrysosporium by
1 Mai C, Majcherczyk A & Huttermann A (2000) multi-copy recombinant Pichia strain. J Environ Sci
Chemo-enzymatic synthesis and characterization of 21, 218–222.
graft copolymers from lignin and acrylic compounds. 16 Valli K, Wariishi H & Gold MH (1990) Oxidation of
Enzyme Microb Technol 27, 167–175. monomethoxylated aromatic compounds by lignin
2 FitzPatrick M, Champagne P, Cunningham MF & peroxidase: role of veratryl alcohol in lignin
Whitney RA (2010) A biorefinery processing biodegradation. Biochemistry 29, 8535–8539.
perspective: treatment of lignocellulosic materials for 17 Hammel KE, Tien M, Kalyanaraman B & Kirk TK
the production of value-added products. Biores (1985) Mechanism of oxidative C alpha-C beta cleavage
Technol 101, 8915–8922. of a lignin model dimer by Phanerochaete chrysosporium
3 Pandey MP & Kim CS (2011) Lignin depolymerization ligninase. Stoichiometry and involvement of free
and conversion: a review of thermochemical methods. radicals. J Biol Chem 260, 8348–8353.
Chem Eng Technol 34, 29–41. 18 Lundell T, Wever R, Floris R, Harvey P, Hatakka A,
4 Zhou CH, Xia X, Lin CX, Tong DS & Beltramini J Brunow G & Schoemaker H (1993) Lignin peroxidase
(2011) Catalytic conversion of lignocellulosic biomass L3 from Phlebia radiata. Pre-steady-state and steady-
to fine chemicals and fuels. Chem Soc Rev 40, state studies with veratryl alcohol and a non-phenolic
5588–5617. lignin model compound 1-(3,4-dimethoxyphenyl)-2-(2-
5 Nanayakkara S, Patti AF & Saito K (2014) methoxyphenoxy)propane-1,3-diol. Eur J Biochem 211,
Chemical depolymerization of lignin involving the 391–402.
redistribution mechanism with phenols and 19 Baciocchi E, Fabbri C & Lanzalunga O (2003) Lignin
repolymerization of depolymerized products. Green peroxidase-catalyzed oxidation of nonphenolic trimeric
Chem 16, 1897–1903. lignin model compounds: fragmentation reactions in
6 Dashtban M, Schraft H, Syed TA & Qin W (2010) the intermediate radical cations. J Org Chem 68,
Fungal biodegradation and enzymatic modification of 9061–9069.
lignin. Int J Biochem Mol Biol 1, 36–50. 20 Choinowski T, Blodig W, Winterhalter KH & Piontek
7 Wong DWS (2009) Structure and action mechanism of K (1999) The crystal structure of lignin peroxidase at
ligninolytic enzymes. Appl Biochem Biotech 157, 1.70 A resolution reveals a hydroxy group on the
174–209. cbeta of tryptophan 171: a novel radical site formed
8 Forney LJ, Reddy CA, Tien M & Aust SD (1982) The during the redox cycle. J Mol Biol 286, 809–827.
involvement of hydroxyl radical derived from hydrogen 21 Edwards SL, Raag R, Wariishi H, Gold MH & Poulos
peroxide in lignin degradation by the white rot fungus TL (1993) Crystal structure of lignin peroxidase.
Phanerochaete chrysosporium. J Biol Chem 257, PNAS 90, 750–754.
11455–11462. 22 Poulos TL, Edwards SL, Wariishi H & Gold MH
9 Bourbonnais R, Leech D & Paice MG (1998) (1993) Crystallographic refinement of lignin peroxidase
Electrochemical analysis of the interactions of laccase at 2 A. J Biol Chem 268, 4429–4440.

1208 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

23 Blodig W, Smith AT, Doyle WA & Piontek K (2001) 36 Paszczynski A, Huynh VB & Crawford R (1986)
Crystal structures of pristine and oxidatively processed Comparison of ligninase-I and peroxidase-M2 from
lignin peroxidase expressed in Escherichia coli and of the white-rot fungus Phanerochaete chrysosporium.
the W171F variant that eliminates the redox active Arch Biochem Biophys 244, 750–765.
tryptophan 171. Implications for the reaction 37 Wariishi H, Valli K & Gold MH (1991) In vitro
mechanism. J Mol Biol 305, 851–861. depolymerization of lignin by manganese peroxidase of
24 Doyle WA, Blodig W, Veitch NC, Piontek K & Smith Phanerochaete chrysosporium. Biochem Biophys Res
AT (1998) Two substrate interaction sites in lignin Commun 176, 269–275.
peroxidase revealed by site-directed mutagenesis. 38 Hammel KE & Cullen D (2008) Role of fungal
Biochemistry 37, 15097–150105. peroxidases in biological ligninolysis. Curr Opin Plant
25 Gelpke MDS, Lee J & Gold MH (2002) Lignin Biol 11, 349–355.
peroxidase oxidation of veratryl alcohol: Effects of the 39 Sundaramoorthy M, Kishi K, Gold MH & Poulos TL
mutants H82A, Q222A, W171A, and F267L. (1997) Crystal structures of substrate binding site
Biochemistry 41, 3498–3506. mutants of manganese peroxidase. J Biol Chem 272,
26 Coconi-Linares N, Maga~ na-Ortız D, Guzman-Ortiz 17574–17580.
DA, Fernandez F, Loske AM & G omez-Lim MA 40 Sundaramoorthy M, Kishi K, Gold MH & Poulos TL
(2014) High-yield production of manganese peroxidase, (1994) The crystal structure of manganese peroxidase
lignin peroxidase, and versatile peroxidase in from Phanerochaete chrysosporium at 2.06-A
Phanerochaete chrysosporium. Appl Microbiol resolution. J Biol Chem 269, 32759–32767.
Biotechnol 98, 9283–9294. 41 Gelpke MD, Youngs HL & Gold MH (2000) Role
27 Nie G, Reading NS & Aust SD (1998) Expression of of arginine 177 in the MnII binding site of
the lignin peroxidase H2 gene from Phanerochaete manganese peroxidase. Studies with R177D, R177E,
chrysosporium in Escherichia coli. Biochem Biophys Res R177N, and R177Q mutants. Eur J Biochem 267,
Commun 249, 146–150. 7038–7045.
28 Gelpke MDS, Mayfield-Gambill M, Cereghino GPL & 42 Sundaramoorthy M, Youngs HL, Gold MH & Poulos
Gold MH (1999) Homologous expression of TL (2005) High-resolution crystal structure of
recombinant lignin peroxidase in Phanerochaete manganese peroxidase: substrate and inhibitor
chrysosporium. Appl Environ Microbiol 65, 1670–1674. complexes. Biochemistry 44, 6463–6470.
29 Wang HK, Lu FP, Sun WF & Du LX (2004) 43 Whitwam R & Tien M (1996) Heterologous expression
Heterologous expression of lignin peroxidase of and reconstitution of fungal Mn peroxidase. Arch
Phanerochaete chrysosporium in Pichia methanolica. Biochem Biophys 333, 439–446.
Biotechnol Lett 26, 1569–1573. 44 Gu L, Lajoie C & Kelly C (2003) Expression of a
30 Xie J, Feng L, Xu N, Zhu GH, Yang J, Xu XL & Fu Phanerochaete chrysosporium manganese peroxidase
SY (2007) Studies on the fusion of ligninolytic enzyme gene in the yeast Pichia pastoris. Biotechnol Progr 19,
cDNAs and their expression. Bioresources 2, 598–604. 1403–1409.
31 Hammel KE, Jensen KA Jr, Mozuch MD, Landucci 45 Jiang F, Kongsaeree P, Charron R, Lajoie C, Xu H,
LL, Tien M & Pease EA (1993) Ligninolysis by a Scott G & Kelly C (2008) Production and separation
purified lignin peroxidase. J Biol Chem 268, of manganese peroxidase from heme amended yeast
12274–12281. cultures. Biotechnol Bioeng 99, 540–549.
32 Haemmerli SD, Leisola MSA & Fiechter A (1986) 46 Timofeevski SL, Nie G, Reading NS & Aust SD
Polymerisation of lignins by ligninases from (1999) Addition of veratryl alcohol oxidase activity to
Phanerochaete chrysosporium. FEMS Microbiol Lett manganese peroxidase by site-directed mutagenesis.
35, 33–36. Biochem Biophys Res Commun 256, 500–504.
33 Brunow G (2005) Methods to reveal the structure of 47 Timofeevski SL, Nie G, Reading NS & Aust SD
lignin. In Lignin, Humic Substances and Coal. (2000) Substrate specificity of lignin peroxidase and a
Biopolymers Online (Hofrichter M & Steinb€ uchel A, S168W variant of manganese peroxidase. Arch
eds), pp. 89–116. Wiley-VHC, Weinheim. Biochem Biophys 373, 147–153.
34 Ghaffar SH & Fan M (2013) Structural analysis for 48 Sutherland GR & Aust SD (1996) The effects of
lignin characteristics in biomass straw. Biomass calcium on the thermal stability and activity of
Bioenergy 57, 264–279. manganese peroxidase. Arch Biochem Biophys 332,
35 Glenn JK & Gold MH (1985) Purification and 128–134.
characterization of an extracellular Mn(II)-dependent 49 Nie G, Reading NS & Aust SD (1999) Relative
peroxidase from the lignin-degrading basidiomycete, stability of recombinant versus native peroxidases from
Phanerochaete chrysosporium. Arch Biochem Biophys Phanerochaete chrysosporium. Arch Biochem Biophys
242, 329–341. 365, 328–334.

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1209


Enzymatic degradation of lignin L. Pollegioni et al.

50 Reading NS & Aust SD (2000) Engineering a disulfide 61 Tsukihara T, Honda Y, Sakai R & Watanabe T (2006)
bond in recombinant manganese peroxidase results in Exclusive overproduction of recombinant versatile
increased thermostability. Biotechnol Progr 16, peroxidase MnP2 by genetically modified white rot
326–333. fungus, Pleurotus ostreatus. J Biotechnol 126, 431–439.
51 Miyazaki C & Takahashi H (2001) Engineering of the 62 Mohorcic M, Bencina M, Friedrich J & Jerala R
H2O2-binding pocket region of a recombinant (2009) Expression of soluble versatile peroxidase of
manganese peroxidase to be resistant to H2O2. FEBS Bjerkandera adusta in Escherichia coli. Bioresour
Lett 509, 111–114. Technol 100, 851–858.
52 Ufot UF & Akpanabiatu MI (2012) An engineered 63 Bao X, Liu A, Lu X & Li JJ (2012) Direct over-
Phlebia radiata manganese peroxidase: expression, expression, characterization and H2O2 stability study
refolding, purification and preliminary of active Pleurotus eryngii versatile peroxidase in
characterization. Am J Mol Biol 2, 359–370. Escherichia coli. Biotechnol Lett 34, 1537–1543.
53 Hofrichter M, Ullrich R, Pecyna MJ, Liers C & 64 Garcia-Ruiz E, Gonzalez-Perez D, Ruiz-Duenas FJ,
Lundell T (2010) New and classic families of secreted Martinez AT & Alcalde M (2012) Directed evolution
fungal heme peroxidases. Appl Microbiol Biotechnol 87, of a temperature-, peroxide- and alkaline pH-tolerant
871–897. versatile peroxidase. Biochem J 441, 487–498.
54 Ruiz-Duenas FJ, Morales M, Perez-Boada M, 65 Bao X, Huang X, Lu X & Li JJ (2014) Improvement
Choinowski T, Martinez MJ, Piontek K & Martinez of hydrogen peroxide stability of Pleurotus eryngii
AT (2007) Manganese oxidation site in Pleurotus versatile ligninolytic peroxidase by rational protein
eryngii versatile peroxidase: a site-directed mutagenesis, engineering. Enz Microb Tech 54, 51–58.
kinetic, and crystallographic study. Biochemistry 46, 66 Roberts JN, Singh R, Grigg JC, Murphy MEP, Bugg
66–77. TDH & Eltis LD (2011) Characterization of dye-
55 Perez-Boada M, Ruiz-Duenas FJ, Pogni R, Basosi R, decolorizing peroxidases from Rhodococcus jostii
Choinowski T, Martinez MJ, Piontek K & Martinez RHA1. Biochemistry 50, 5108–5119.
AT (2005) Versatile peroxidase oxidation of high 67 Ahmad M, Taylor CR, Pink D, Burton K, Eastwood
redox potential aromatic compounds: site-directed D, Bending GD & Bugg TD (2010) Development of
mutagenesis, spectroscopic and crystallographic novel assays for lignin degradation: comparative
investigation of three long-range electron transfer analysis of bacterial and fungal lignin degraders. Mol
pathways. J Mol Biol 354, 385–402. BioSyst 6, 815–821.
56 Ruiz-Duenas FJ, Martinez MJ & Martinez AT (1999) 68 Singh R, Grigg JC, Qin W, Kadla JF, Murphy ME &
Heterologous expression of Pleurotus eryngii Eltis LD (2013) Improved manganese-oxidizing activity
peroxidase confirms its ability to oxidize Mn(2+) and of DypB, a peroxidase from a lignolytic bacterium.
different aromatic substrates. Appl Environ Microbiol ACS Chem Biol 8, 700–706.
65, 4705–4707. 69 Brown ME, Barros T & Chang MC (2012)
57 Tien M, Kirk TK, Bull C & Fee JA (1986) Steady- Identification and characterization of a multifunctional
state and transient-state kinetic studies on the dye peroxidase from a lignin-reactive bacterium. ACS
oxidation of 3,4-dimethoxybenzyl alcohol catalyzed by Chem Biol 7, 2074–2081.
the ligninase of Phanerocheate chrysosporium Burds. J 70 Giardina P, Faraco V, Pezzella C, Piscitelli A,
Biol Chem 261, 1687–1693. Vanhulle S & Sannia G (2010) Laccases: a never-
58 Caramelo L, Martinez MJ & Martinez AT (1999) A ending story. Cell Mol Life Sci 67, 369–385.
search for ligninolytic peroxidases in the fungus 71 Thurston CF (1994) The structure and function of
Pleurotus eryngii involving a-keto-c-thiomethylbutyric fungal laccases. Microbiol-Sgm 140, 19–26.
acid and lignin model dimers. Appl Environ Microbiol 72 Piontek K, Antorini M & Choinowski T (2002)
65, 916–922. Crystal structure of a laccase from the fungus
59 Fernandez-Fueyo E, Ruiz-Duenas FJ, Martinez MJ, Trametes versicolor at 1.90-A resolution containing a
Romero A, Hammel KE, Medrano FJ & Martinez AT full complement of coppers. J Biol Chem 277,
(2014) Ligninolytic peroxidase genes in the oyster 37663–37669.
mushroom genome: heterologous expression, molecular 73 Bertrand T, Jolivalt C, Briozzo P, Caminade E, Joly
structure, catalytic and stability properties, and lignin- N, Madzak C & Mougin C (2002) Crystal structure of
degrading ability. Biotechnol Biofuels 7, 2. a four-copper laccase complexed with an arylamine:
60 Tsukihara T, Honda Y & Watanabe T (2006) insights into substrate recognition and correlation with
Molecular breeding of white rot fungus Pleurotus kinetics. Biochemistry 41, 7325–7333.
ostreatus by homologous expression of its versatile 74 Skalova T, Dohnalek J, Ostergaard LH, Ostergaard
peroxidase MnP2. Appl Microbiol Biotechnol 71, PR, Kolenko P, Duskova J, Stepankova A & Hasek J
114–120. (2009) The structure of the small laccase from

1210 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

Streptomyces coelicolor reveals a link between laccases Engineering platforms for directed evolution of laccase
and nitrite reductases. J Mol Biol 385, 1165–1178. from Pycnoporus cinnabarinus. Appl Environ Microbiol
75 Xu F (1996) Catalysis of novel enzymatic iodide 78, 1370–1384.
oxidation by fungal laccase. Appl Biochem Biotech 59, 88 Leontievsky A, Myasoedova N, Pozdnyakova N &
221–230. Golovleva L (1997) ‘Yellow’ laccase of Panus tigrinus
76 Xu F (1997) Effects of redox potential and hydroxide oxidizes non-phenolic substrates without electron-
inhibition on the pH activity profile of fungal laccases. transfer mediators. FEBS Lett 413, 446–448.
J Biol Chem 272, 924–928. 89 Leontievsky AA, Vares T, Lankinen P, Shergill JK,
77 Kawai S, Nakagawa M & Ohashi H (1999) Aromatic Pozdnyakova NN, Myasoedova NM, Kalkkinen N,
ring cleavage of a non-phenolic beta-O-4 lignin model Golovleva LA, Cammack R, Thurston CF et al.
dimer by laccase of Trametes versicolor in the presence (1997) Blue and yellow laccases of ligninolytic fungi.
of 1-hydroxybenzotriazole. FEBS Lett 446, 355–358. FEMS Microbiol Lett 156, 9–14.
78 Kawai S, Asukai M, Ohya N, Okita K, Ito T & 90 Palmieri G, Giardina P, Bianco C, Scaloni A, Capasso
Ohashi H (1999) Degradation of a non-phenolic beta- A & Sannia G (1997) A novel white laccase from
O-4 substructure and of polymeric lignin model Pleurotus ostreatus. J Biol Chem 272, 31301–31307.
compounds by laccase of Coriolus versicolor in the 91 Zhao D, Zhang X, Cui D & Zhao M (2012)
presence of 1-hydroxybenzotriazole. FEMS Microbiol Characterisation of a novel white laccase from the
Lett 170, 51–57. deuteromycete fungus Myrothecium verrucaria NF-05
79 LaFayette PR, Eriksson KE & Dean JF (1999) and its decolourisation of dyes. PLoS ONE 7, e38817.
Characterization and heterologous expression of 92 Hullo MF, Moszer I, Danchin A & Martin-Verstraete
laccase cDNAs from xylem tissues of yellow-poplar I (2001) CotA of Bacillus subtilis is a copper-dependent
(Liriodendron tulipifera). Plant Mol Biol 40, 23–35. laccase. J Bacteriol 183, 5426–5430.
80 Bollag JM & Leonowicz A (1984) Comparative studies 93 Martins LO, Soares CM, Pereira MM, Teixeira M,
of extracellular fungal laccases. Appl Environ Microbiol Costa T, Jones GH & Henriques AO (2002) Molecular
48, 849–854. and biochemical characterization of a highly stable
81 Otterbein L, Record E, Longhi S, Asther M & bacterial laccase that occurs as a structural component
Moukha S (2000) Molecular cloning of the cDNA of the Bacillus subtilis endospore coat. J Biol Chem
encoding laccase from Pycnoporus cinnabarinus I-937 277, 18849–18859.
and expression in Pichia pastoris. Eur J Biochem 267, 94 Enguita FJ, Martins LO, Henriques AO & Carrondo
1619–1625. MA (2003) Crystal structure of a bacterial endospore
82 Molina-Guijarro JM, Perez J, Munoz-Dorado J, coat component. A laccase with enhanced
Guillen F, Moya R, Hernandez M & Arias ME (2009) thermostability properties. J Biol Chem 278,
Detoxification of azo dyes by a novel pH-versatile, 19416–19425.
salt-resistant laccase from Streptomyces ipomoea. Int 95 Durao P, Chen Z, Fernandes AT, Hildebrandt P,
Microbiol 12, 13–21. Murgida DH, Todorovic S, Pereira MM, Melo EP &
83 Li Q, Ge L, Cai J, Pei J, Xie J & Zhao L (2014) Martins LO (2008) Copper incorporation into
Comparison of two laccases from Trametes versicolor recombinant CotA laccase from Bacillus subtilis:
for application in the decolorization of dyes. characterization of fully copper loaded enzymes. J Biol
J Microbiol Biotechnol 24, 545–555. Inorg Chem 13, 183–193.
84 Li Q, Pei J, Zhao L, Xie J, Cao F & Wang G (2014) 96 Koschorreck K, Richter SM, Ene AB, Roduner E,
Overexpression and characterization of laccase from Schmid RD & Urlacher VB (2008) Cloning and
Trametes versicolor in Pichia pastoris. Appl Biochem characterization of a new laccase from Bacillus
Microbiol 50, 140–147. licheniformis catalyzing dimerization of phenolic acids.
85 Theerachat M, Emond S, Cambon E, Bordes F, Marty Appl Microbiol Biotechnol 79, 217–224.
A, Nicaud JM, Chulalaksananukul W, Guieysse D, 97 Koschorreck K, Schmid RD & Urlacher VB (2009)
Remaud-Simeon M & Morel S (2012) Engineering and Improving the functional expression of a Bacillus
production of laccase from Trametes versicolor in the licheniformis laccase by random and site-directed
yeast Yarrowia lipolytica. Bioresour Technol 125, mutagenesis. BMC Biotechnol 9, 12.
267–274. 98 Reiss R, Ihssen J & Thony-Meyer L (2011) Bacillus
86 Mate D, Garcia-Burgos C, Garcia-Ruiz E, Ballesteros pumilus laccase: a heat stable enzyme with a wide
AO, Camarero S & Alcalde M (2010) Laboratory substrate spectrum. BMC Biotechnol 11, 9.
evolution of high-redox potential laccases. Chem Biol 99 Machczynski MC, Vijgenboom E, Samyn B & Canters
17, 1030–1041. GW (2004) Characterization of SLAC: a small laccase
87 Camarero S, Pardo I, Canas AI, Molina P, Record E, from Streptomyces coelicolor with unprecedented
Martinez AT, Martinez MJ & Alcalde M (2012) activity. Protein Sci 13, 2388–2397.

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1211


Enzymatic degradation of lignin L. Pollegioni et al.

100 Eugenio ME, Hernandez M, Moya R, Martin- dioxygenases are involved in syringate degradation by
Sampedro R, Villar JC & Arias ME (2011) Evaluation Sphingomonas paucimobilis SYK-6. J Bacteriol 187,
of a new laccase produced by Streptomyces Ipomoea 5067–5074.
on biobleaching and ageing of Kraft pulps. 111 Masai E, Sasaki M, Minakawa Y, Abe T, Sonoki T,
Bioresources 6, 3231–3241. Miyauchi K, Katayama Y & Fukuda M (2004) A
101 Dube E, Shareck F, Hurtubise Y, Daneault C & novel tetrahydrofolate-dependent O-demethylase gene
Beauregard M (2008) Homologous cloning, is essential for growth of Sphingomonas paucimobilis
expression, and characterisation of a laccase from SYK-6 with syringate. J Bacteriol 186, 2757–2765.
Streptomyces coelicolor and enzymatic decolourisation 112 Martinez AT, Speranza M, Ruiz-Duenas FJ, Ferreira
of an indigo dye. Appl Microbiol Biotechnol 79, P, Camarero S, Guillen F, Martinez MJ, Gutierrez A
597–603. & del Rio JC (2005) Biodegradation of
102 Otsuka Y, Sonoki T, Ikeda S, Kajita S, Nakamura M lignocellulosics: microbial, chemical, and enzymatic
& Katayama Y (2003) Detection and characterization aspects of the fungal attack of lignin. Int Microbiol 8,
of a novel extracellular fungal enzyme that catalyzes 195–204.
the specific and hydrolytic cleavage of lignin 113 Guillen F, Martinez MJ, Munoz C & Martinez AT
guaiacylglycerol beta-aryl ether linkages. Eur J (1997) Quinone redox cycling in the ligninolytic fungus
Biochem 270, 2353–2362. Pleurotus eryngii leading to extracellular production of
103 Sato Y, Moriuchi H, Hishiyama S, Otsuka Y, Oshima superoxide anion radical. Arch Biochem Biophys 339,
K, Kasai D, Nakamura M, Ohara S, Katayama Y, 190–199.
Fukuda M et al. (2009) Identification of three alcohol 114 Henriksson G, Ander P, Pettersson B & Pettersson G
dehydrogenase genes involved in the stereospecific (1995) Cellobiose dehydrogenase (Cellobiose oxidase)
catabolism of arylglycerol-beta-aryl ether by from Phanerochaete chrysosporium as a wood
Sphingobium sp. strain SYK-6. Appl Environ Microbiol degrading enzyme – studies on cellulose, xylan and
75, 5195–5201. synthetic lignin. Appl Microbiol Biotechnol 42,
104 Meux E, Prosper P, Masai E, Mulliert G, Dumarcay 790–796.
S, Morel M, Didierjean C, Gelhaye E & Favier F 115 Henriksson G, Johansson G & Pettersson G (2000) A
(2012) Sphingobium sp SYK-6 LigG involved in lignin critical review of cellobiose dehydrogenases. J Biotechnol
degradation is structurally and biochemically related 78, 93–113.
to the glutathione transferase omega class. FEBS Lett 116 Eltis LD & Bolin JT (1996) Evolutionary relationships
586, 3944–3950. among extradiol dioxygenases. J Bacteriol 178,
105 Reiter J, Strittmatter H, Wiemann LO, Schieder D & 5930–5937.
Sieber V (2013) Enzymatic cleavage of lignin beta-O-4 117 Williams PA, Assinder SJ & Shaw LE (1990)
aryl ether bonds via net internal hydrogen transfer. Construction of hybrid xylE genes between the two
Green Chem 15, 1373–1381. duplicate homologous genes from TOL plasmid
106 Tanamura K, Abe T, Kamimura N, Kasai D, pWW53: comparison of the kinetic properties of the
Hishiyama S, Otsuka Y, Nakamura M, Kajita S, gene products. J Gen Microbiol 136, 1583–1589.
Katayama Y, Fukuda M et al. (2011) Characterization 118 Junca H, Plumeier I, Hecht HJ & Pieper DH (2004)
of the third glutathione S-transferase gene involved in Difference in kinetic behaviour of catechol 2,3-
enantioselective cleavage of the beta-aryl ether by dioxygenase variants from a polluted environment.
Sphingobium sp. strain SYK-6. Biosci Biotechnol Microbiology 150, 4181–4187.
Biochem 75, 2404–2407. 119 Carboni-Oerlemans C, Dominguez de Maria P, Tuin
107 Sonoki T, Masai E, Sato K, Kajita S & Katayama Y B, Bargeman G, van der Meer A & van Gemert R
(2009) Methoxyl groups of lignin are essential carbon (2006) Hydrolase-catalysed synthesis of
donors in C1 metabolism of Sphingobium sp. SYK-6. peroxycarboxylic acids: biocatalytic promiscuity for
J Basic Microbiol 49 (Suppl. 1), S98–S102. practical applications. J Biotechnol 126, 140–151.
108 Bugg TD, Ahmad M, Hardiman EM & Rahmanpour 120 Duncan S, Jing Q, Katona A, Kazlauskas RJ,
R (2011) Pathways for degradation of lignin in Schilling J, Tschirner U & Aldajani WW (2010)
bacteria and fungi. Nat Prod Rep 28, 1883–1896. Increased saccharification yields from aspen biomass
109 Masai E, Katayama Y, Nishikawa S & Fukuda M upon treatment with enzymatically generated peracetic
(1999) Characterization of Sphingomonas paucimobilis acid. Appl Biochem Biotechnol 160, 1637–1652.
SYK-6 genes involved in degradation of lignin-related 121 Wiermans L, Perez-Sanchez M & Dominguez de
compounds. J Ind Microbiol Biotechnol 23, 364–373. Maria P (2013) Lipase-mediated oxidative
110 Kasai D, Masai E, Miyauchi K, Katayama Y & delignification in non-aqueous media: formation of
Fukuda M (2005) Characterization of the gallate de-aromatized lignin-oil and cellulase-accessible
dioxygenase gene: three distinct ring cleavage polysaccharides. ChemSusChem 6, 251–255.

1212 FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS


L. Pollegioni et al. Enzymatic degradation of lignin

122 Varnai A, Huikko L, Pere J, Siika-Aho M & Viikari L 126 Solomon EI, Penfield KW, Gewirth AA, Lowery MD,
(2011) Synergistic action of xylanase and mannanase Shadle SE, Guckert JA & LaCroix LB (1996)
improves the total hydrolysis of softwood. Biores Electronic structure of the oxidized and reduced blue
Technol 102, 9096–9104. copper sites: contributions to the electron transfer
123 Kudanga T & Le Roes-Hill M (2014) Laccase pathway, reduction potential, and geometry. Inorg
applications in biofuels production: current status and Chim Acta 243, 67–78.
future prospects. Appl Microbiol Biotechnol 98, 127 Kawai S, Nakagawa M & Ohashi H (2002)
6525–6542. Degradation mechanisms of a nonphenolic beta-O-4
124 Chen YR, Sarkanen S & Wang YY (2012) Lignin- lignin model dimer by Trametes versicolor laccase in
degrading enzyme activities. Methods Mol Biol 908, the presence of 1-hydroxybenzotriazole. Enzy Microb
251–268. Tech 30, 482–489.
125 Wariishi H, Akileswaran L & Gold MH (1988) 128 Heinfling A, Ruiz-Duenas FJ, Martinez MJ,
Manganese peroxidase from the basidiomycete Bergbauer M, Szewzyk U & Martinez AT (1998) A
Phanerochaete chrysosporium: spectral characterization study on reducing substrates of manganese-oxidizing
of the oxidized states and the catalytic cycle. peroxidases from Pleurotus eryngii and Bjerkandera
Biochemistry 27, 5365–5370. adusta. FEBS Lett 428, 141–146.

FEBS Journal 282 (2015) 1190–1213 ª 2015 FEBS 1213

You might also like