Download Deciphering the Enigma of Lignification: Precursor Transport

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Metabolism wikipedia , lookup

Lipid signaling wikipedia , lookup

Magnesium in biology wikipedia , lookup

Biochemical cascade wikipedia , lookup

Gene regulatory network wikipedia , lookup

Plant breeding wikipedia , lookup

Polyclonal B cell response wikipedia , lookup

Paracrine signalling wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Vectors in gene therapy wikipedia , lookup

Signal transduction wikipedia , lookup

Transcript
Molecular Plant
•
Volume 5
•
Number 2
•
Pages 304–317
•
March 2012
REVIEW ARTICLE
Deciphering the Enigma of Lignification:
Precursor Transport, Oxidation, and the
Topochemistry of Lignin Assembly
Chang-Jun Liu1
ABSTRACT Plant lignification is a tightly regulated complex cellular process that occurs via three sequential steps: the
synthesis of monolignols within the cytosol; the transport of monomeric precursors across plasma membrane; and the
oxidative polymerization of monolignols to form lignin macromolecules within the cell wall. Although we have a reasonable understanding of monolignol biosynthesis, many aspects of lignin assembly remain elusive. These include the precursors’ transport and oxidation, and the initiation of lignin polymerization. This review describes our current knowledge
of the molecular mechanisms underlying monolignol transport and oxidation, discusses the intriguing yet leastunderstood aspects of lignin assembly, and highlights the technologies potentially aiding in clarifying the enigma of plant
lignification.
Key words:
lignification; monolignol transport; ABC transporter; laccase; peroxidase.
INTRODUCTION
Lignin is complex phenypropanoid polymer derived primarily
from three cinnamyl alcohols: p-coumaryl, coniferyl, and
sinapyl alcohols (termed monolignols). It is a crucial structural
component preserving the integrity of plant cell wall, imparting stiffness and strength of vascular plants, enabling the
transport of water and solutes through the tracheary elements
in the vasculature system, and affording physical barriers
against invasions of phytopathogens, and other environmental stresses (Boerjan et al., 2003; Ralph et al., 2004a). However,
its presence contributes to the recalcitrance of cell wall to degradation, and thus is detrimental to using cellulosic fibers in
cattle feedstock, for pulping and paper making, and for producing liquid biofuels (Chen and Dixon, 2007; Li et al., 2008;
Weng et al., 2008).
Plant lignification is a cellular process generating lignin
polymer in the cell wall. In general, it occurs in three stages:
the biosynthesis of monolignols in the cytosol; the transport
of these monolignols to the cell wall; and their subsequent oxidative dehydrogenation and polymerization to form heterogeneous macromolecules. Over several decades, extensive
studies have centered on monolignol biosynthesis, particularly
on elucidating its pathways and the molecular regulation of
the biosynthetic processes. The progresses have been intensively reviewed correspondingly (Anterola and Lewis, 2002;
Boerjan et al., 2003; Ralph et al., 2004a; Ralph, 2007; Li
et al., 2008; Vanholme et al., 2008; Zhong and Ye, 2009; Umezawa, 2010; Vanholme et al., 2010; Zhao and Dixon, 2011).
Compared to the knowledge of monolignol biosynthesis,
our understanding of lignin assembly, namely the transport
of lignin precursors, their deposition, and subsequent activation and polymerization, is fragmentary. In this review, we outline the different speculations on the sequestration, transport,
and subsequent spatial deposition of the monolignols, and describe recent progresses in molecular genetics and biochemical
studies aimed at gaining an understanding of the underlying
molecular mechanisms for translocating monolignols across
cell membranes, advances in identifying the oxidative dehydrogenation enzymes, and progresses and issues in detailing
the topochemistry of lignification. Meanwhile, we highlight
the technological developments that might be applied productively to explore lignin assembly.
1
To whom correspondence should be addressed. E-mail [email protected],
tel. 631-344-2966, fax 631-344-3407
ª The Author 2012. Published by the Molecular Plant Shanghai Editorial
Office in association with Oxford University Press on behalf of CSPB and
IPPE, SIBS, CAS.
doi: 10.1093/mp/ssr121, Advance Access publication 3 February 2012
Received 29 September 2011; accepted 13 December 2011
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
Biology Department, Brookhaven National Laboratory, Upton, NY 11973, USA
Liu
MONOLIGNOL BIOSYNTHESIS AND THE
POTENTIAL PHYSICAL ORGANIZATION
OF THE ENZYMES
Enigma of Plant Lignification
|
305
are membrane-anchored cytochrome P450 proteins that predictably associate with the outer surface of the endoplasmic reticulum (ER) by virtue of their N-terminal membrane anchor (Li
et al., 2008). Nevertheless, many other enzymes, such as phenylalanine ammonia lyase (PAL), 4-hydroxycinnamoyl CoA ligase
(4CL), caffeoyl CoA O-methyltransferase (CCoAOMT), hydroxycinnamoyl CoA reductase (CCR), caffeic acid/5-hydroxyferulic
acid 3/5-O-methyltransferase (COMT), and (hydroxy)cinnamyl
alcohol dehydrogenase (CAD) found in diverse species operationally are soluble proteins and likely to be located within
the cytosol (Takabe et al., 1985; Nakashima et al., 1997; Chen
et al., 2000). Apparently, the different compartmental propensity of the monolignol biosynthetic enzymes, together with the
exclusive plastidic localization of the shikimate pathway that
leads to the synthesis of aromatic amino acids for phenylpropanoids (Herrmann and Weaver, 1999; Rippert et al., 2009), implies
either the occurrence of multiple subcellular sequestration of
metabolic intermediates or the ideal physical organization of
Figure 1. The Scheme of the Simplified Shikimate–Phenylpropanoid–Lignin Biosynthetic Pathway, Illustrating Different Compartmentalization of the Biosynthesis.
CM, chorismate mutase; TYRA, arogenage dehydrogenase; ADT, arogenate dehydratase; PAT, prephenate aminotransferase; PAL, phenylalanine ammonia lyase; TAL, tyrosine ammonia lyase; C4H, cinnamic acid 4-hydroxylase; 4CL, 4-hydroxycinnamoyl CoA ligase; HCT, hydroxycinnamoyl; CoA, shikimate/quinate hydroxycinnamoyltransferase; C3#H, p-coumaroylshikimate 3#-hydroxylase; CCoAOMT, caffeoyl CoA
O-methyltransferase; CCR, cinnamoyl CoA reductase; Cald5H/F5H, coniferaldehyde/ferulate 5-hydroxylase; COMT, caffeic acid/5-hydroxyferulic
acid O-methyltransferase; CAD, (hydroxy)cinnamyl alcohol dehydrogenase; POX, peroxidase; LAC, laccase.
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
Originating from the shikimate pathway, wherein plant produces aromatic amino acids, the biosynthesis of monolignols
starts from the deamination of phenylalanine by the entry
point enzyme phenylalanine ammonia lyase (PAL), and then
undergoes subsequent aromatic-ring modifications via hydroxylation and methylation, and the transformation of the
carboxylic moiety of the propane tail through esterification
and reduction. This process yields primarily three hydroxycinnamoyl alcohols, p-coumaryl, coniferyl, and sinapyl alcohols
(i.e. monolignols) (Umezawa, 2010; Vanholme et al., 2010).
More than 10 enzymes sequentially catalyze this synthetic
pathway (Figure 1). Three of them, namely cinnamic acid
4-hydroxylase (C4H), p-coumaroylshikimate 3#-hydroxylase
(C3#H), and coniferaldehyde/ferulic acid 5-hydroxylase (F5H),
d
306
|
Liu
d
Enigma of Plant Lignification
physically organized and/or compartmentalized and how such
macromolecular organizations affect lignin deposition are intriguing aspects for future explorations.
TRANSPORT OF MONOLIGNOLS ACROSS
CELL MEMBRANE
After monolignols are synthesized in the cytosol, they must be
moved into the cell wall, where they are oxidized, and integrated into the matrix of the secondary cell wall. The exact
mechanism whereby this occurs remains unclear; at least three
models or pathways are envisaged, as illustrated in Figure 2: (1)
exocytosis via ER-Golgi-derived vesicles; (2) passive diffusion
via hydrophobic reactions through the plasma membrane;
and (3) active transport via plasma membrane-located transporters or the other facilitators. We have discussed these potential mechanisms in detail in a recent review (Liu et al., 2011).
To keep the integrity of this paper, we reiterate some related
contents here but focus more on the up-to-date progresses on
this subject.
Monolignol Glucosylation and Deglucosylation
Monolignols in cells occur both as free aglycones (non-sugar
compounds) and as water-soluble 4-O-b-D-glucosides. The latter,
namely p-coumaryl alcohol glucoside, coniferin, and syringin,
are primarily found in gymnosperms (Terazawa and Miyake,
1984; Whetten and Sederoff, 1995). A set of angiosperm species
also can generate those glucoconjugates in particular tissues; for
instance, coniferin and syringin accumulate in Arabidopsis roots
and their levels can be raised by light treatment (Hemm et al.,
2004). In conifers, the build-up of coniferin correlates spatially
and temporally with the beginning of secondary growth
(Savidge, 1988, 1989). Presumably, these glucoconjugates are
stored within the vacuoles of cambial cells (Leinhos and Savidge,
1993; Dharmawardhana et al., 1995). The possible presence of
the glucosides of monolignols (or of cinnamaldehydes) in
vacuoles of differentiating xylem cells led to the suggestion that
those glucoconjugates may serve the storage and transport
functions of monolignols at least in gymnosperms and some evolutionarily less-advanced angiosperms (Steeves et al., 2001; Tsuji
and Fukushima, 2004). Presumably, those glucosides are sequestrated first from the cytosol to vacuoles and then exported to the
cell wall (Dharmawardhana et al., 1995, 1999; Samuels et al.,
2002; Escamilla-Treviño et al., 2006).
Monolignol glucosylation is catalyzed by coniferyl/sinapyl alcohol glucosyltransferase (Steeves et al., 2001). The activity of
this enzyme was detected in many gymnosperms and angiosperms; the woody angiosperms exhibited higher enzyme activities than did the herbaceous species (Ibrahim, 1977). On the
other hand, monolignol deglucosylation, catalyzed by specific
b-glucosidase, is thought to occur in the cell wall at the
site of lignin polymerization. The apoplastic location of
b-glucosidase was demonstrated by immunohistochemical
studies (Marcinowski et al., 1979; Burmeister and Hosel,
1981). The enzymes were identified in Norway spruce (Picea
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
the pathways to eliminate the phytotoxicity of aromatic intermediates and to enhance the transformation efficiency of the
intermediate compounds. Such physical organization of metabolic pathways is variously termed ‘metabolic compartment’,
‘metabolic channeling’, or ‘metabolon’. The metabolic channels
may contain arrays of consecutive, physically associated enzymes
assembled on membranes or other physical structures, or interacted directly with each other. This association can channel biosynthetic intermediate from one enzyme to the next without
releasing them into the total cellular metabolic pool (Srere,
1987; Hrazdina and Jensen, 1992; Burbulis and Winkel-Shirley,
1999; Winkel-Shirley, 1999). Earlier immunolocalization studies
on monolignol enzymes in the lignifying mesophyll cells of Zinnia elegans and the differentiating xylem cells of Eucalyptus and
Populus revealed that the isoforms of PAL, CAD, OMT, and 4CL
could associate with ER-Golgi-derived vesicles, and then dispersed into the cytosol (Takabe et al., 2001; Takeuchi et al.,
2001). In contrast, the peroxidase involved in the dehydrogenative polymerization of monolignols was synthesized in the
rough ER and then transported to the lignifying cell wall (Takabe
et al., 1985; Nakashima et al., 1997; Takabe et al., 2001; Takeuchi
et al., 2001). Metabolic labeling experiments also revealed
that a PAL isoform, PAL1, in tobacco was in close physical
association with the membrane-bound P450 enzyme C4H on
the microsomes for the synthesis of the active intermediate,
p-coumaric acid (Czichi and Kindl, 1977; Hrazdina and Jensen,
1992; Deshpande et al., 1993; Rasmussen and Dixon, 1999; Dixon,
2000). Membrane fractionation and protein gel-blot analyses
corroborated this observation. Furthermore, the transient coexpression of the PAL1 or PAL2 fused with green fluorescence
protein in the cells expressing C4H demonstrated that two
PAL isoforms exhibited different affinities in associating with
the ER microsome and C4H likely organized the enzyme complex
containing PAL1 and/or PAL2 (Achnine et al., 2004). These data
suggest that monolignol biosynthetic enzymes are partially
organized within the lignifying cells. Consequently, Dixon
et al. (2001) speculated that there was a complicated metabolic
channel for producing syringyl lignin, wherein three cytochrome
P450 enzymes, C4H, C3#H, and F5H, anchor other operationally
soluble enzymes, including COMT and perhaps specific forms of
PAL, 4CL, and/or CCR, to the surface of ER. However, experimentally substantializing this appealing hypothesis proved challenging and technically demanding. Recently, many biochemical,
biophysical, molecular, and cellular approaches were developed
and applied for dissecting protein–protein interactions or protein complex organization (Lalonde et al., 2008; Miernyk and
Thelen, 2008). Among them, several may offer the possibility
to precisely evaluate the potential physical/spatial organization
of lignin biosynthesis at cellular and subcellular levels. Particularly promising are the split ubiquitin membrane yeast
two-hybrid screening (Obrdlik et al., 2004), surface resonance
spectroscopy, fluorescence-based technology, such as bimolecular fluorescence complementation (BiFC) and the fluorescence
energy resonance transfer (FRET), and three-dimensional
super-solution microscopy. How lignin biosynthetic enzymes
Liu
d
Enigma of Plant Lignification
|
307
Symplastic transport of monolignols may export them to the cell wall through active transport or by passive diffusion. Alternatively, they
may be sequestrated and stored as glucoconjugates into the vacuoles in gymnosperms, before their subsequent transport to the cell wall
and hydrolysis to free monolignols for polymerization. The deposited monolignols in the cell wall diffuse to initiation sites where the
polymerization process begins. The polymerization to form different bond-linkages of lignin is known to be a random chemical process.
However, the nature of initiating sites and the way in which the amount and type of lignin formation is controlled across the cell wall are
poorly understood.
abies) (Marcinowski and Grisebach, 1978), pinus species
(Leinhos et al., 1994; Dharmawardhana et al., 1995), and
A. thaliana (Escamilla-Treviño et al., 2006). The cDNAs encoding coniferin b-glucosidase were isolated from both lodgepole pine and Arabidopsis; their expression was proved
highly specific at the site of lignification (Dharmawardhana
et al., 1999; Escamilla-Treviño et al., 2006). Based on the discovery of UDP-glucose:coniferyl alcohol glucosyltransferase/
coniferin b-glucosidase, a monolignol glucosyltransferase and
b-glucosidase-mediated transport pathway was proposed,
where the two enzymes coordinately regulate the storage
and mobilization of monolignols for lignin biosynthesis
(Dharmawardhana et al., 1995, 1999; Samuels et al., 2002;
Escamilla-Treviño et al., 2006). However, recent experiments
failed to support this speculation. A small cluster of glucosyltransferase genes were functionally characterized from
Arabidopsis. The encoded enzymes exhibited ability to glucoconjugate monolignols in vitro (Lim et al., 2001). However, disrupting those genes’ expression did not perturb lignin
deposition (Vanholme et al., 2008), although a corresponding
reduction or accumulation of soluble monolignol glucosides
in the roots or leaves of transgenic plants was recorded (Lanot
et al., 2006). Moreover, feeding [3H]-Phe to the dissected xylem
of lodgepole pine (Pinus contorta) revealed that the radiolabels
were incorporated directly into the monolignols and the lignins
accumulated in the cell wall. The substantial amount of the
radiolabels was associated neither with monolignol glucoside
(coniferin) nor with the interior of the central vacuole, where,
expectedly, coniferin is stored (Kaneda et al., 2008). These data
argue against the proposed roles of monolignol glucosylation–
deglucosylation in exporting lignin precursors across plasma
membrane, and imply that monolignol aglycone is the chemical
form used for such transport.
Mechanisms for Transporting Monolignols across Cell
Membranes
Exocytosis via Vesicles Derived from Endoplasmic
Reticulum-Golgi Bodies
The non-cellulosic polysaccharides of plant cell wall, such as pectin and hemicellulose, are known to be synthesized within the
Golgi bodies and exported to the cell wall through an exocytosis
mechanism (Lerouxel et al., 2006; Sandhu et al., 2009). Early autoradiographic and immunochemical studies implied a similar
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
Figure 2. The Scheme for Monolignol Deposition and the Subsequent Initiation of Lignin Polymerization within the Cell Wall.
308
|
Liu
d
Enigma of Plant Lignification
Passive Diffusion
Genetic engineering of lignin biosynthesis and chemical analysis of the resulting transgenic cell walls reveal that many nontraditional phenolics can be incorporated into the lignin
polymer. In transgenic plants and/or mutants deficient
in monolignol biosynthesis, some metabolite intermediates
or precursors, such as 5-hydroxyconiferyl alcohol (Van
Doorsselaere et al., 1995; Ralph et al., 2001), hydroxycinnamaldehydes (Baucher et al., 1996), and hydroxycinnamic acids
(Leple et al., 2007; Ralph et al., 2008), were integrated into lignin. In addition, in monocot and some dicot species, monolignols are often further acylated with p-hydroxybenzoate,
p-coumarate (Lu et al., 2004), and acetate (Lu and Ralph,
2002). These derivatives occur in the cell wall lignin (del Rio
et al., 2007, 2008). Such accommodation of alternative monomers in lignification implies a potential non-selective passive
diffusion of lignin monomers through plasma membrane
(Vanholme et al., 2008, 2010). This possibility was further evidenced from in vitro partitioning experiments using immobilized liposomes or lipid-bilayer discs as the model membranes.
Incubating monolignol-like compounds with artificial model
membranes caused them to freely partition into the membrane phase due to hydrophobic–hydrophilic interactions
(Boija et al., 2007, 2008). However, so far, we lack definitive
biochemical evidence for such transport in plant tissues.
Transporter-Mediated Active Sequestration and Export of
Monolignols
Global transcriptomic, proteomic studies, or functional genomics
in gymnosperms and angiosperms have frequently uncovered
highly expressed membrane transporters in lignifying or lignified tissues. For example, several ESTs encoding for ATP-binding
cassette (ABC) transporters were disclosed from gene expression
profiling during wood transformation (Allona et al., 1998;
Hertzberg et al., 2001; Kirst et al., 2003; Egertsdotter et al.,
2004). Transcriptional profiling based on the first-generation
loblolly pine microarray revealed the abundance of distinct
ABC transporter-like genes in the mature and compression
woods, wherein different types of lignins accumulated (Whetten
et al., 2001; Egertsdotter et al., 2004). In global transcript profiling of A. thaliana stems, seven ABC transporter genes exhibited
similar expression patterns with the functionally known monolignol biosynthetic genes (Ehlting et al., 2005). ABC transporter
genes were also recognized in the Eucalyptus xylem-subtractive
library (Paux et al., 2004) and in the cultures of lignifying Zinnia
tracheary elements (Pesquet et al., 2005). Also, the cell wall
macro-array analysis of maize (Zea mays) brown mid-rib (bm)
mutants indicated that the lower levels of guaiacyl lignin in
bm2 mutants might reflect the decreased transcription of one
ABC transporter gene (Guillaumie et al., 2007). Furthermore,
in a recent proteomics study on plasma membranes from poplar
leaf, xylem, and cambium/phloem, a set of ABC transporters
showed a rather ‘specific’ tissue distribution; some members
of the ABC transporter subfamily B and G were identified only
in the cambium/phloem (Nilsson et al., 2010). These data imply
that ABC transporters might play a significant role in cell wall
lignification, presumably in transporting monolignols to the wall
(Samuels et al., 2002; Douglas and Ehlting, 2005).
ABC transporters constitute a large, diverse family of proteins
that translocate a broad range of substances across cell membranes (Higgins, 1992; Sánchez-Fernández et al., 2001; Verrier
et al., 2008). The genomes of Arabidopsis and rice (Oryza sativa)
each contain more than 120 putative ABC transporters (SánchezFernández et al., 2001; Jasinski et al., 2003). Many of these proteins exhibited diverse transport activities to a variety of small
molecule metabolites, such as glutathione conjugates (Lu
et al., 1998; Tommasini et al., 1998), auxins (Geisler et al.,
2005; Strader and Bartel, 2009; Ruzicka et al., 2010), abscisic acid
(Kang et al., 2010; Kuromori et al., 2010), inorganic ions (Suh
et al., 2007), malate (Lee et al., 2008), wax and cutin precursors
(Pighin et al., 2004; McFarlane et al., 2010), defensive secondary
metabolites such as antifungal terpenoids (Jasinski et al.,
2001), tropane alkaloids (hyoscyamine and scopolamine)
(Goossens et al., 2003), benzylisoquinoline alkaloid (berberine)
(Shitan et al., 2003; Shitan and Yazaki, 2007), flavone glucuronides (Klein et al., 2000; Frangne et al., 2002), isoflavone genistein (Sugiyama et al., 2007), and many xenobiotics (Baerson
et al., 2005).
To ascertain whether the transport of monolignol is an active process that involves ABC transporters, we recently
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
mechanism for transporting lignin precursors. A vesicle trafficking between cytosol and plasmalemma in the differentiating
tracheids of wheat xylem was observed (Pickett-Heaps, 1968).
Feeding the developing xylem with isotopically labeled Phe,
Tyr, and cinnamic acid resulted in their incorporation into the
rough ER, the Golgi apparatus, and some vesicles that were associated with plasma membrane, or aggregated near the bands
of cell wall microtubules (Pickett-Heaps, 1968; Fujita and
Harada, 1979; Takabe et al., 1985). These findings suggested
that the ER-Golgi apparatus are likely to be involved in synthesizing and transporting monolignols to the cell wall. However,
since the applied radiotracers are assimilated efficiently into lignin and into proteins, it might complicate the interpretation of
these autoradiographic data. Indeed, in a recent experiment,
Kaneda et al. (2008) fed the [3H]-Phe to the cambium/developing xylem tissues of lodgepole pine that were isolated by cryofixation and freezing substitution before sectioning (those
treatments substantially minimize the damage to the cells
and thus prevent pseudo-autographic imaging). By selectively
inhibiting phenylpropanoid or protein biosynthesis, they discovered that the radioactivity presented in the ER-Golgi was
primarily derived from proteins rather than phenylpropanoids/monolignols. Further, they observed that the abundant
Golgi and Golgi-vesicle clusters in the developing xylem cells
did not load with phenylpropanoids (Kaneda et al., 2008). Their
study suggests that ER-Golgi vesicle-mediated exocytosis is
unlikely to play a significant role in transporting monolignols.
Liu
Issues in Identifying Particular Transporters Responsible
for Depositing Lignin
To identify the specific ABC transporters functioning in lignification, Kaneda et al. (2011) analyzed a subset of Arabidopsis
ABC transporters whose expressions have a good correlation
Enigma of Plant Lignification
|
309
with phenylpropanoid–lignin biosynthesis and are higher in
the vascular or interfascicular tissues of the primary stem
(Ehlting et al., 2005). However, mutant lines singly deficient
in those ABC transporter genes did not attenuate xylem lignification. Instead, two mutant lines respectively disrupting
ABCB14 and ABCB15 impaired auxin polar transport in the
stem, even though the former was recently characterized as
a malate importer in leaf guard cells (Lee et al., 2008). These
findings highlight the difficulty in characterizing the specific
monolignol transporters. Since lignification is a very complicated cellular process, orchestrating a large set of gene expression during cell differentiation and secondary cell wall
thickening, the ABC transporters found in vascular tissues
and correlated with their lignification may be involved in distinct biological processes. Furthermore, lignin deposition and
distribution are known to be tissue/cell-type specific, which
may require a set of distinct membrane transporters. Many
ABC transporter genes have overlapping tissue expression pattern and therefore possible functional redundancy. Analyzing
plants with a single gene deletion might fail to distinguish its
consequence on lignification, particularly when lignin content
is measured in large-sized, pooled cell wall samples. In addition, ABC transporters typically exhibit low substrate specificity (Verrier et al., 2008; Ruzicka et al., 2010); some members
may transport a set of unrelated molecules and fulfill disparate
functions, as implied by the researches on ABCB14 (Lee et al.,
2008; Kaneda et al., 2011). Besides creating multiple genedeficient lines for the related ABC transporters, we must
precisely measure the deficit of lignification in a tissue or
cell-type-specific manner. Laser capture microdissection of vascular bundles and interfascicular fibers, followed by microthioacidolysis or pyrolytic GC/MS analysis (Patten et al.,
2010) may offer a practical, higher-resolution strategy for
identifying the particular set of ABC transporters or other
membrane transporters involved in transporting monolignols
and depositing lignin.
ENZYMES RESPONSIBLE FOR OXIDATIVE
DEHYDROGENATION
After their deposition in the cell wall, monolignols undergo
dehydrogenation, producing resonance-stabilized radicals.
Then, the resulting phenoxy radicals couple each other or to
the growing lignin polymer, if the polymer was also oxidized,
to form different types of sub-linkages. Monolignol oxidation/
dehydrogenation is proposed to be catalyzed by peroxidases,
laccases, or other phenol oxidases. Those enzymes are widely
found in lignin-forming tissues and can oxidize a variety of
phenolics, including monolignols in vitro (Freudenberg,
1959; Ralph et al., 2004a; Davin et al., 2008).
Peroxidase uses hydrogen peroxide as its electron receptor
to oxidize a variety of phenolics (Harkin and Obst, 1973). Different peroxidases have distinct predilection in oxidizing the
conventional monolignols. Most peroxidases readily oxidize
coniferyl alcohol, and the resulting radicals can transfer to,
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
conducted the in vitro uptake assay to monolignols and/or
their glucosides by using plasma and vacuolar membrane
vesicles prepared from both Arabidopsis rosette leaves and
poplar roots (Miao and Liu, 2010). Our study reveals several
lines of evidence demonstrating the involvement of ABC-like
transporters in monolignol sequestration and transport: (1)
the transport of monolignols across either plasma or vacuolar
membranes largely depends on the presence of nucleotide
phosphates, particularly ATP; lacking it severely impairs the
transport activity of either type of vesicles, suggesting that
transport is an active process requiring energy; (2) a set of specific ABC transporter inhibitors largely reduce the uptake activity of both plasma and vacuolar membrane vesicles for
lignin precursors in the presence of ATP; (3) the uptake of
monolignols or their glucosides by membrane vesicles displays
typical protein–ligand binding kinetics, indicating that it is
a membrane protein-mediated biochemical process rather
than passive diffusion; and (4) the uptake to lignin precursors
by membrane vesicles shows obvious selectivity. In the presence of ATP, the vacuolar membrane vesicles actively sequester
monolingol glucoconjugates, whereas the preference of the
inverted plasma membrane vesicles is for the monolignol aglycones; this difference suggests that the glucosylation of monolignols is a prerequisite for vacuolar sequestration, while the
aglycones are the form for direct export into cell walls. Also,
the findings implicate the distinct classes of ABC transporters
as being involved in either partitioning the polarized ‘storage
form’ glucosides of monolignols into vacuoles or exporting the
hydrophobic aglycones across plasmalemma. Compared to the
vacuolar membrane vesicles, plasma membrane ones exhibit
a more relaxed substrate specificity, being able to convey
p-hydroxycinnamyl alcohols and other related cinnamaldehydes, thereby suggesting the basis for the observed plasticity
of lignin biosynthesis (Miao and Liu, 2010). Promiscuous
substrate-transport activity and/or the low intrinsic level of
plasmalemma diffusion may engender the deposition of
non-classic lignin precursors in the cell wall. In addition, we
found that the chemical character of the para-hydroxyl of
monolignol substantially influences the nature of transport.
While the 4-O-glucosylation of monolignols likely dominates
the vacuolar sequestration, the etherification of the parahydroxyl with a methyl moiety largely releases the dependence of monolignol transport on the ATP-energized
mechanism. When incubated with 4-O-methyl coniferyl
alcohol in the absence of ATP, the amount of this etherified
monolignol derivative accumulates in the inverted plasmamembrane vesicles is comparable to that of conventional
monolignols accumulated in the presence of ATP (Y. Miao
and C.-J. Liu, unpublished data).
d
310
|
Liu
d
Enigma of Plant Lignification
their expressions closely concur across a wide variety of tissues/
development stages (Turlapati et al., 2011). A few family members display unique tissue/cell-type expression patterns, suggesting their specific roles in either lignification or other
oxidative processes (Pourcel et al., 2005; Turlapati et al.,
2011). Determining the role of laccases in lignification yielded
ambiguous findings until a recent genetic study wherein the
functions of the Arabidopsis laccase genes LAC4 and LAC17
in xylem lignification were demonstrated conclusively (Berthet
et al., 2011). About a 30–40% reduction of lignin content in the
stem of mutant plants simultaneously disrupting LAC4 and
LAC17 was registered. A deficiency of LAC17 primarily affected
the deposition of guaiacyl lignin and resulted in a change of
the S/G ratio in mutant lines, implying the preference of LAC17
for oxidizing coniferyl alcohol. However, definitive specifications of the biochemical properties of those LACs are lacking
presently. It will be interesting to see whether the identified
LACs are able to oxidize sinapyl alcohol and other higher molecule oligomers/polymers.
Interestingly, down-regulation of LAC broadly changed the
phenylpropanoid biosynthetic pathway. Almost the entire
batch of monolignol biosynthetic genes was suppressed coordinately with the disruption of LAC4 and 17 (Berthet et al.,
2011). These data imply the existence of some form of feedback regulatory mechanisms at the plasma membrane/cell wall
interior that prevents the accumulation/transport of monolignols when oxidative capacity is diminished. It is interesting
to decipher whether the accumulated monolignols or their
derivatives serve as signaling molecules, and how the set of
biosynthetic genes in the pathway are organized as a particular
‘regulon’.
SPATIAL DISTRIBUTION OF LIGNIN AND
TOPOCHEMISTRY OF LIGNIFICATION
Lignin is known to be selectively deposited in the walls of different cell types. For example, the vessels of birch wood are
enriched in G units, while the fiber wall incorporates both
S and G units (Fergus and Goring, 1970). This feature is also
demonstrated in Arabidopsis stems, where the lignin of the
vascular bundle rich in vessel cells primarily contains G units,
while the interfascicular fibers are enriched in S units (Chapple
et al., 1994). In a more subtle way, lignin is differentially deposited in discrete regions of the cell walls. In spruce wood,
the lignin of the middle lamella embeds more p-coumaryl
alcohol-derived units than does the secondary wall lignin that
mainly is derived from coniferyl alcohol (Whiting and Goring,
1982). Feeding labeled monolignols into developing xylem
reveals the preferential deposition of radiolabeled p-coumaryl
alcohol in the middle lamella or cell corners, whereas coniferyl
alcohol is located mostly in the secondary wall (Terashima
et al., 1993). Moreover, during the development of secondary
cell wall to form S1-, S2-, and S3- layers, the three kinds of
monolignols exhibit sequential deposition, in the order of
p-coumaryl alcohol (for the H unit), coniferyl alcohol (G unit),
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
and thus oxidize, sinapyl alcohol. Sinapyl alcohol itself is a poor
substrate for oxidation, and is catalyzed only by a very specific
peroxidase (Sasaki et al., 2006, 2008). These properties led to
a radical mediation model to explain lignin polymerization, in
which the coniferyl alcohol radicals generated through the interaction of the cell wall-bound peroxidases with monolignols
diffuses to the growing lignin polymer; thereafter, they either
transfer their higher oxidation state to a lignin molecule, if the
lignin polymer is at a base oxidative state, or undergo a coupling reaction to form a covalent bond with the polymer, if the
latter is at a higher oxidation state (Hatfield and Vermerris,
2001). Further, a complementary model was derived from this
radical-transfer hypothesis by suggesting the operation of
a diffusible redox shuttle-peroxidase system in which a membrane-bound peroxidase oxidizes manganese, the oxidized
ions diffuse through small pores in the cell wall to the lignification site, and then they oxidize both the monolignols and
growing lignin polymers. As the redox ions directly oxidize
the monolignols and polymers, the radicals as the primary oxidation products from both constitute a high concentration, thus
facilitating the coupling frequency of the monolignols with
growing polymers (Onnerud et al., 2002). However, the need
for having radical or redox-shuttle mediation was challenged
by the identification of a Populus peroxidase specific for sinapyl
alcohol that can directly and efficiently oxidize both sinapyl alcohol and the higher molecular polymers (Sasaki et al., 2004).
Peroxidases in plant are encoded by a multigene family. In
Arabidopsis, 73 class III peroxidase genes were identified
(Tognolli et al., 2002; Fagerstedt et al., 2010). Despite their
high degree of sequence and function redundancy that complicate the identification of lignin-specific enzymes, the biological functions of class III plant peroxidases in lignification
were demonstrated in a few genetic studies with altered expression of the peroxidase genes. Down-regulation of the
gene coding for a cationic peroxidase NtPrx60 (TP60) in tobacco resulted in up to a 50% reduction in total acetyl bromide
lignin, and the corresponding reduction in both the G and S
units (Blee et al., 2003); down-regulation of an anionic peroxidase PkPrx03 (PrxA3a) in aspen resulted in a decline of up to
20% of total lignin content in transgenic lines (Li et al., 2003).
These data suggest that species-specific oxidative enzymes are
involved in plant lignification; on the other hand, different
types of peroxidases may have overlapping biological functions in lignin polymerization.
Laccases (p-diphenol:dioxygen oxidoreductase, EC 1.10.3.2)
are copper-containing extracellular glycoproteins that require
O2 as secondary substrate to oxidize various phenolic, inorganic, and/or aromatic amine substrates (Reinhammar and
Malmstroem, 1981). Similar to peroxidase, laccases oxidize
monolignols in vitro (Freudenberg, 1959; Sterjiades et al.,
1992; Bao et al., 1993; Takahama, 1995; Richardson et al.,
1997), and laccase-like activities were detected in the lignifying cell walls of differentiating xylem (Driouich et al., 1992;
Bao et al., 1993; Liu et al., 1994; Richardson et al., 2000). Arabidopsis contains 17 putative laccase homologous genes, and
Liu
Enigma of Plant Lignification
|
311
ticularly for ferulate, these ‘wall-bound’ phenolics can dimerize or polymerize with each other, or with other cell wall
phenolics, forming ester-to-ether linkages to cross-link the adjacent polysaccharides, lignins, and/or structural proteins
(Ralph et al., 2004a). These ‘wall-bound’ phenolics purportedly
also act as nucleation sites for lignin polymerization (Ralph
et al., 2004a; Ralph, 2010). Peroxidase activity was noted, catalyzing the formation of ferulate-mediated ester to ether linkages between polysaccharide and lignin composition, or the
cross-links of ferulate residues in polysaccharides (Burr and
Fry, 2009a, 2009b). However, conclusive evidence to establish
the role of ‘wall-bound’ ferulate in initiating lignin polymerization is needed.
Since lignin assembly depends on the quantity and type of
monolignol radicals within the particular region of the cell
wall (Hatfield and Vermerris, 2001; van Parijs et al., 2010),
the potential sub-compartmental localization of the radicalproducing enzymes, namely peroxidase and laccase, within
the cell wall may affect the initiation of lignin polymerization
and control the spatial deposition of a particular type of lignin.
Both enzymes are extracellular proteins, and are known from
immunohistochemical analyses to locate in middle lamellae,
cell corner, and secondary cell wall of the lignifying cells
(Bao et al., 1993; Liu et al., 1994; Richardson et al., 2000; Sasaki
et al., 2006). However, there is little detail about their potential
sub-compartmentalization and spatial movement within the
cell wall during developmental programming and the lignification process. An early study found that the sycamore maple
laccase is far less active on phenolic substrates containing multiple aromatic rings than is peroxidase (Sterjiades et al., 1993).
This leads to the assumption that laccases may polymerize
monolignols into oligo-lignols during the early stages of lignification, whereas the cell wall peroxidases may function when
H2O2 is generated during the later stages of xylem cell development or in response to environmental stresses (Sterjiades
et al., 1993). Recently, intriguing findings from several pieces
of cell wall immunohistochemical studies revealed the existence of diverse microenvironments within cell walls that
are likely to influence the access of proteins or enzymes to specific targets (Lee et al., 2011). Some small-molecular-weight
carbohydrate binding modules or immuno-epitopes can be effectively blocked or restricted to access to their ligands with
the presence of pectic homogalacturonan (Herve et al.,
2010; Marcus et al., 2010). It is possible that the distribution
and localization of oxidative enzymes within the cell wall
are affected with the similar ‘polymer masking’ effects. Along
with characterizing lignin-specific laccases and peroxidases, it
now might be able to evaluate this appealing hypothesis. Detailed biochemical analyses of the identified laccases and peroxidases should clarify whether different types of oxidative
enzymes exhibit substrate specificity, particularly towards oxidizing high-molecule oligomers or polymers. Moreover, advanced microscopic techniques now enable us to dissect the
details of single protein behaviors in a particular cell compartment, as recently demonstrated in a series of researches
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
and sinapyl alcohol (S unit) (Terashima et al., 1986a, 1986b,
1993). Disturbing lignin biosynthesis by down-regulating particular monolignol biosynthetic genes, such as CCR, selectively
affects the spatial deposition of lignin in the fiber and xylem
cells, and even within the different sub-layers of secondary cell
wall (Chabannes et al., 2001; Ruel et al., 2009; Thevenin et al.,
2011). These data strongly suggest that lignification is tightly
regulated during developmental programs and that lignin deposition within the wall is a highly organized process. Although the exact molecular mechanism in controlling lignin
deposition remains elusive, the lignin composition of each individual cell appears to be controlled by a complex array of
gene expression in the differentiating cells or in adjacent cells.
This complexity would determine the number and type of lignin precursors deposited at lignification sites, which were predicted as the key factors controlling lignin structure (van Parijs
et al., 2010). Accordingly, the spatial and temporal expression
of the monolignol transporters and their substrate specificity
may greatly contribute to the tissue and cell-type-specific lignification.
Once monolignols are laid down in the cell wall, they are
proposed to diffuse to the initiation site for polymerization.
Lignification, in general, begins at the cell corner of the middle
lamella and the S1 region of the secondary cell wall before
spreading across the secondary wall towards the lumen
(Donaldson, 2001; Moller et al., 2006). Lignification of primary
cell wall typically begins after the start of the formation of the
secondary cell wall, while lignification of secondary cell wall
usually starts when secondary cell wall formation is completed.
Currently, it is unclear yet how the precise topochemistry of
lignification occurs and is controlled. A type of non-enzymatic
dirigent protein was characterized functioning as a template
guiding the formation of the bond linkage of two monolignol
molecules, and defining the stereochemistry of the resulted
optically active compound lignan (Davin et al., 1997). This proteinaceous control mechanism was envisaged for lignin polymerization, where proteins bearing dirigent sites guide the
phenoxy radical coupling within the cell wall (Davin and Lewis,
2000, 2005; Davin et al., 2008). The dirigent sites were then predicted to be the lignin initiation sites (Donaldson, 2001). However, lignin is structurally known as an optical inactive, racemic
polymer containing many different types of bonds connecting
monolignol residues in a somewhat random pattern
(Vanholme et al., 2010). These facts apparently challenge
the involvement of dirigent proteins in polymerization. In
addition, although a set of dirigent homolog genes were
found in plants, so far, none was definitively defined in lignin
polymerization, either biochemically or genetically.
Because lignification starts at the region furthest from the
protoplast, it was suggested that the initiation sites may have
been bound to specific regions of the existing cell wall
(Donaldson, 1994). The cell walls of monocot grasses and some
dicot species, including Arabidopsis thaliana, contain significant quantities of ester-bound hydroxycinnamates, primarily
p-coumarate and ferulate (Ishii, 1997; Ralph et al., 2004b). Par-
d
312
|
Liu
d
Enigma of Plant Lignification
CONCLUDING REMARKS
Lignification is an extremely complicated but tightly controlled biological process, even though the bond formation
in lignin polymerization is a random chemical process. Lignin
deposition and assembly are spatially and temporally programmed in differentiating cell walls. Two important events
therein are the transport of precursors across the protoplast
membrane and the subsequent oxidative dehydrogenation
in the cell walls. In concert with a complex array of transcriptional regulation (reviewed by Zhong and Ye, 2009; Umezawa,
2010; Zhao and Dixon, 2011) and potential metabolic organization of monolignol synthesis, spatially and temporally controlled monolignol transport and oxidation in distinct cell
types and/or the particular sub-compartments of the cell wall
are envisioned, which may play a critical role in controlling the
elaborate processes of lignin deposition. Moving the monolignols across cell membranes takes an energy-requiring, selec-
tive transport process that involves ABC transporters. Despite
the complexity of this process, research is underway towards
clarifying and identifying particular monolignol transporters.
Aided by modern microscopic technologies, together with the
identification of the monolignol transporters and the specific
oxidative enzymes for generating phenox radicals, the enigma
of lignification in plants will be better elucidated.
FUNDING
This work was supported by the Division of Chemical Sciences, Geosciences, and Biosciences, Office of Basic Energy Sciences of the US
Department of Energy through Grant DEAC0298CH10886; the National Science Foundation through grant MCB-1051675; the Laboratory Directed Research and Development Program of Brookhaven
National Laboratory (No 11–007); the CAS/SAFEA International Partnership Program for Creative Research Teams in Plant Metabolisms,
and the National Science Foundation of China for Oversea Distinguished Young Scholars (31028003). No conflict of interest declared.
REFERENCES
Achnine, L., Blancaflor, E.B., Rasmussen, S., and Dixon, R.A. (2004).
Colocalization of L-phenylalanine ammonia-lyase and cinnamate 4-hydroxylase for metabolic channeling in phenylpropanoid biosynthesis. Plant Cell. 16, 3098–3109.
Allona, I., et al. (1998). Analysis of xylem formation in pine by cDNA
sequencing. Proc. Natl Acad. Sci. U S A. 95, 9693–9698.
Anterola, A.M., and Lewis, N.G. (2002). Trends in lignin modification: a comprehensive analysis of the effects of genetic manipulations/mutations on lignification and vascular integrity.
Phytochemistry. 61, 221–294.
Baerson, S.R., et al. (2005). Detoxification and transcriptome response in Arabidopsis seedlings exposed to the allelochemical
benzoxazolin-2(3H)-one. J. Biol. Chem. 280, 21867–21881.
Bao, W., O’Malley, D.M., Whetten, R., and Sederoff, R.R. (1993).
A laccase associated with lignification in loblolly pine xylem.
Science. 260, 672–674.
Bates, M., Huang, B., and Zhuang, X. (2008). Super-resolution microscopy by nanoscale localization of photo-switchable fluorescent probes. Curr. Opin. Chem. Biol. 12, 505–514.
Baucher, M., et al. (1996). Red xylem and higher lignin extractability
by down-regulating a cinnamyl alcohol dehydrogenase in Poplar. Plant Physiol. 112, 1479–1490.
Berthet, S., et al. (2011). Disruption of LACCASE4 and 17 results in
tissue-specific alterations to lignification of Arabidopsis thaliana
stems. Plant Cell. 23, 1124–1137.
Blee, K.A., Choi, J.W., O’Connell, A.P., Schuch, W., Lewis, N.G., and
Bolwell, G.P. (2003). A lignin-specific peroxidase in tobacco
whose antisense suppression leads to vascular tissue modification. Phytochem. 64, 163–176.
Boerjan, W., Ralph, J., and Baucher, M. (2003). Lignin biosynthesis.
Annu. Rev. Plant Biol. 54, 519–546.
Boija, E., Lundquist, A., Edwards, K., and Johansson, G. (2007). Evaluation of bilayer disks as plant cell membrane models in partition studies. Anal. Biochem. 364, 145–152.
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
applying spin disc confocal microscopy to delineate the functional association of the cellulose synthase complex with cortical microtubules (Paredez et al., 2006; Gutierrez et al., 2009).
Aided with this type of microscopic technique, fluorescently
labeled cellulose synthase was clearly observed as motile punctates at the plasma membrane. They moved bi-directionally at
constant rates in linear tracks aligned and coincident with cortical microtubules, whereby they orient nascent microfibrils in
the secondary thickening cell walls (Paredez et al., 2006). In an
earlier ultrastructural study employing transmission electron
microscopy, Donaldson (1994) observed that, during the initiation of lignification in the cell wall of Pinus radiate, lignin
particles deposited in the S1 sub-layer appeared as elongated
structures, quite distinct from the round-shaped particles in
the middle lamellae. This finding suggests that the orientation
of lignin deposition in the secondary cell wall is affected by
either the deposited cellulose microfibrils or the wall’s microtubules. The same approach for monitoring the mobility of cellulose synthase might be applicable to probing the subcellular
action of oxidative enzymes during lignification. Moreover,
the emergence of super-resolution fluorescence microscopy
offers the capability of visualizing the three-dimensional fine
structure of the cell and to efficiently detect a single nanosized fluorophor with the remarkable spatiotemporal resolution of ,20 nm (Bates et al., 2008; Huang et al., 2009; Gutierrez
et al., 2010). This technique permits the imaging of features
that previously were only resolved via electron microscopy. Applying one type of super-resolution microscopy, Structured Illumination Microscopy (SIM), Fitzgibbon et al. (2010) were
able to see the apertures of individual plasmodesmal pores
and trace the fine structure of endoplasmic reticulum-derived
tubules through the pores and between adjacent sieve element cells. With such spectroscopic techniques, we might be
able to pinpoint the functional sites and action modes of
the specific laccases, peroxidases, and any other types of cell
wall proteins possibly related to lignification.
Liu
Boija, E., Lundquist, A., Nilsson, M., Edwards, K., Isaksson, R., and
Johansson, G. (2008). Bilayer disk capillary electrophoresis:
a novel method to study drug partitioning into membranes. Electrophoresis. 29, 3377–3383.
Burbulis, I.E., and Winkel-Shirley, B. (1999). Interactions among
enzymes of the Arabidopsis flavonoid biosynthetic pathway.
Proc. Natl Acad. Sci. U S A. 96, 12929–12934.
Burmeister, G., and Hosel, W. (1981). Immunohistochemical localization of beta-glucosidase in lignin and isoflavone metabolism
in Cicer arietinum L. seedlings. Planta. 152, 578–586.
Burr, S.J., and Fry, S.C. (2009b). Feruloylated arabinoxylans are oxidatively cross-linked by extracellular maize peroxidase but not
by horseradish peroxidase. Mol. Plant. 2, 883–892.
Enigma of Plant Lignification
|
313
Dharmawardhana, D.P., Ellis, B.E., and Carlson, J.E. (1995).
A b-glucosidase from lodgepole pine xylem specific for the lignin
precursor coniferin. Plant Physiol. 107, 331–339.
Dharmawardhana, D.P., Ellis, B.E., and Carlson, J.E. (1999). cDNA
cloning and heterologous expression of coniferin b-glucosidase.
Plant Mol. Biol. 40, 365–372.
Dixon, R.A., Chen, F., Guo, D., and Parvathi, K. (2001). The biosynthesis of monolignols: a ‘metabolic grid’, or independent pathways to guaiacyl and syringyl units? Phytochem. 57, 1069–1084.
Donaldson, L.A. (1994). Mechanical constrains on lignin deposition
during lignification. Wood Sci. Technol. 28, 111–118.
Donaldson, L.A. (2001). Lignification and lignin topochemistry: an
ultrastructural view. Phytochem. 57, 859–873.
Douglas, C.J., and Ehlting, J. (2005). Arabidopsis thaliana full genome longmer microarrays: a powerful gene discovery tool
for agriculture and forestry. Transgenic Res. 14, 551–561.
Chabannes, M., et al. (2001). In situ analysis of lignins in transgenic
tobacco reveals a differential impact of individual transformations on the spatial patterns of lignin deposition at the cellular
and subcellular levels. Plant J. 28, 271–282.
Driouich, A., Laine, A.C., Vian, B., and Faye, L. (1992). Characterization and localization of laccase forms in stem and cell cultures of
sycamore. Plant J. 2, 13–24.
Chapple, C., Vogt, T., Ellis, B.E., and Somerville, C. (1994). Arabidopsis mutant defective in phenylpropanoid pathway. Plant Cell. 4,
1413–1424.
Egertsdotter, U., et al. (2004). Gene expression during formation of
earlywood and latewood in loblolly pine: expression profiles of
350 genes. Plant Biol. (Stuttg.). 6, 654–663.
Chen, C.Y., et al. (2000). Cell-specific and conditional expression of
caffeoyl-coenzyme A-3-O-methyltransferase in poplar. Plant
Physiol. 123, 853–867.
Ehlting, J., et al. (2005). Global transcript profiling of primary stems
from Arabidopsis thaliana identifies candidate genes for missing
links in lignin biosynthesis and transcriptional regulators of fiber
differentiation. Plant J. 42, 618–640.
Chen, F., and Dixon, R.A. (2007). Lignin modification improves fermentable sugar yields for biofuel production. Nature Biotechnol. 25, 759–761.
Czichi, U., and Kindl, H. (1977). Phenylalanine ammonia-lyase and
cinnamic acid hydroxylase as assembled consecutive enzymes on
microsomal membranes of cucumber cotyledons: cooperation
and subcellular distribution. Planta. 134, 133–143.
Davin, L.B., and Lewis, N.G. (2000). Dirigent proteins and dirigent
sites explain the mystery of specificity of radical precursor coupling in lignan and lignin biosynthesis. Plant Physiol. 123,
453–461.
Davin, L.B., and Lewis, N.G. (2005). Lignin primary structures and
dirigent sites. Curr. Opin. Biotechnol. 16, 407–415.
Davin, L.B., et al. (1997). Stereoselective bimolecular coupling by an
auxillary (dirigent) protein without an active center. Science.
275, 362–366.
Davin, L.B., Jourdes, M., Patten, A.M., Kim, K.W., Vassao, D.G., and
Lewis, N.G. (2008). Dissection of lignin macromolecular configuration and assembly: comparison to related biochemical processes in allyl/propenyl phenol and lignan biosynthesis. Nat.
Prod. Rep. 25, 1015–1090.
Escamilla-Treviño, L.L., Chen, W., Card, M.L., Shih, M.-C.,
Cheng, C.-L., and Poulton, J.E. (2006). Arabidopsis thaliana
b-glucosidases BGLU45 and BGLU46 hydrolyse monolignol
glucosides. Phytochem. 67, 1651–1660.
Fagerstedt, K.V., Kukkola, E.M., Koistinen, V.V., Takahashi, J., and
Marjamaa, K. (2010). Cell wall lignin is polymerised by class III
secretable plant peroxidases in Norway spruce. J. Integr. Plant
Biol. 52, 186–194.
Fergus, B.J., and Goring, D.A.I. (1970). The distribution of lignin in
birch wood as determined by ultraviolet microscopy. Holzforschung. 24, 118–124.
Fitzgibbon, J., Bell, K., King, E., and Oparka, K. (2010). Superresolution imaging of plasmodesmata using three-dimensional
structured illumination microscopy. Plant Physiol. 153,
1453–1463.
Frangne, N., Eggmann, T., Koblischke, C., Weissenböck, G.,
Martinoia, E., and Klein, M. (2002). Flavone glucoside uptake into barley mesophyll and Arabidopsis cell culture vacuoles: energization occurs by H+-antiport and ATP-binding cassettetype
mechanisms. Plant Physiol. 128, 726–733.
del Rio, J.C., et al. (2008). Highly acylated (acetylated and/or
p-coumaroylated) native lignins from diverse herbaceous plants.
J. Agric. Food Chem. 56, 9525–9534.
Freudenberg, K. (1959). Biosynthesis and constitution of lignin.
Nature. 183, 1152–1155.
del Rio, J.C., Marques, G., Rencoret, J., Martinez, A.T., and
Gutierrez, A. (2007). Occurrence of naturally acetylated lignin
units. J. Agric. Food Chem. 55, 5461–5468.
Fujita, M., and Harada, H. (1979). Autoradiographic investigations of
cell wall development. II. Tritiated phenylalanine and ferulic acid assimilation in relation to lignification. Mokuzai Gakkaishi. 25, 89–94.
Deshpande, A.S., Surendranathan, K.K., and Nair, P.M. (1993). The
phenylpropanoid pathway enzymes in Solanum tuberosum exist
as a multienzyme complex. Indian Journal of Biochemistry and
Biophysics. 30, 36–41.
Geisler, M., et al. (2005). Cellular efflux of auxin catalyzed by the
Arabidopsis MDR/PGP transporter AtPGP1. Plant J. 44, 179–194.
Goossens, A., Hakkinen, S.T., Laakso, I., Oksman-Caldentey, K.M.,
and Inze, D. (2003). Secretion of secondary metabolites by
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
Burr, S.J., and Fry, S.C. (2009a). Extracellular cross-linking of maize
arabinoxylans by oxidation of feruloyl esters to form oligoferuloyl esters and ether-like bonds. Plant J. 58, 554–567.
d
314
|
Liu
d
Enigma of Plant Lignification
ATP-binding cassette transporters in plant cell suspension cultures. Plant Physiol. 131, 1161–1164.
Guillaumie, S., Pichon, M., Martinant, J.P., Bosio, M., Goffner, D.,
and Barriere, Y. (2007). Differential expression of phenylpropanoid and related genes in brown-midrib bm1, bm2, bm3, and
bm4 young near-isogenic maize plants. Planta. 226, 235–250.
Gutierrez, R., Grossmann, G., Frommer, W.B., and Ehrhardt, D.W.
(2010). Opportunities to explore plant membrane organization
with super-resolution microscopy. Plant Physiol. 154, 463–466.
Harkin, J.M., and Obst, J.R. (1973). Lignification in trees: indication
of exclusive peroxidase participation. Science. 180, 296–298.
Hatfield, R., and Vermerris, W. (2001). Lignin formation in plants:
the dilemma of linkage specificity. Plant Physiol. 126, 1351–1357.
Hemm, M.R., Rider, S.D., Ogas, J., Murry, D.J., and Chapple, C.
(2004). Light induces phenylpropanoid metabolism in Arabidopsis roots. Plant J. 38, 765–778.
Herrmann, K.M., and Weaver, L.M. (1999). The Shikimate Pathway.
Annu. Rev. Plant Physiol. Plant Mol. Biol. 50, 473–503.
Hertzberg, M., et al. (2001). A transcriptional roadmap to wood formation. Proc. Natl Acad. Sci. U S A. 98, 14732–14737.
Herve, C., Rogowski, A., Blake, A.W., Marcus, S.E., Gilbert, H.J., and
Knox, J.P. (2010). Carbohydrate-binding modules promote the
enzymatic deconstruction of intact plant cell walls by targeting
and proximity effects. Proc. Natl Acad. Sci. U S A. 107,
15293–15298.
Higgins, C.F. (1992). ABC transporters: from microorganisms to
man. Annu. Rev. Cell Biol. 8, 67–113.
Hrazdina, G., and Jensen, R.A. (1992). Spatial organization of
enzymes in plant metabolic pathways. Annu. Rev. Plant Physiol.
Plant Mol. Biol. 43, 241–267.
Huang, B., Bates, M., and Zhuang, X. (2009). Super-resolution fluorescence microscopy. Annu. Rev. Biochem. 78, 993–1016.
Ibrahim, R.K. (1977). Glucosylation of lignin precursors by uridine
diphosphate glucose: coniferyl alcohol glucosyltransferase in
higher plants. Z. Pflanzenphysiol. 85, 253–262.
Kirst, M., et al. (2003). Apparent homology of expressed genes from
wood-forming tissues of loblolly pine (Pinus taeda) with Arabidopsis thaliana. Proc. Natl Acad. Sci. U S A. 100, 7383–7388.
Klein, M., Martinoia, E., Hoffmann-Thoma, G., and Weissenbock, G.
(2000). A membrane-potential dependent ABC-like transporter
mediates the vacuolar uptake of rye flavone glucuronides: regulation of glucuronide uptake by glutathione and its conjugates.
Plant J. 21, 289–304.
Kuromori, T., et al. (2010). ABC transporter AtABCG25 is involved in
abscisic acid transport and responses. Proc. Natl Acad. Sci. U S A.
107, 2361–2366.
Lalonde, S., Ehrhardt, D.W., Loque, D., Chen, J., Rhee, S.Y., and
Frommer, W.B. (2008). Molecular and cellular approaches for
the detection of protein–protein interactions: latest techniques
and current limitations. Plant J. 53, 610–635.
Lanot, A., et al. (2006). The glucosyltransferase UGT72E2 is responsible for monolignol 4-O-glucoside production in Arabidopsis
thaliana. Plant J. 48, 286–295.
Lee, K.J., Marcus, S.E., and Knox, J.P. (2011). Cell wall biology: perspectives from cell wall imaging. Mol. Plant. 4, 212–219.
Lee, M., et al. (2008). The ABC transporter AtABCB14 is a malate
importer and modulates stomatal response to CO2. Nat. Cell Biol.
10, 1217–1223.
Leinhos, V., and Savidge, R.A. (1993). Isolation of protoplasts from
developing xylem of Pinus banksiana and Pinus strobos. Can. J.
For. Res. 23, 343–348.
Leinhos, V., Udagama-Randeniya, P.V., and Savage, R.A. (1994). Purification of an acidic coniferin-hydrolysing glucosidase from developing xylem of Pinus banksiana. Phytochem. 37, 311–315.
Leple, J.C., et al. (2007). Downregulation of cinnamoyl-coenzyme
A reductase in poplar: multiple-level phenotyping reveals effects
on cell wall polymer metabolism and structure. Plant Cell. 19,
3669–3691.
Lerouxel, O., Cavalier, D.M., Liepman, A.H., and Keegstra, K. (2006).
Biosynthesis of plant cell wall polysaccharides: a complex process. Curr. Opin. Plant Biol. 9, 621–630.
Ishii, T. (1997). Structure and functions of feruloylated polysaccharides. Plant Sci. 127, 111–127.
Li, X., Weng, J.K., and Chapple, C. (2008). Improvement of biomass
through lignin modification. Plant J. 54, 569–581.
Jasinski, M., Ducos, E., Martinoia, E., and Boutry, M. (2003). The
ATP-binding cassette transporters: structure, function, and gene
family comparison between rice and Arabidopsis. Plant Physiol.
131, 1169–1177.
Li, Y.H., Kajita, S., Kawai, S., Katayama, Y., and Morohoshi, N.
(2003). Down-regulation of an anionic peroxidase in transgenic
aspen and its effect on lignin characteristics. J. Plant Res. 116,
175–182.
Jasinski, M., Stukkens, Y., Degand, H., Purnelle, B., MarchandBrynaert, J., and Boutry, M. (2001). A plant plasma membrane
ATP binding cassette-type transporter is involved in antifungal
terpenoid secretion. Plant Cell. 13, 1095–1107.
Lim, E.K., Li, Y., Parr, A., Jackson, R., Ashford, D.A., and Bowles, D.J.
(2001). Identification of glucosyltransferase genes involved in
sinapate metabolism and lignin synthesis in Arabidopsis. J. Biol.
Chem. 276, 4344–4349.
Kaneda, M., et al. (2011). ABC transporters coordinately expressed
during lignification of Arabidopsis stems include a set of ABCBs
associated with auxin transport. J. Exp. Bot. 62, 2063–2077.
Liu, C.J., Miao, Y.C., and Zhang, K.W. (2011). Sequestration and
transport of lignin monomeric precursors. Molecules. 16,
710–727.
Kaneda, M., Rensing, K.H., Wong, J.C., Banno, B., Mansfield, S.D.,
and Samuels, A.L. (2008). Tracking monolignols during wood development in lodgepole pine. Plant Physiol. 147, 1750–1760.
Liu, L., Dean, J.F.D., Friedman, W.E., and Eriksson, K.-E.L. (1994).
A laccase like phenol oxidase is correlated with lignin biosynthesis in Zinnia elegans stem tissues. Plant J. 6, 213–224.
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
Gutierrez, R., Lindeboom, J.J., Paredez, A.R., Emons, A.M., and
Ehrhardt, D.W. (2009). Arabidopsis cortical microtubules position
cellulose synthase delivery to the plasma membrane and interact
with cellulose synthase trafficking compartments. Nat. Cell Biol.
11, 797–806.
Kang, J., et al. (2010). PDR-type ABC transporter mediates cellular
uptake of the phytohormone abscisic acid. Proc. Natl Acad. Sci.
U S A. 107, 2355–2360.
Liu
Lu, F., and Ralph, J. (2002). Preliminary evidence for sinapyl acetate
as a lignin monomer in kenaf. Chem. Commun. (Camb.). 1, 90–91.
Lu, F., Ralph, J., Morreel, K., Messens, E., and Boerjan, W. (2004).
Preparation and relevance of a cross-coupling product between
sinapyl alcohol and sinapyl p-hydroxybenzoate. Org. Biomol.
Chem. 2, 2888–2890.
Lu, Y.P., Li, Z.S., Drozdowicz, Y.M., Hortensteiner, S., Martinoia, E.,
and Rea, P.A. (1998). AtMRP2, an Arabidopsis ATP binding cassette transporter able to transport glutathione S-conjugates
and chlorophyll catabolites: functional comparisons with
Atmrp1. Plant Cell. 10, 267–282.
Marcinowski, S., Falk, H., Hammer, D.K., Hoyer, B., and Grisebach, H.
(1979). Appearance and localization of a beta-glucosidase hydrolyzing coniferin in spruce (Picea abies) seedlings. Planta. 144,
161–165.
Marcus, S.E., et al. (2010). Restricted access of proteins to mannan
polysaccharides in intact plant cell walls. Plant J. 64, 191–203.
McFarlane, H.E., Shin, J.J., Bird, D.A., and Samuels, A.L. (2010).
Arabidopsis ABCG transporters, which are required for export
of diverse cuticular lipids, dimerize in different combinations.
Plant Cell. 22, 3066–3075.
Miao, Y.C., and Liu, C.J. (2010). ATP-binding cassette-like transporters are involved in the transport of lignin precursors across
plasma and vacuolar membranes. Proc. Natl Acad. Sci. U S A.
107, 22728–22733.
Miernyk, J.A., and Thelen, J.J. (2008). Biochemical approaches for
discovering protein–protein interactions. Plant J. 53, 597–609.
Moller, R., Koch, G., Nanayakkara, B., and Schmitt, U. (2006). Lignification in cell cultures of Pinus radiata: activities of enzymes
and lignin topochemistry. Tree Physiol. 26, 201–210.
Nakashima, J., Mizuno, T., Takabe, K., Fujita, M., and Saiki, H.
(1997). Direct visualization of lignifying secondary wall thickenings in Zinna elegans cells in culture. Plant Cell Physiol. 38,
818–827.
Nilsson, R., Bernfur, K., Gustavsson, N., Bygdell, J., Wingsle, G., and
Larsson, C. (2010). Proteomics of plasma membranes from poplar
trees reveals tissue distribution of transporters, receptors, and
proteins in cell wall formation. Mol. Cell Proteomics. 9, 368–387.
Obrdlik, P., et al. (2004). K+ channel interactions detected by a genetic system optimized for systematic studies of membrane protein interactions. Proc. Natl Acad. Sci. U S A. 101, 12242–12247.
Onnerud, H., Zhang, L., Gellerstedt, G., and Henriksson, G. (2002).
Polymerization of monolignols by redox shuttle-mediated enzymatic oxidation: a new model in lignin biosynthesis. Plant Cell.
14, 1953–1962.
Paredez, A.R., Somerville, C.R., and Ehrhardt, D.W. (2006). Visualization of cellulose synthase demonstrates functional association
with microtubules. Science. 312, 1491–1495.
Patten, A.M., et al. (2010). Probing native lignin macromolecular
configuration in Arabidopsis thaliana in specific cell wall types:
further insights into limited substrate degeneracy and assembly
of the lignins of ref8, fah 1–2 and C4H::F5H lines. Mol. Biosyst. 6,
499–515.
Enigma of Plant Lignification
|
315
Paux, E., Tamasloukht, M., Ladouce, N., Sivadon, P., and GrimaPettenati, J. (2004). Identification of genes preferentially
expressed during wood formation in Eucalyptus. Plant Mol. Biol.
55, 263–280.
Pesquet, E., et al. (2005). Novel markers of xylogenesis in Zinnia are
differentially regulated by auxin and cytokinin. Plant Physiol.
139, 1–19.
Pickett-Heaps, J.D. (1968). Xylem wall deposition: radioautographic
investigations using lignin precursors. Protoplasma. 65, 181–205.
Pighin, J.A., et al. (2004). Plant cuticular lipid export requires an ABC
transporter. Science. 306, 702–704.
Pourcel, L., Routaboul, J.M., Kerhoas, L., Caboche, M., Lepiniec, L.,
and Debeaujon, I. (2005). TRANSPARENT TESTA10 encodes
a laccase-like enzyme involved in oxidative polymerization of
flavonoids in Arabidopsis seed coat. Plant Cell. 17, 2966–2980.
Ralph, J. (2007). Perturbing lignification. In The Compromised
Wood Workshop 2007, Entwistle, K., Harris, P.J., Walker, J., eds
(Canterbury, New Zealand: Wood Technology Research Centre,
University of Canterbury), pp. 85–112.
Ralph, J. (2010). Hydroxycinnamates in lignification. Phytochem.
Revs. 9, 65–83.
Ralph, J., et al. (2001). NMR evidence for benzodioxane structures
resulting from incorporation of 5-hydroxyconiferyl alcohol into
lignins of O-methyltransferase-deficient poplars. J. Agri. Food
Chem. 49, 86–91.
Ralph, J., et al. (2004a). Lignins: natural polymers from oxidative
coupling of 4-hydroxyphenylpropanoids. Phytochem. Rev. 3,
29–60.
Ralph, J., et al. (2004b). Peroxidase dependent cross-linking reactions of p-hydroxycinnamates in plant cell walls. Phytochem.
Rev. 3, 79–96.
Ralph, J., et al. (2008). Identification of the structure and origin of
a thioacidolysis marker compound for ferulic acid incorporation
into angiosperm lignins (and an indicator for cinnamoyl CoA reductase deficiency). Plant J. 53, 368–379.
Rasmussen, S., and Dixon, R.A. (1999). Transgene-mediated and
elicitor-induced perturbation of metabolic channeling at the entry point into the phenylpropanoid pathway. Plant Cell. 11,
1537–1551.
Reinhammar, B., and Malmstroem, B.G. (1981). In ‘Blue’ coppercontaining oxidases, Copper Proteins, Spiro,T.G., eds (New York,
NY: Wiley), pp. 109–149.
Richardson, A., Duncan, J., and McDougall, G.J. (2000). Oxidase activity in lignifying xylem of a taxonomically diverse range of
trees: identification of a conifer laccase. Tree Physiol. 20,
1039–1047.
Richardson, A., Stewart, D., and McDougall, G.J. (1997). Identification and partial characterization of a coniferyl alcohol oxidase
from lignifying xylem of Sitka spruce (Picea sitchensis). Planta.
203, 35–43.
Rippert, P., Puyaubert, J., Grisollet, D., Derrier, L., and Matringe, M.
(2009). Tyrosine and phenylalanine are synthesized within the
plastids in Arabidopsis. Plant Physiol. 149, 1251–1260.
Ruel, K., et al. (2009). Impact of CCR1 silencing on the assembly of
lignified secondary walls in Arabidopsis thaliana. New Phytol.
184, 99–113.
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
Marcinowski, S., and Grisebach, H. (1978). Enzymology of lignification: cell wall-bound b-glucosidase from coniferin from spruce
(Picea abies) seedlings. Eur. J. Biochem. 87, 37–44.
d
316
|
Liu
d
Enigma of Plant Lignification
Ruzicka, K., et al. (2010). Arabidopsis PIS1 encodes the ABCG37
transporter of auxinic compounds including the auxin precursor
indole-3-butyric acid. Proc. Natl Acad. Sci. U S A. 107,
10749–10753.
Sugiyama, A., Shitan, N., and Yazaki, K. (2007). Involvement of
a soybean ATP-binding cassette-type transporter in the secretion
of genistein, a signal flavonoid in legume–Rhizobium symbiosis.
Plant Physiol. 144, 2000–2008.
Samuels, A.L., Rensing, K.H., Douglas, C.J., Mansfield, S.D.,
Dharmawardhana, D.P., and Ellis, B.E. (2002). Cellular machinery
of wood production: differentiation of secondary xylem in Pinus
contorta var. Iatifolia. Planta. 216, 72–82.
Suh, S.J., et al. (2007). The ATP binding cassette transporter AtMRP5
modulates anion and calcium channel activities in Arabidopsis
guard cells. J. Biol. Chem. 282, 1916–1924.
Sánchez-Fernández, R., Davies, E.T.G., Coleman, J.O.D., and
Rea, P.A. (2001). The Arabidopsis thaliana ABC protein superfamily: a complete inventory. J. Biol. Chem. 276, 30231–30244.
Takabe, K.F.M., Harada, H., and Saiki, H. (1985). Autoradiographic
investigations of lignification in the cell walls of cryptomeria
(Cryptomeria japonica d. Don). Mokuzai Gakkaishi. 31, 613–619.
Takabe, K., Takeuchi, M., Sato, T., Ito, M., and Fujita, M. (2001). Immunocytochemical localization of enzymes involved in lignification of the cell wall. J. Plant Res. 114, 509–515.
Sasaki, S., Baba, K., Nishida, T., Tsutsumi, Y., and Kondo, R. (2006).
The cationic cell-wall-peroxidase having oxidation ability for
polymeric substrate participates in the late stage of lignification
of Populus alba L. Plant Mol. Biol. 62, 797–807.
Takahama, U. (1995). Oxidation of hydroxycinnamic acid and hydroxycinnamyl alcohol derivatives by laccase and peroxidase: interactions among p-hydroxyphenyl, guaiacyl and syringyl groups
during the oxidation reactions. Physiol. Plantarum. 93, 61–68.
Sasaki, S., Nishida, T., Tsutsumi, Y., and Kondo, R. (2004). Lignin
dehydrogenative polymerization mechanism: a poplar cell wall
peroxidase directly oxidizes polymer lignin and produces in vitro
dehydrogenative polymer rich in b-O-4 linkage. FEBS Lett. 562,
197–201.
Takeuchi, M., Takabe, K., and Fujita, M. (2001). Immunolocalization
of O-methyltransferase and peroxidase in differentiating xylem
of poplar. Holzforschung. 55, 146–150.
Sasaki, S., Nonaka, D., Wariishi, H., Tsutsumi, Y., and Kondo, R.
(2008). Role of Tyr residues on the protein surface of cationic
cell-wall-peroxidase (CWPO-C) from poplar: potential oxidation
sites for oxidative polymerization of lignin. Phytochem. 69,
348–355.
Savidge, R.A. (1988). A biochemical indicator of commitment to
tracheid differentiation in Pinus contorta. Can. J. Bot. 66,
2009–2012.
Savidge, R.A. (1989). Coniferin, a biochemical indicator of commitment to tracheid differentiation in conifers. Can. J. Bot. 67,
2663–2668.
Terashima, N., Fukushima, K., and Tsuchiya, S. (1986a). Heterogeneity in formation of lignin. VII: An autoradiographic study
on the formation of guaiacyl and syringyl lignin in poplar.
J. Wood Chem. Technol. 6, 495–504.
Terashima, N., Fukushima, K., and Takabe, K. (1986b). Heterogeneity in formation of lignin. VIII: An autoradiographic study on the
formation of guaiacyl and syringyl lignin in Magnolia kobus dc.
Holzforschung. 40, 101–105.
Terashima, N., Fukushima, K., He, L.-F., and Takabe, K. (1993). Comprehensive model of the lignified plant cell wall. In Forage Cell
Wall Structure and Digestibility, Jung, H.G., Buxton, D.R., Hatfield, R.D., Ralph, J., eds (Madison: ASA-CSSA-SSSA), pp. 247–270.
Shitan, N., and Yazaki, K. (2007). Accumulation and membrane
transport of plant alkaloids. Curr. Pharm. Biotechnol. 8, 244–252.
Terazawa, M., and Miyake, M. (1984). Phenolic compounds in living
tissue of woods. II. Seasonal variations of phenolic glycosides in
the cambial sap of woods. Mokuzai Gakkaishi. 30, 329–334.
Shitan, N., et al. (2003). Involvement of CjMDR1, a plant multidrugresistance-type ATP-binding cassette protein, in alkaloid transport in Coptis japonica. Proc. Natl Acad. Sci. U S A. 100, 751–756.
Thevenin, J., et al. (2011). The simultaneous repression of CCR and
CAD, two enzymes of the lignin biosynthetic pathway, results in
sterility and dwarfism in Arabidopsis thaliana. Mol. Plant. 4, 70–82.
Srere, P.A. (1987). Complexes of sequential metabolic enzymes.
Annu. Rev. Biochem. 56, 89–124.
Tognolli, M., Penel, C., Greppin, H., and Simon, P. (2002). Analysis
and expression of the class III peroxidase large gene family in
Arabidopsis thaliana. Gene. 288, 129–138.
Steeves, C., Förster, H., Pommer, U., and Savidge, R. (2001). Coniferyl alcohol metabolism in conifers. I. Glucosidic turnover of cinnamyl aldehydes by UDPG: coniferyl alcohol glucosyltransferase
from pine cambium. Phytochem. 57, 1085–1093.
Sterjiades, R., Dean, J.F., and Eriksson, K.E. (1992). Laccase from sycamore maple (Acer pseudoplatanus) polymerizes monolignols.
Plant Physiol. 99, 1162–1168.
Sterjiades, R., Dean, J.F.D., Gamble, G., Himmelsbach, D.S., and
Eriksson, K.E.L. (1993). Extracellular laccases and peroxidases
from sycamore maple (Acer pseudoplatanus) cell-suspension cultures: reactions with monolignols and lignin model compounds.
Planta. 190, 75–87.
Strader, L.C., and Bartel, B. (2009). The Arabidopsis PLEIOTROPIC
DRUG RESISTANCE8/ABCG36 ATP binding cassette transporter
modulates sensitivity to the auxin precursor indole-3-butyric
acid. Plant Cell. 21, 1992–2007.
Tommasini, R., et al. (1998). An ABC-transporter of Arabidopsis
thaliana has both glutathione-conjugate and chlorophyll catabolite transport activity. Plant J. 13, 773–780.
Tsuji, Y., and Fukushima, K. (2004). Behavior of monolignol glucosides in angiosperms. J. Agric. Food Chem. 52, 7651–7659.
Turlapati, P.V., Kim, K.W., Davin, L.B., and Lewis, N.G. (2011). The
laccase multigene family in Arabidopsis thaliana: towards
addressing the mystery of their gene function(s). Planta. 233,
439–470.
Umezawa, T. (2010). The cinnamate/monolignol pathway. Phytochem. Rev. 9, 1–17.
Van Doorsselaere, J., Baucher, M., Chognot, E., Chabbert, B., Tollier,
M.-T., Petit-Conil, M., Leplé, J.-C., Pilate, G., Cornu, D., Monties,
B., Van Montagu, M., Inzé, D., Boerjan, W., and Jouanin, L.
(1995). A novel lignin in poplar trees with a reduced caffeic
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
Sandhu, A.P., Randhawa, G.S., and Dhugga, K.S. (2009). Plant cell
wall matrix polysaccharide biosynthesis. Mol. Plant. 2, 840–850.
Liu
d
Enigma of Plant Lignification
|
317
acid/5-hydroxyferulic acid O-methyltransferase activity. Plant J.
8, 855–864.
Whetten, R., and Sederoff, R. (1995). Lignin biosynthesis. Plant Cell.
7, 1001–1013.
van Parijs, F.R., Morreel, K., Ralph, J., Boerjan, W., and Merks, R.M.
(2010). Modeling lignin polymerization. I. Simulation model of
dehydrogenation polymers. Plant Physiol. 153, 1332–1344.
Whetten, R., Sun, Y.H., Zhang, Y., and Sederoff, R. (2001). Functional genomics and cell wall biosynthesis in loblolly pine. Plant
Mol. Biol. 47, 275–291.
Vanholme, R., Demedts, B., Morreel, K., Ralph, J., and Boerjan, W.
(2010). Lignin biosynthesis and structure. Plant Physiol. 153,
895–905.
Whiting, P., and Goring, D.A.I. (1982). Chemical characterization of
tissue fractions from the middle lamella and secondary wall of
black spruce tracheids. Wood Sci. Technol. 16, 261–267.
Vanholme, R., Morreel, K., Ralph, J., and Boerjan, W. (2008). Lignin
engineering. Curr. Opin. Plant Biol. 11, 278–285.
Winkel-Shirley, B. (1999). Evidence for enzyme complexes in the
phenylpropanoid and flavonoid pathways. Physiol. Plantarum.
107, 142–149.
Verrier, P.J., et al. (2008). Plant ABC proteins: a unified nomenclature and updated inventory. Trends Plant Sci. 13, 151–159.
Zhao, Q., and Dixon, R.A. (2011). Transcriptional networks for lignin
biosynthesis: more complex than we thought? Trends Plant Sci.
16, 227–233.
Zhong, R., and Ye, Z.H. (2009). Transcriptional regulation of lignin
biosynthesis. Plant Signal. Behav. 4, 1028–1034.
Downloaded from http://mplant.oxfordjournals.org/ at FMRP/USP/BIBLIOTECA CENTRAL on September 23, 2014
Weng, J.K., Li, X., Bonawitz, N.D., and Chapple, C. (2008). Emerging strategies of lignin engineering and degradation for cellulosic biofuel production. Curr. Opin. Biotechnol. 19,
166–172.