You are on page 1of 19

Molecular Plant • Volume 3 • Number 1 • Pages 2–20 • January 2010 REVIEW ARTICLE

Phenylpropanoid Biosynthesis
Thomas Vogt1
Leibniz-Institute of Plant Biochemistry, Department of Secondary Metabolism, Weinberg 3, D-06120 Halle (Saale), Germany

ABSTRACT The general phenylpropanoid metabolism generates an enormous array of secondary metabolites based on
the few intermediates of the shikimate pathway as the core unit. The resulting hydroxycinnamic acids and esters are am-
plified in several cascades by a combination of reductases, oxygenases, and transferases to result in an organ and devel-
opmentally specific pattern of metabolites, characteristic for each plant species. During the last decade, methodology
driven targeted and non-targeted approaches in several plant species have enabled the identification of the participating
enzymes of this complex biosynthetic machinery, and revealed numerous genes, enzymes, and metabolites essential for
regulation and compartmentation. Considerable success in structural and computational biology, combined with the an-
alytical sensitivity to detect even trace compounds and smallest changes in the metabolite, transcript, or enzyme pattern,
has facilitated progress towards a comprehensive view of the plant response to its biotic and abiotic environment. Trans-
genic approaches have been used to reveal insights into an apparently redundant gene and enzyme pattern required for
functional integrity and plasticity of the various phenylpropanoid biosynthetic pathways. Nevertheless, the function and
impact of all members of a gene family remain to be completely established. This review aims to give an update on the various
facets of the general phenylpropanoid pathway, which is not only restricted to common lignin or flavonoid biosynthesis, but
feeds into a variety of other aromatic metabolites like coumarins, phenolic volatiles, or hydrolyzable tannins.

Key words: Phenylpropanoid; biosynthetic pathway; flavonoid; anthocyanin; tannin; coumarin; volatiles; lignin.

INTRODUCTION tases, and various superfamilies of transferases (Figure 1).


Phenylpropanoids contribute to all aspects of plant responses Some of these enzymes might exhibit overlapping specificities
towards biotic and abiotic stimuli. They are not only indicators in vitro, but their developmentally and spatially controlled ex-
of plant stress responses upon variation of light or mineral pression specifically contributes to tissue and plant-specific
treatment, but are also key mediators of the plants resistance chemical phenotypes.
towards pests (La Camera et al., 2004). They promote invasion In the last decade, knockout mutations and RNAi-mediated
of new habitats (Bais et al., 2003) and provide the biochemical suppression in Arabidopsis and other model plants have facil-
resources for successful reproduction (Dudareva et al., 2004). itated the identification and functional characterization of
Phenylpropanoid-based polymers, like lignin, suberin, or con- genes and enzymes (Alonso et al., 2003; Ossowski et al.,
densed tannins, contribute substantially to the stability and 2008). A plethora of mutants is available for candidate genes
robustness of gymnosperms and angiosperms towards me- central to the various branches of the phenylpropanoid path-
chanical or environmental damage, like drought or wounding. way. Identification of single loss-of-function mutants may not
The magnificent diversity of phenylpropanoids, in principle, be immediately accessible based on a clear phenotype such as
is the result of efficient modification and amplification of impaired growth or fertility (Briggs et al., 2006). In the absence
a very limited set of core structures, derived from the shikimate of any phenotype, high throughput screening of phenylpropa-
pathway (Herrmann, 1995). Aromatic amino acid lyases link noid patterns by GC– or UPLC–MS may enable the correlation
the pool of phenylalanine and tyrosine to biosynthetic path- of metabolite pattern with transcript profiles in an organ or
ways sometimes downgraded as secondary metabolism,
which, however, turn out to be equivalently relevant to plant
1
survival as photosynthesis or the citrate cycle. A clear differen- To whom correspondence should be addressed. E-mail tvogt@ipb-halle.de,
tiation between secondary and primary metabolites may seem fax +49 345 5582 1509, tel. +49 345 5582 1530.

obsolute (Hartmann, 2007). The diversity and plasticity of the ª The Author 2009. Published by the Molecular Plant Shanghai Editorial
Office in association with Oxford University Press on behalf of CSPP and
resulting natural products, more specifically the phenylpropa-
IPPE, SIBS, CAS.
noid profile, are then achieved by a set of enzymes organized doi: 10.1093/mp/ssp106, Advance Access publication 24 December 2009
usually in superfamilies, like oxygenases, ligases, oxidoreduc- Received 30 September 2009; accepted 19 November 2009
Vogt d Phenylpropanoid Biosynthesis | 3

like sinapate ester formation in Arabidopsis, have been clari-


fied in detail in the last decade and may have direct practical
impacts on current breeding programs, resulting in consumer
or farmer benefits (Bhinu et al., 2009). Several new ideas and
insights into the complex organization of pathways are due to
the advancement of analytical, chemical, and biomolecular
tools, like whole genome tilling arrays and application of fluo-
rescent biosensors (Herbst et al., 2009). These techniques will
be a critical driving force towards understanding plant biology
(Yonekura-Sakakibara et al., 2008; Zeller et al., 2009). They
promote rapid annotation of genes, the structural and func-
tional characterization of enzymes, realize quantification of
even trace amounts of metabolites, and enable identifica-
tion of signal transduction pathways in living cells towards
a comprehensive view of the plant organism.

Figure 1. Schematic View of the Diversity of the Secondary Metab-


olite Formation. THE CENTRAL PHENYLPROPANOID
PATHWAY
tissue specificity manner (Matsuda et al., 2008). In some cases, The plant shikimate pathway is the entry to the biosynthesis of
this approach may reveal the impact of a single phenylpropa- phenylpropanoids. Its intracellular, plastidial location and
noid pathway gene on other metabolic pathways (Rohde complex regulation have been investigated for more than a de-
et al., 2004). Metabolite profiling can be combined with immu- cade (Schmid and Amrhein, 1995). Regulation of the entire
nofluorescence or imaging techniques based on intrinsically pathway in plants can be performed at multiple levels.
fluorescent proteins, like GFP, to follow also the expression Transcriptional control was shown for 3-deoxy-D-arabinose-
and interaction of participating enzymes down to the cellular heptulosonate (DHAP) synthase (Herrmann and Weaver,
level (Dixit et al., 2006). 1999), whereas arogenate and prephenate dehydratase are
In Arabidopsis and poplar, nearly all genes required for phe- inhibited by phenylalanine, one end-product of the pathway
nylpropanoid and lignin formation are encoded by small gene (Yamada et al., 2008). The plastidal location of two elusive
families (Hamberger et al., 2007). Clustering of these sequen- steps of the shikimate pathway, arogenate dehydratase and
ces in cladograms suggests a variety of functions other than dehydrogenase, leading to phenylalanine and tyrosine biosyn-
the bona fide suggested roles of selected members of a gene thesis has recently been proven (Rippert et al., 2009). No recent
family based on sequence information. If genes of a single molecular evidence is available for the location of prephenate
pathway are compared, whole genome data indicate that in aminotransferase, although its activity within a plastidal frac-
poplar, bean, or alfalfa, the pattern of genes involved in fla- tion of Sorghum bicolor was shown (Siehl et al., 1986). There-
vonoid hydroxylation appears more complex as compared to fore, a location of the complete pathway within plastids
Arabidopsis (Hamberger et al., 2007). appears plausible. It is noteworthy that genes of the shikimate
In recent years, various excellent reviews summarized the pathway in plants apparently do not originate from a single
current knowledge on structural genes involved in phenylpro- prokaryotic ancestor of cyanobacterial origin, but are likely
panoid, specifically lignin and flavonoid formation, regulatory derived from at least three different sources. A complex,
transcription factors, hormonal control of the whole pathways multi-step shuffling of loss and gain of function occurred dur-
by jasmonate or auxin and evolution of pathway genes from ing phylogeny of this pathway, which might explain the mul-
primary metabolism (Weisshaar and Jenkins, 1998; Pichersky tiple factors contributing to the genomic organization and
and Gang, 2005; D’Auria and Gershenzon, 2005; Ober, 2005; expression of the pathway genes in Arabidopsis and other
Boudet, 2007). This report tries to emphasize that phenylpro- plants (Richards et al., 2006). With respect to its important con-
panoid derived metabolites comprise and contribute to multi- tribution to the carbon flow within the plant, the response of
ple biosynthetic branches other than lignin and flavonoid individual genes of the shikimate pathway to variations in
biosynthesis only (Figure 2). Some decisive steps in important light or nutrient content appears tightly regulated and more
pathways were resolved only recently, such as 2#-hydroxylation complex than the transcriptional responses of the entry genes
in coumarin biosynthesis (Kai et al., 2008). In a few cases, new of phenylpropanoid and flavonid biosynthesis, encoding phe-
organ-specific pathways could be established in the last couple nylalanine ammonia lyase (PAL) or chalcone synthase (CHS)
of years based on functional transcriptomics (Alves-Ferreira (Lillo et al., 2008).
et al., 2007; Ehlting et al., 2008) including the formation of ta- PAL and tyrosine ammonia lyase (TAL) catalyze the non-
petum-specific trisacyl-polyamine conjugates of Arabidopsis oxidative deamination of phenylalanine to trans-cinnamate
(Matsuno et al., 2009; Fellenberg et al., 2009). Other pathways, and direct the carbon flow from the shikimate pathway to
4 | Vogt d Phenylpropanoid Biosynthesis

CHO
H OH COOH
H OH + H2C
CH2O P O P

Erythose-4-Phosphat PEP

Gallotannins COOH COOH


+
Ellagitannins
HO OH O OH
OH OH

Gallic acid Dihydroshikimate

HO O O COOH COOH

COOH
CH3O

Coumarins HO OH O CH2
OH
Isochorismate
Shikimate

OH
HO O
HO O
COOH
COOH OH

OH O
OH OH O NH3+

Isoflavonoids Flavonoids Phenylalanine Salicylic acid

PAL

S CoA
OH OH COOH

O O
HO
OH
HO Cinnamate Benzoic acid
p-Coumaroyl CoA
OH
Stilbenes

OH

O glucose
HO O OH CH3O
O
OH

O CH3O HO
OH HO O
CH3O
Aurones
OH HO Phenylpropanoid esters
OH
Catechin Lignans Phenylpropenes
+
Cutin Acylated Polyamines
Lignin
Suberin
Sporopollein

Proanthocyanidins

Figure 2. Diversification of Phenylpropanoids Based on the General Phenylpropanoid Pathway.


The metabolites of the shikimate pathway and the central metabolite, 4-coumaroyl CoA, are shaded in gray.
Vogt d Phenylpropanoid Biosynthesis | 5

the various branches of the general phenypropanoid metab- encoding cytochrome P450s in Arabidopsis, 4-hydroxylation
olism. A crystal structure became available for a codon-opti- of trans-cinnamate to 4-coumarate is encoded by a single gene
mized, heterologously expressed enzyme from Petroselinum only, At2g30490, encoding CYP73A5. As expected, mutations
crispum (Ritter and Schulz, 2004). The structural data are con- in At2g30490 severely affect growth and development of the
sistent with the unusual deamination reaction, which is inde- plants and result in the accumulation of the quite unusual cin-
pendent of any co-factor. The required electrophilic prosthetic namoylmalate in leaves (Schilmiller et al., 2009).
group, 4-methylidene-imidazole-5-one, is auto-catalytically The subsequent step, encoded by the small gene family of
formed. These data also provide evidence for the requirements 4-CL ligases, channels the resulting aromatic CoA-esters to dif-
of antioxidative SH-protective reagent, like dithiothreitol ferent biosynthetic pathways. 4-coumaroyl CoA probably rep-
(DTT), for full enzyme activity in vitro, since an oxidized resents the most important branchpoint within the central
DTT-derivative was detected in the electron density map of phenylpropanoid biosynthesis in plants. It is either the direct
the crystal, when DTT was present into the crystallization precursor for flavonoid or H-lignin biosynthesis, or is fed into
buffer. Replacing the characteristic amino acid His89 of the ac- the production of increasingly methoxylated guaiacyl (G)- and
tive site of TAL in bacteria by phenylalanine, a characteristic syringyl (S)-monolignols. A decisive step is initiated by the en-
feature of PALs, converts TAL to PAL and vice versa (Watts zyme hydroxycinnamoyl CoA:shikimate/quinate hydroxycin-
et al., 2006; Louie et al., 2006). This is one rare example that namoyltransferase (HCT). HCT belongs to the large family of
a single amino acid exchange leads to a drastic change in sub- BAHD-like acyltransferases (D’Auria, 2006) and may require
strate specificity. From an evolutionary perspective, plant and regulation at multiple levels. Its down-regulation results in
microbial PALs and TALs are part of a superfamily of enzymes an increase in H-units, a depletion in S-units, and a significant
from plants, funghi, and bacteria, and are likely derived from increase in caffeoylquinate esters in stems, with only minor
a precursor of the widespread histidine ammonia lyase (HAL) effects in leaves (Hoffmann et al., 2004). Silencing of the
family in the histidine degradation pathway (Röther et al., SHT gene in tracheary elements forming Pinus radiata cell cul-
2002). Both can be distinguished from HALs by an additional tures demonstrated the importance of this gene for lignin
N-terminal domain and by a more effective shielding of their quantity and quality, most notably a drastic shift towards
active sites from molecular oxygen. the H-type of monolignols (Wagner et al., 2007).
Several copies of the PAL-genes are found in all plant spe- Only the shikimate and quinate esters, but not the free phe-
cies, comprising four genes in Arabidopsis, to five in poplar nylpropanoid acid or its CoA and glucose ester, are the sub-
and nine in rice (Hamberger et al., 2007). The individual genes strates of Cyp98A3, which catalyze the hydroxylation of the
may respond differentially to biotic and abiotic stressors and meta-position of the hydroxycinnamoyl shikimate or quinate
their expression is developmentally and spatially controlled esters, respectively (Schoch et al., 2001; Abdulrazzak et al.,
(Bhuiyan et al., 2009; Lillo et al., 2008). In poplar, an organ- 2006). Supported by reports of a high in vitro KM of HCT for
specific expression of three of the PAL-genes has been sugges- shikimate and quinate, two explanations for this fuzzy and ap-
ted—all display a different function (Kao et al., 2002). Two parently complicated meta-hydroxylation appear most likely:
genes are expressed in lignifying tissues; only one is involved (1) effective protection and channeling of the reactive cate-
in lignin formation. The second PAL-gene is specifically tar- chol moiety for subsequent methylation by CCoAOMT and
geted to condensed tannin formation. The third gene is asso- (2) a versatile metabolic control for channeling of the shiki-
ciated with flowering, although its functional role remains to mate pathway intermediates into the biosynthesis of second-
be established (Hamberger et al., 2007). In contrast to Arabi- ary metabolites only when photosynthesis is running well.
dopsis and poplar, in the Solanaceae, the entry step of the phe- Under sub-optimal photosynthesis, when plants are stressed,
nylpropanoid pathway is represented by a remarkable set of low concentration of shikimate pathway intermediates may
an estimated 20 putative PAL-genes, as shown for Lycopersicon re-direct the entire phenylpropanoid pathway towards the
esculentum (Chang et al., 2008). However, only a single gene production of phytoalexins, volatiles, flavonoids, and antho-
appears to be strongly expressed in all tissues, whereas the cyanins and de novo synthesis of (defence?) proteins (Schoch
remaining genes appear effectively silenced. This underutiliza- et al., 2006; Abdulrazzak et al., 2006). A sophisticated cross-
tion is peculiar and the precise genetic mechanisms of this talk and effective transport system between photosynthetic
extensive gene silencing remain to be established. active parenchyma cells and the cambial cells or tracheary ele-
The initial three steps of the pathway catalyzed by PAL, cin- ments of the vascular bundles would, however, be required.
namate 4-hydroxylase, and 4-coumaroyl CoA-Ligase are man- Caffeoyl quinate (chlorogenic acid) and caffeoyl phenyl lac-
datory and provide the basis for all subsequent branches and tic acid (rosmarinic acid) are the predominant soluble phenyl-
resulting metabolites. Subsequent steps downstream of PAL, propanoids in the Solanaceae and Lamiales, respectively.
leading to phenylpropanoid monomers (with exception of cin- Possible roles for both esters as defence compounds or as po-
namate 4-hydroxylase), are encoded by small gene families in tential antioxidants are discussed (Petersen et al., 2009). To ex-
all species investigated so far (Hamberger et al., 2007). With plore the potential benefits as a nutraceutical antioxidant,
respect to its impact on the phenylpropanoid metabolism, it overproduction of chlorogenic acid was achieved in tomato ei-
seems surprising that among nearly 300 functional genes ther by functional expression of a BAHD-like HCT (Niggeweg
6 | Vogt d Phenylpropanoid Biosynthesis

et al., 2004) or specifically in potato tubers by overexpressing tions of the plant. Partly due to the substrate specificity of
the StMtf1-MYB-transcription factor (Rommens et al., 2009). most 4CL-ligases earlier in the pathway, sinapoyl CoA is usually
Diversification of this small subset of phenylpropanoids like considered not to be a direct precursor for lignin monomer
coumaroyl- and caffeoyl CoA esters is achieved by an array of production—this may, however, depend on the species inves-
subsequent oxygenases, reductases, and transferases and tigated (Hamada et al., 2004). Comparison of annuals and per-
results in the observed complexity and heterogeneity of plant ennials with respect to lignin biosynthesis has revealed an
phenylpropanoid metabolite profiles. apparent functional redundancy of most participating genes
and enzymes. Although CAD is considered a key enzymatic
step in monolignol biosynthesis, a detailed analysis of 17
LIGNIN CAD encoding genes in Arabidopsis revealed that most inser-
Besides cellulose, lignin is the most prominent polymer on tion lines of individual CAD-genes only had a marginal effect
Earth. Lignin is nearly exclusively based on phenylpropanoid on the overall lignin composition (Kim et al., 2004). This sug-
units derived from the oxidative polymerization of hydroxycin- gests at least some functional redundancy among these genes
namoyl alcohol derivatives. As early as 1960, the contribution in case of Arabidopsis and single CADs are not rate-limiting.
of the shikimate pathway, the reduction of ferulic acid to con- On the other hand, it is less likely that all of these genes encode
iferyl alcohol, and the polymerization of monolignols had al- bona fide CADs and actually contribute to monolignol forma-
ready been realized (Neish, 1960). The term lignocellulose tion. A complex grid of temporal and organ-specific regulatory
implies that the phenolic polymer can be cross-linked to the mechanisms may exist to respond rapidly to various biotic and
cellulosic material by feruloyl-linkages, a characteristic feature abiotic stressors and ensure the growth and development un-
of grasses (Burr and Fry, 2009). der adverse conditions. The development of functional
The various aspects of lignin and lignan formation have genomics and genetic mapping of several species like poplar,
been summarized in a recent review by Davin et al. (2008), pine, eucalyptus, and alfalfa has already enabled to decipher
including a detailed biochemical evaluation of the require- the numerous possibilities of various metabolic routes towards
ments leading to controlled versus random coupling of lignin monomer and oligomer formation, oxidative phenol
4-hydroxyphenyl (H)-, guaiacyl (G)-, and syringyl (S)-derived coupling, and analysis of its polymeric structure (Boerjan
monolignols. This review also illustrates progress and experi- et al., 2003; Bhalerao et al., 2003; Davin et al., 2008).
mental limitations in structural elucidation of the various Remarkably, information on enzymatic reactions from con-
forms of lignin in monocots and dicots. Due to its economical iferyl alcohol leading to the various types of lignans and neo-
value for timber and biofuel formation, lignin biosynthesis lignans is limited. Data on the 8–8#-stereospecific coupling of
and manipulation has been a central research focus (Rubin, two coniferyl alcohol moieties resulting in the formation of
2008; Li et al., 2008). Numerous efforts to manipulate the lig- pinoresinol have been reported (Davin and Lewis, 2003).
nin content have largely focused on structural and regulatory The lignans, in contrast, received considerable interest based
genes leading the accumulation of gymnosperm and angio- on pharmacological properties, such as in the case of the cyto-
sperm lignin (Sederoff et al., 1999; Boudet et al., 2003; static and anti-cancerous, but also severely toxic, podophyllo-
Vanholme et al., 2008). The contribution of the corresponding toxin. (Enzymatically) controlled, regioselective 8–8’ coupling
enzymes leading to different monomer composition and of isoeugenol-derived lignans without an 9–9#-hydroxygroup
resulting in various ratios of guaiacyl or syringyl units have was recently demonstrated in roots of the Aristolochiaceae
been summarized and the effects of the knockout mutations Holystis reniformis upon feeding radiolabeled precursors
of the corresponding genes have been analyzed in detail (Messiano et al., 2009). The identification of two pinoresinol
(Anterola and Lewis, 2002; Dixon and Reddy, 2003; Vanholme reductase genes, catalyzing specifically the formation of lari-
et al., 2008). The activation of phenylpropanoid precursors, by ciresinol from pinoresinol in Arabidopsis, may facilitate a more
the initial formation of coenzyme A-thioesters and the subse- thorough investigation of the enzymatic steps of this pathway
quent reduction in two successive steps by cinnamoyl CoA re- on a molecular level in a model organism (Nakatsubo et al.,
ductase (CCR) and cinnamyol alcohol dehydrogenase (CAD), is 2008).
thoroughly documented (reviewed in Davin et al., 2008). Both In addition to the entry enzymes into lignin biosynthesis,
enzymes are considered the committed steps into monolignol, hydroxylation and methylation steps by cytochrome
lignan, and lignin biosynthesis. Arabidopsis knockout mutants P450s and S-adenosyl-L-methionine (AdoMet)-dependent
of the CCR1-gene show a reduced lignin formation, which can O-methyltransferases essentially determine the contribution
be partly compensated for by CCR-2, but they also display an of guaiacyl and syringyl monomers to the gymnosperm and
enhanced pool of feruloyl malate at the expense of sinapoyl angiosperm pattern. Overexpression of ferulate-5 hydroxylase
malate accumulation (Mir Derikvand et al., 2008). This again (F5H), also termed coniferylaldehyde hydroxylase, leads nearly
indicates that metabolic fluxes within a complex pathway exclusively to shift to syringyl-lignin in poplar, emphasizing the
can be, at least in part, redirected towards a pool of soluble critical role of this gene for the structure and physical proper-
metabolites by eliminating a single, central member of this ties of angiosperm lignin (Stewart et al., 2009), whereas knock-
pathway, without otherwise affecting the essential, vital func- down of the f5h gene reveals pleiotropic changes in the
Vogt d Phenylpropanoid Biosynthesis | 7

Arabidopsis metabolome (Huang et al., 2009). Decoration of biosynthesis, where hydroxycinnamic acids are activated by
the various hydroxylated monolignol precursors in Arabidopsis CoA-esters, in sinapic ester metabolism, 1-O-acyl-glucose esters
is performed either by cation-dependent CCoAOMT1, re- serve as the energy-rich metabolites during seed maturation
quired for the 3-O-methylation of caffeoyl-CoA, or by a cation- and germination (Mock and Strack, 1992). The corresponding
independent caffeic acid OMT (COMT), which is involved in glucosyltransferases have been investigated in molecular de-
the formation of sinapoyl residues and S-lignin from the tail (Milkowski et al., 2000; Sinlapadech et al., 2007; Meissner
5-hydroxyferuloyl derivatives (Davin et al., 2008). The role of et al., 2008). Two transferases, sinapoylglucose:choline sina-
CCoAOMT1, originally considered to be only associated with poyltransferase (SCT) and sinapoylglucose:malate sinapoyl-
monolignol formation, has changed considerably in the last transferase (SMT), are involved in the accumulation of
couple of years (Do et al., 2007; Kai et al., 2008; Fellenberg sinapine in the seed or sinapoylmalate in the seedlings, respec-
et al., 2009). Its involvement in various pathways, including tively. Compared with the presumed ancestors, the serine car-
sinapoylmalate formation, scopoletin biosynthesis, and meth- boxypeptidases (SCPs), the SMT from Arabidopsis, has
ylation of acylspermidine precursors, has proven its central maintained the catalytic triad, Ser–His–Asp (Milkowski and
role in the modification of phenylpropanoids. Engineering Strack, 2004; Stehle et al., 2009). A single Asp/Glu substitution
substrate promiscuous OMTs to act as monolignol 4-O- at the amino acid position close to the catalytic Ser enables
methyltransferases and impaire oxidative radical coupling is binding of 1-O-glucose (Figure 3). In the absence of L-malate,
a novel approach to modify lignin composition (Bhuiya and a low esterase activity resulting in the formation of sinapic acid
Liu, 2009). The key issue, if lignin formation proceeds via a met- can still be observed in this SCP-like protein, which is reminis-
abolic grid or by precise channeling of individual precursors cent of the hydrolytic ancestor. The drastic consequences of
through metabolons as proposed for other pathways in plant only slight differences in the catalytic site can be perceived
secondary metabolism, remains to be resolved (Humphreys as one of the driving forces resulting in a total pool of 51 genes
and Chapple, 2002; Jørgensen et al., 2005; Conrado et al., encoding SCPLs in Arabidopsis (Fraser et al., 2007). Even a small
2008). Based on current data, a clear decision favoring either subset of this reservoir of novel catalytic activities may already
one of both possibilities cannot be made. However, one might be able to amplify existing metabolites and prove advanta-
assume that precise channeling would be a superior mode of geous to the plant under changing environmental conditions.
control. Additional work is also required on the transport of The existence of acylated polyamine conjugates has been
monolignols prior to assembly. Two possibilities are discussed, described for several decades and their widespread occur-
either golgi-mediated vesicle transport (Samuels et al., 2002) rence, predominantly in flowers and pollen grains of wind-
or specific translocation of monolignols by unknown (ATP pollinated species, has been attributed to UV-protection, fer-
binding cassette?) transporters suggested for xylem formation tility, as well as defence (Meurer et al., 1988; Kakkar and Rai,
in angio- and gymnosperms (Ehlting et al., 2005; Kaneda et al., 1993; Walters, 2003). The rationale of their existence has been
2008).
Since the basic principles in lignin and lignocellulose forma-
tion may be similar in dicots and monocots, molecular work on
fast-growing crops with high biomass, severe drought resis-
tance, or low fertilizer requirements, like Miscanthus spec.,
should be encouraged to grow lignin and lignocellulose-
producing plants under extreme climatic conditions
(Dohleman and Long, 2009) and enable developing countries
to benefit from recent progress in biotechnology.

PHENYLPROPANOID ESTERS AND


AMIDES (ACYLATED POLYAMINES)
Arabidopsis has been an excellent tool to investigate the bio- Figure 3. Model Structure of the Active Site of the AtSMT Modified
synthesis of a specific class of phenylpropanoids characteristic from Stehle et al. (2006) with the Docking Interaction between the
Substrates Sinapoylglucose and L-malate (A) and the Mutated Pep-
for the Brassicaceae: the sinapic acid esters. Sinapoyl choline tidase Motif Gly–Asp–Ser–Tyr–Ser (B).
(sinapine) accumulates to high levels in the seeds of Brassica The amino acids replaced by the residues of serine carboxypepti-
napus (rapeseed), and, besides tannins and glucosinolates, dases are marked in green. The bulky side chain of Glu at position
can negatively affect its use as a protein supplement (Bhinu 172 would restrict binding of the glucose moiety of sinapoylglucose
et al., 2009). The reduction of sinapine by genetic engineering sterically. Therefore, the Glu to Asp substitution enables the acyl-
trasferases to accommodate 1-O-glucose esters as substrates and
of various biosynthetic enzymes was one of the driving forces
is the hallmark of glucose ester-dependent acyltransferases. Resi-
in the analysis of the pathway, but also led to important dis- dues forming the catalytic triad are colored in black. The substrates
coveries with respect to structural information and evolution sinapoylglucose and L-malate are colored in orange. Dotted lines
of the participating enzymes. In contrast to lignin or flavonoid indicate hydrogen bonds.
8 | Vogt d Phenylpropanoid Biosynthesis

disputed since then. The recent identification of phenylpropa- coumaric acid in lignin biosynthesis (Schoch et al., 2001).
noid polyamine conjugates in different organs of Arabidopsis The new, pathway-specific CYP98A8 and its paralog CYP98A9
wild-type and insertion lines by LC–MS analysis was accompa- belong to a monophyletic CYP98A clade and are duplications
nied by simultaneous identification of specific genes and of an ancestor of CYP98A3, which evolved rapidly via retropo-
enzymes in the acyl conjugate formation (Böttcher et al., sition and positive Darwinian selection (Matsuno et al., 2009).
2008; Matsuda et al., 2008; Luo et al., 2009). The rapid Knockout mutants of CYP98A8 display a shift in the major ac-
annotation of several genes and enzymes involved in the ylated spermidine conjugate pattern, from 5-hydroxyferuloyl
biosynthesis of these conjugates illustrates the combined to feruloyl conjugates. Likewise, AtTSM1, a cation-dependent
power of Arabidopsis knockout and knockdown insertion lines OMT with significant similarity to CCoAOMT1 in lignin biosyn-
with sensitive UPLC–MS analyses (Fellenberg et al., 2008; thesis, catalyzes the final step in acyl spermidine conjugate for-
Grienenberger et al., 2009). In the tapetum of Arabidopsis mation. The enzyme methylates a single 5-hydroxyferuloyl
flower buds, a new pathway for the biosynthesis of acylated moiety and yields N,N’-bis(5-hydroxyferuloyl)-N’’ sinapoyl sper-
spermidine conjugates was established. Hydroxycinnamic midine (Fellenberg et al., 2008). CCoAOMT1 and potentially
acids are transferred to spermidine by a BAHD-like spermidine CYP98A3 play additional roles earlier in methylation and
hydroxycinnamic acid transferase (SHT). Knockout mutation of hydroxylation of the quinate/shikimate esters, respectively.
this gene results in anthers and pollen grains, completely de- Corresponding knockouts of CCoAOMT1 also resulted in
void of trisacyl conjugates (Grienenberger et al., 2009). Two a severe decrease in acyl-spermidine formation (Figure 4;
other enzymes critical for this pathway, a cytochrome P450 Fellenberg et al., 2009). This pathway convincingly demon-
and a methyltransferase (Figure 4), were recruited from strates that genetic mechanisms like retroposition and gene
enzymes in an established biosynthetic pathway to catalyze duplication are also active in plant secondary metabolism
the precise modifications of acylspermidine derivatives. and may result in an accelerated establishment of organ-
CYP98A3 hydroxylates quinate and shikimate esters of 4- specific new pathways from ancestral genes (Matsuno et al.,
2009). The reason for the exclusive appearance of the pathway
in the tapetum is not yet understood, since neither the overall
plant phenotype nor the plant fertility appears affected
(Grienenberger et al., 2009). Regulatory elements of the
pathway have up to now not been identified, although can-
didate MYB transcription factors are described that regulate
phenylpropanoid–polyamine conjugate accumulation in other
plants (Gális et al., 2006).

ANTHOCYANINS, AURONES,
(ISO)FLAVONOIDS, AND STILBENES
Due to ease of visualization and control of mutants and
genetic imbalances, anthocyanin biosynthesis was one of
the first branches of the general phenylpropanoid metabo-
lism, for which biosynthetic enzymes and corresponding tran-
scription factors were identified and has been reviewed
extensively (Allan et al., 2008). The key enzymatic step of
the whole pathway, the formation of anthocyanidins via
(2R-3R)-dihydroflavonols and oxidation of (2R, 3S, 4S)-
leucoanthocyanidins by anthocyanidin synthase (ANS) has
recently been questioned due the apparent flavonol syn-
thase activity of Arabidopsis ANS and the identification of
(+)-catechin as a substrate of Gerbera hybrida ANS (Wellmann
Figure 4. A Simplified Version of the Biosynthetic Pathway Result-
et al., 2006). In a monomeric state, anthocyanins never display
ing in Acylated Spermidine Conjugates in flower buds Is Based on a blue color. Therefore, investigation of blue color formation
Recent Data by Fellenberg et al. (2008, 2009), Grienenberger et al. of anthocyanins either by pH-shift, metal complex formation,
(2009), and Matsuno et al. (2009). modification of the core unit by additional hydroxylation,
The exact sequence of the initial hydroxylation and Werner meth- methylation, or acylation and subsequent molecular stacking
ylation steps by CYP98A3 and CCoAOMT1, the subsequent hydrox-
has been a major focus of anthocyanin research in the last de-
ylation by Cyp98A9 and the transferase reaction by SHT is not
resolved yet. Sequential steps catalyzed by CYP98A9 and AtTSM1
cade (Yoshida et al., 2009). This research was partly driven by
are established. AtTSM1, localized exclusively in the tapetum, the cut-flower industry to develop blue varieties of those flow-
may serve as a tapetal marker in young flower buds of Arabidopsis. ers, which naturally do not exhibit this phenotype, like roses or
Vogt d Phenylpropanoid Biosynthesis | 9

carnations. During the process of developing a bluish tone, im- et al., 2009). UGT89C1 was linked to the late steps in flavo-
portant contributions to anthocyanin biosynthesis have been nol biosynthesis based on transcript and metabolite
made in terms of structural and regulatory genes. It was shown pattern and therefore correctly identified as Arabidopsis 7-
that roses, in contrast to most other plants, have a unique bi- O-rhamnosyltransferase (7RhaT) (Yonekura-Sakakibara et al.,
functional glucosyltransferase that catalyses the glycosylation 2007).
of both the 3- and the 5-OH group (Ogata et al., 2005). How- Several other late steps in the formation of the flavonoid
ever, they are devoid of a 5#-hydroxylase activity, which partly skeleton have only been poorly investigated. B-ring hydroxyl-
explains the lack of any natural bluish color. Expression of the ation at positions 6 and 8 of the flavonoid skeleton is limited to
Viola spp. 3#, 5#-hydroxylase and functional replacement of a few plant families only, like the Asteraceae. For Chrysanthe-
the endogenous rose dihydroflavonol reductase (DFR) for an mum segetum, Halbwirth and Stich (2006) identified a new
orthologous enzyme from Iris hollandica resulted in plants type of 8-monohydroxylase dependent on both NADPH and
that nearly exclusively accumulated delphinidin in the petals FAD, and was, therefore, clearly different from classical cyto-
and displayed a bluish, slightly purple tone (Katsumoto et al., chrome P450s. Obviously, there is a lack of enzymatic and ge-
2007). Myricetin, in addition to quercetin, was detected as netic information for this as well as other exotic plants.
a co-pigment in the transgenic plants, indicating that the hy- Scutellaria baicalensis is one of several prominent examples
droxylase modifies both flavonols and anthocyanins. Further used in traditional folk medicine. Its roots harbor flavonoids
modification by co-expression of an additional acyltransferase with unusual hydroxylation, methylation, or glycosylation pat-
did not result in the desired increase of acylated compounds tern in the flavonoid A- and B-ring (Li-Weber, 2009). Exponen-
and intensified color formation (Katsumoto et al., 2007). To tially growing databases and sequence information may
further enhance color formation, additional methylation of enable the identification and annotation of the corresponding
these B-ring hydroxyl groups may be required. Methylation genes and enzymes in due course, if we manage to prevent
of vicinal dihydroxy groups of anthocyanins from a petunidin these valuable resources from eradication.
to malvidin type was recently demonstrated to proceed by Three other elusive modifications or decorations of the fla-
a cation-dependent OMT in grapevine (Hugueney et al., vonoid molecules have been recently established at the molec-
2009). It further extends the biosynthetic niches of this type ular level. These modifications include the characterization
of enzyme towards modification of chromogenic compounds and identification of the first soluble, rare plant glucuronosyl-
with an aromatic cis-diol structure. transferase (GlurT) UGT94B1 from Bellis perennis (Sawada
Yellow flower colors are not necessarily due to carotinoid ac- et al., 2005). The key amino acid distinguishing this type of en-
cumulation, as in sunflowers. In Anthirrhinum majus, the yel- zyme from the common glycosyltransferases was identified as
low aurones display a bright yellow color with an additional an arginine at position 25 outside of the conserved UDP–sugar
fluorescence and are considered to help attract pollinators. binding-motif, the PSPG-box (Osmani et al., 2008).
Ono and co-workers (2006) were able to produce ornamental A second modification has been known to phytochemists
Torrenia hybrida flowers with bright yellow petals. This was for more than 20 years, but the corresponding enzymatic
achieved by the combined down-regulation of the anthocyanin step has not been resolved: the C-glycosylation of the
pathway and the co-expression of two genes involved in aurone flavonoid A-ring. Unequivocal annotation of flavone 6- and
formation, the vacuolar localized aureosidin synthase (AmSA1) 8 C-glycosylating enzyme was recently achieved for
and a cytoplasmatic chalcone 4#-O-glucosyltransferase. Co- a C-glucosyltransferase (CGT) from rice (Brazier-Hicks et al.,
epxression of the modifying UGT together with the AmSA1 2009). A biochemical and molecular approach was used in
was absolutely required, in order to enable transport of the the case of the Oryza sativa CGT. This CGT could not be iden-
4#-O-glycosylated chalcones to the vacuoles, where enzymatic tified by a simple sequence similarity approach, since no CGT-
conversion to the aurone 6-O-glucosides was performed (Ono specific amino acid sequence motif of OsCGT1 was detected
et al., 2006). and the most similar enzymes with 30% overall identity were
The complex structures of flavonoids and anthocyanins re- described as UGTs involved in monolignol formation in Arabi-
quire precise annotation of the participating modifying and dopsis. The authors propose that C-glycosylation activity in
decorating enzymes. Most acyl-, glucosyl-, and methyltrans- rice and wheat may proceed via a mechanism involving a 2-
ferases identified up to now exhibit overlapping or promiscu- hydroxyflavone intermediate, although this hydroxylation ac-
ous substrate specificities in vitro (Bowles et al., 2006). If tivity requires further confirmation. The 2-hydroxyflavone
a potential gene function is not immediately apparent by exists in equilibrium with its open chain form, which is pro-
sequence analysis of a multigene family, comparison of its posed as the substrate C-glucosyltransferase.
gene expression with the expression profiles of known genes, Last but not least, the cloning and characterization of enzy-
complementation of the metabolite pattern combined with matically active naringenin 8-prenyltransferase pave the way
reverse genetics help to annotate a single member from large for subsequent investigations of this important, yet poorly
superfamilies, like CYPs or UGTs. CYP98A8/A9 was correlated characterized class of membrane enzymes (Sasaki et al., 2008).
to the phenylpropanoid polyamine conjugate biosynthesis Regulation of flavonoid and anthocyanin formation by ba-
in the Arabidopsis tapetum (Ehlting et al., 2008; Matsuno sic TTG/basic helix-loop-helix bHLH and MybR2R3-type
10 | Vogt d Phenylpropanoid Biosynthesis

transcription factors has been the topic of many reports, such the 4#- hydroxyl group of 2,7,4#-trihydroxyisoflavanone at
as Stracke et al. (2007), and several reviews (Broun, 2005; Allan an initial step of the pathway as well as towards the 3-hydroxy
et al., 2008). The complex pattern of factors regulating antho- group of the pterocarpan 6a-hydroxymaackiain resulting in
cyanin formation has recently been extended by Matsui et al. the formation of pisatin has been identified and structurally
(2008) and Zhu et al. (2009). They discovered that single repeat characterized (Liu et al., 2006). Careful analysis of both sub-
R3 Myb transcription factors may also repress anthocyanin for- strates revealed their close to identical three-dimensional
mation, probably by competing with R2R3-Mybs for the bind- structure. A single enzyme able to convert two metabolites
ing sites of bHLH proteins, thereby down-regulating structural of the same pathway suggests that substrate promiscuity does
genes and counteracting, such as the Myb-like PAP1/2 en- not only provide evolutionary advantages in response to
hancer of anthocyanin formation in Arabidopsis (Gonzales novel, exogenous phytoalexins, but may also be maintained
et al., 2008). The bHLH-domain of the maize R-gene, partici- to serve those rare cases of metabolic serendipity. This situa-
pating in anthocyanin formation, links flavonoid formation tion is somewhat different from the methylation step in
to histone modification by interacting with an EMSY-like fac- acyl-spermidine formation in Arabidopsis flowers, where not
tor, which is involved in chromatin repair (Hernandez et al., CCoAOMT1, but a similar cation-dependent OMT (AtTSM1)
2007). with distinct properties was recruited to catalyze the final dec-
Identification of metabolite pattern and elucidation of the orating step of conjugate formation (Fellenberg et al., 2009).
structural genes and enzymes, deciphering the complex net- As in case of the isoflavonoids, also the stilbenes like resver-
work of hundreds of transcription factors and their hormonal atrol and its glycoside piceid are of limited distribution in the
control (Gális et al., 2006) in Arabidopsis and other plant spe- plant kingdom. Both have been shown to display numerous
cies, will provide a tremendous challenge to understand phe- biological activities in vitro, including antitumor activity and
nylpropanoid metabolism and its contribution to plant induction of apoptosis (Pervaiz, 2004; Crozier et al., 2009).
biology. Non-targeted approaches correlating co-expression Trans-resveratrol apparently retards the ageing process and
data and metabolomics at the organ or tissue level can provide increases longevity in various organisms from yeast to insects
solutions to unravel parts of this intriguing puzzle (Tohge and rodents (Orallo, 2008). Since dietary polyphenolics usually
et al., 2005). display low absorption and are chemically modified by the
Terpenophenolics, stilbenes, and isoflavonoids are limited small intestine, the contribution to human health in vivo will
to a few species only, but have attracted major interest due require profound epidemiological studies (Crozier et al., 2009).
to potential health-promoting affects on mammals (Stevens The relative ease to produce transgenic edible crops by intro-
and Page, 2004; Barnes and Prasain, 2005; Shen et al., 2009; ducing only a single gene, in this case stilbene synthase under
Dixon, 1999; Veitch, 2009). The biosynthesis of prenylchalcone the control of a 35S-promoter, results in substantial levels of
and other prenylated chalcones represent a deviation from the stilbenes in tomato fruits and may encourage other scientists
standard flavonoid biosynthesis. Xanthohumol is a major me- to explore the proposed positive in vitro effects in controlled
tabolite accumulating specifically in glandular trichomes of experimental set ups in vivo (Giovinazzo et al., 2005; Schijlen
hops, Humulus lupulus. By an EST-approach, Nagel et al. et al., 2006).
(2008) could recently show that in the glandular trichomes,
naringenin chalcone is first prenylated and subsequently
methylated by a substrate-specific 6#-O-desmethylxanthohu-
HYDROLYZABLE AND NON-
mol OMT, which is responsible for the stabilization of the
HYDROLYZABLE PROANTHOCYANIDINS
resulting xanthohumol. This result provides the basis for fur-
AND TANNINS
ther investigations of this important pathway, leading to Proanthocyanidin biosynthesis branches off the flavonoid
the formation of a family of compounds with considerable pathway from either 2,3-cis or trans flavanols and is initiated
phytoestrogenic properties. by anthocyanidin reductase or leucoanthocyanidin reductase,
The biosynthesis of isoflavonoid and related pterocarpans in respectively (Dixon et al., 2005; Tian et al., 2008). The resulting
model legumes like Medicago sativa and M. truncatula or Lo- epicatechin and catechin derivates can be oxidized to qui-
tus japonicus has been a major research focus for the last two nones, which are polymerized. However, it is not clear whether
decades. The oxidative migration of the aryl ring from 2S- this polymerization of the soluble precursors proceeds enzy-
liquiritigenin, the 5-deoxynaringenin formation catalyzed by matically by a laccase type of enzyme (encoded by the TT10
the cytochrome P450 isoflavone synthase and the subsequent gene in Arabidopsis) or non-enzymatically (Pourcel et al.,
4#-O-methylation are the decisive branch points leading to the 2005; Dixon et al., 2005). Transport of precursors across
formation of isoflavonoids. In contrast, stabilization of the iso- vacuoles membranes is facilitated by multidrug-extrusion
flavanone by a 4#-OMT, a dehydratase and second cytochrome MATE-type transporters in Arabidopsis (encoded by the
P450, isoflavone 2#-hydroxylase, are required for the pathway TT12-locus) and Medicago truncatula (Marinova et al., 2007;
to proceed towards the biosynthesis of pterocarpan-type phy- Zhao and Dixon, 2009). Similar to anthocyanin formation,
toalexins, like pisatin (Aoki et al., 2000). A promiscuous 4#- the deposition of proanthocyanidin oligomers in the seed coat
isoflavonoid OMT with identical catalytic properties towards of Arabidopsis is regulated by the interaction of a series of
Vogt d Phenylpropanoid Biosynthesis | 11

transcriptional regulators R2R3 MYB-type proteins (TT2), 2009). From the sum of data, it is evident that several path-
bHLH-factors (TT8), and WD40-repeat proteins (TTG1), which ways for benzoic acid biosynthesis may co-exist in a single
act synergistically. Overexpression of those transcription fac- plant and relative fluxes from one or the other pathway
tors may create improved crop varieties with enhanced proan- may be controlled by hormones or abiotic stressors. Feeding
thocyanidin contents (Baudry et al., 2004; Mellway et al., 2009; deuterium-labeled phenylalanine showed that metabolic
Terrier et al., 2009). The reported health-promoting activities fluxes of benzaldehyde formation of the non b-oxidative
of 3-O-esterified proanthocyanidins like epigallocatechin-gal- pathway from cinnamate to benzaldehyde to benzoic acid
late (EGCG) (Rezai-Zadeh et al., 2005), a major metabolite in was about twice that of the b-oxidative pathway involving
green tea leaves, can be hampered by a low bioavailability benzoyl-CoA in P. hybrida petals (Boatright et al., 2004).
upon modification of the core structure by sulfatation, meth- The 2-hydroxyl derivative of benzoic acid, salicylic acid (SA),
ylation, and glucuronylation (Henning et al., 2008). has long been known as a key component of plant innate immu-
Similar to the proanthocyanidins (non-hydrolyzable tan- nity (Lee and Raskin, 1999; Nawrath and Metreux, 1999) and also
nins), the hydrolyzable tannins, like galloyl-glucose esters plays a central role in systemic-acquired resistance (SAR) (Vlot
and ellagitannins, characterized by oxidative coupling of et al., 2008). Biochemical evidence for a direct pathway from
neighboring galloyl groups, display a strong tendency to scav- BA to SA was obtained first by the identification of benzoic acid
enge radicals. Due to multiple aromatic hydroxyl groups, they 2-hydroxylase, a cytochrome P450 monooxygenase from to-
cross-link and precipitate enzymes, making them interesting bacco (León et al., 1995), and was also based on an increase
defence compounds specifically when present in high concen- of PAL-transcripts and activity in tobacco plants upon tobacco
trations in oat or in sumac (Barbehenn et al., 2006; Gross, mosaic virus (TMV) infection (Ogawa et al., 2006). However,
2008). Remarkably little information is available on the biosyn- in contrast to the prokaryotic BA 2-hydroxylase encoded by
thesis of gallic acid, which respresents the core structure in the NahG gene, there is no Arabidopsis insertion line indicating
gallo- and ellagitannin structures. Retrobiosynthetic 13C- a similar gene and no CYP-candidate genes have been anno-
NMR-studies, by feeding 13C-labeled glucose to young leaves tated yet. Transgenic tobacco plants overexpressing PAL also
of Rhus typhina and measurments of oxygen isotope abun- did not show any significant change in the SA levels (Shadle
dance, performed by Werner and colleagues (1997, 2004), et al., 2003), suggesting that SA biosynthesis is not derived from
showed that gallic acid is derived from the shikimate pathway. trans-cinnamate and benzoic acid. In contrast, experimental ev-
The formation via phenylalanine and hydroxylated cinnamic idence from the Arabidopsis sid2 mutant, encoding isochoris-
acid derivatives could be ruled out by these experiments. Based mate synthase 1 (ICS1) (Wildermuth et al., 2001; Strawn et al.,
on these studies, it appears most likely that gallic acid is 2007), as well as virus-mediated RNA-silencing in N. benthami-
formed via dehydrogenation of 5-dihydroshikimate, although ana (Catinot et al., 2008) suggest that chorismate and ICS may be
a dehydratation step via protocatechuic acid and subsequent central to and required for salicylic acid biosynthesis. The exact
monooxygenation cannot be ruled out. Enzymatic evidence contribution of the several pathways to SA bioynthesis may
for the existence of a specific cytochrome P450 monooxyge- thereforebespecies-specificandresearchmayrequireacompar-
nase or a 2-oxoglutarate dependent dioxygenase as shown ative investigation of several mono- and dicot families.
in coumarin formation would help to resolve the contribution Modification of SA results in a significant increase in suscep-
of both possibilities. Clearly, a lot of research remains to be tibility to pathogens (Vlot et al., 2008). Modification of salicylic
done specifically on the hydrolyzable tannins, although enzy- acid can be achieved either by methylation (Ross et al., 1999) or
matic studies appear difficult due to the protein precipitant by glycosylation, at the carboxy- or at the 2-hydroxy group, re-
properties of both, substrates and products. spectively (Lee and Raskin, 1999; Lim et al., 2002). The meth-
yltransferase activity towards the carboxygroup of benzoic
acid and salicylic acid is part of the SABATH-family of enzymes
C6–C1 METABOLITES that are generally implicated in hormone homoeostasis or
In contrast to the established biosynthesis of phenylpropa- phenylpropanoid volatile formation in plants (Zhao et al.,
noids with a C6–C3 carbon skeleton, the formation of phe- 2008; Barkman et al., 2007; Kapteyn et al., 2007). Methyl salic-
nolics with a C6–C1 skeleton, like benzoic acid (BA), ylate has been identified as the critical mobile signal for SAR.
appears complex and is still fragmentary. Starting from Its perception may include additional regulatory steps, such as
trans-cinnamate, two biosynthetic scenarios are proposed. by a salicylic acid binding protein (SABP2), identified as a meth-
A b-oxidative pathway via CoA-activation and a 3-keto-acyl ylesterase (Forouhar et al., 2005; Zhao et al., 2009). Glycosyla-
CoA-thiolase was described for Petunia hybrida (Van tion of salicylic acid is much less investigated. A systematic
Moerkercke et al., 2009). An alternative non-oxidative path- screen within Arabidopsis plant natural product UGTs revealed
way, via benzoyl CoA and benzaldehyde, with a final step only enzymes with activities towards the carboxyl group, but
catalyzed by a benzaldehyde dehydrogenase, was identified none towards the 2-OH group (Lim et al., 2002). This suggests
in Anthirrhinum majus (Long et al., 2009). Additional routes, that regiospecific SA-glycosylating enzymes share limited dis-
specific for certain plants or specific organs, have also been tribution in plants and that glycosylation is not the common
described (Abd El-Mawla and Beerhues, 2002; Ibdah et al., way of inactivating this hormone.
12 | Vogt d Phenylpropanoid Biosynthesis

PHENYLPROPENE VOLATILES a drastic increase in phenylpropanoid volatile formation when


expressed in Petunia flowers (Ben-Zvi et al., 2008). This fasci-
Benzenoid volatile formation in Petunia flowers has been nating aspect of simultaneously enhancing flower color and
shown to be under strict diurnal and hormonal control, scent production illustrates only one part of the biotechnolog-
and appears to follow the transcriptional control of central ical prospects this kind of research may lead into for potential
shikimate pathway genes, like 5-enol-pyruvylshikimate-3- benefits to the consumer. Volatile emission from (genetically)
phosphate synthase (EPSPS) (Schuurink et al., 2006). Methyl enhanced scent producers might be used to ‘prime’ and
salicylate and methyl benzoate are not the only phenolic thereby protect non-transgenic crops against pests (Dudareva
compounds emitted as volatiles by flowers. They are part of and Pichersky, 2008) and pave the way to an alternative way of
a complex species-specific bouquet of compounds, including molecular farming approaches in an ongoing debate on the
terpenoids and phenylpropenes (Dudareva et al., 2004). These sense and non-sense of transgenic crops.
volatiles are also either emitted as floral attractants to pollina-
tors, as defence compounds against microorganisms, insects,
and mammalian predators, or can display allelopathic effects COUMARINS
on other plants (Dudareva et al., 2004; Horiuchi et al., 2007). High expression levels in all organs, from seeds to roots to
For phenylpropenes from the Lamiales and simple catechol flowers, suggested participation of CCoAOMT1 not only in lig-
derivatives, like orcinol from Chinese roses, several species nin biosynthesis, but in several biosynthetic pathways leading
and organ-specific biosynthetic pathways have been eluci- to soluble phenylpropanoids (Do et al., 2007). Knockout of the
dated in the last decade (Scalliet et al., 2006; Koeduka gene encoding CCoAOMT1 also led to a reduction in the cou-
et al., 2006; Kapteyn et al., 2007). The biosynthesis of anti- marin scopoletin in Arabidopsis roots (Kai et al., 2008), sug-
microbial phenylpropenes like eugenol, isoeugenol, or chavi- gesting that feruloyl CoA is the phenylpropanoid-derived
col could be resolved due to the high production levels of precursor of the UV-fluorescent coumarin scopoletin. The de-
these compounds either in the corolla of Petunia hybrida cisive step in coumarin biosynthesis in Arabidopsis roots was
and the peltate glands of Ocymum basilikum. Both com- verified by T-DNA insertion mutants (Kai et al., 2008). The
pounds are formed via the general phenylpropanoid path- key ortho-oxygenation step is not performed by a CYP-like
way, including the conversion of coniferyl alcohol by a monooxygenase, but by a position specific 2-oxoglutarate de-
BAHD-like coniferyl alcohol acetyltransferase and a subse- pendent dioxygenase (F6#H1). This enzyme catalyzes the for-
quent NADPH-dependent reduction of the coniferyl acetate mation of the 6#-O-hydroxylated trans-isomer from feruloyl
(Dexter et al., 2007; Koeduka et al., 2006). The structure of CoA, which undergoes isomerization of an enolate anion
the key enzyme, eugenol synthase (EGS), a member of the and lactonization to form scopoletin (Kai et al., 2008; Figure 5).
short chain dehydrogenases/reductases, catalyzing the reduc- The assumption that substrate specificities of orthologous
tive replacement of the propenyl side chain has been resolved F6#H1-dioxygenase-like enzymes affect the substitution pat-
by X-ray crystallography (Louie et al., 2007). Methylation of the tern of coumarins justifies investigations of exotic plant species
4-hydroxy group of phenylpropenes is achieved by a specific containing unusual highly methoxylated coumarins, such as in
subfamily of methyltransferases. They may be either localized the roots of Pelargonium sidoites (Kolodziej, 2007). This might
specifically in the glandular trichomes as observed for sweet lead to the detection of orthologous enzymes earlier in the
basil or could be part of the phenylpropene pathway in the phenylpropanoid pathway, with novel substrate specificities,
leaves, as in the case of Pimpinella anisum (anise) (Koeduka like unusual CoA-ligases. The resulting coumarin aglycones
et al., 2009). As far as coniferylacetate-derived phenylpropenes can be modified by a set of UDP-glucose-dependent glucosyl-
are concerned, a CCoAOMT1-like protein is again responsible transferases to result, for example, in scopolin or esculin, the
for the methylation of the 3-OH group, presumably at the most prominent soluble coumarins in plants. Alternatively, in
caffeoyl CoA level. Several levels of metabolic control at the Apiaceae, simple coumarins can be prenylated (Stanjek
the branch points of volatile emission are apparent when se- et al., 1999) and after a series of subsequent cytochrome
lected basil lines were tested by a systems biology approach P450-catalyzed reactions (Larbat et al., 2007), result in the for-
combining transciptome, proteome, and volatile metabolite mation of furanocoumarins, generally considered as photo-
profiles (Xie et al., 2008). With the exception of cinnamate toxic defense compounds.
4-hydroxylase, all other transcript and protein levels of phenyl-
propanoid pathway genes correlate well with metabolite for-
mation. Nevertheless, additional control by post-translational
EXOTIC PHENYLPROPANOIDS
phosphorylation or ubiquintinylation is described for PAL and Several phenylpropanoid-derived metabolites are restricted
chavicol OMT (CVOMT), respectively. to a few species or families only. These include the
Engineering of transcription factors has not only been used curcumins, benzophenones, biphenyls, or phenylphenale-
to control expression of flower color. Co-engineering scent nones. Although of exotic distribution, some possess pharma-
and flower color by a single transcription factor PAP1, known cological and health-promoting effects, which make them an
to regulate anthocyanin formation, surprisingly also led to interesting target for biosynthetic analyses (Bengmark et al.,
Vogt d Phenylpropanoid Biosynthesis | 13

O SCoA O SCoA O SCoA netic resources and metabolic diversity of plants with exotic
2-oxoglutarate
dependent phenylpropanoid-derived compounds are worth exploring
dioxygenase Rotation
Feruloyl-CoA to detect enzymes with unusual specificities for biotechnolog-
OH O
6' F6’H1 6' H ical or molecular farming approaches.
MeO MeO MeO
OH OH OH

O
MIXED POLYMERS, SUBERIN, CUTIN,
O SCoA AND SPOROPOLLENIN
O SCoA
MeO
O The contribution of the general phenylpropanoid metabolism
OH
SCoA towards the biosynthesis of mixed polymers is a challenging
O O
task for structural biochemists and physiologists. The complex-
MeO MeO ity of the pathways and robustness of cutin, suberin, or sporo-
CoAS OH OH
pollenin towards degradation requires harsh oxidative
chemical breakdown. Inherently, there is considerable chance
O
of artefact formation, which hampers the precise determina-
O tion of the contribution of the individual constituents of phe-
MeO
nylpropanoid metabolism to the structural integrity of these
OH polymers. These polymers essentially serve completely differ-
Scopoletin ent functions, as a water barrier for the sessile plant in case
of suberin and cutin, or as a highly protective and rigid scaffold
Figure 5. Recently Established Decisive Steps of Scopoletin Biosyn-
for the male gametophyte. Nevertheless, their complex and
thesis, Slightly Modified According to Kai et al. (2008).
variable, highly rigid structures are based on only two building
Starting from feruloyl–CoA, the 2-oxoglutarate dependent dioxy-
genase F6#H1 catalyzes the formation of the 6#-O-hydroxylated blocks, aliphatic and aromatic. The contribution of the latter is
trans-isomer from feruloyl CoA, which undergoes subsequent isom- apparently lower than in the case of lignin. The production of
erization, rotation, and lactonization to form scopoletin and free suberin is limited to a few plant tissues, like the endodermis
coenzyme A. with the casparian strip or the phellogen or cork cambium;
its principle components have been established (Bernards,
2002). As summarized in recent reviews by Pollard et al.
2009). The biosynthesis of curcumins with linear diarylhepta- (2008) and Franke and Schreiber (2007), besides linear ali-
noid structure is not sufficiently clarified. Specifically, the phatic, oxygenated long chain dicarboxylic acids and esters,
introduction of the oxygenation pattern, either before or after in Arabidopsis, feruloyl esters comprise a significant part of
the formation of the curcumin skeleton remains to be estab- the suberin structure. These are virtually absent in cutin. Poor
lished (Kita et al., 2008). That plant polyketide synthase type III substrate availability and problematic detection of intermedi-
(PKS III), which catalyzes the formation of bisdemethoxycurcu- ates render any biochemical investigations and identification
min, from p-coumaroyl and malonyl CoA, was described orig- of the participating enzymes of the phenylpropanoid pathway
inally from rice (Katsuyama et al., 2007). Its presence in the difficult. This includes the acyltransferase(s) connecting feru-
actual curcumin producing Curcuma longa has most recently loyl CoA to the aliphatic, oxygenated building blocks either
been demonstrated and the corresponding enzymes have in or outside the symplast. A genomic approach towards su-
been characterized (Katsuyama et al., 2009). Benzophenone- berin biosynthesis and cork differentiation in Quercus robur
synthase (BPS) and biphenyl-synthase (BIS) are specific PKS (cork oak) stresses the major contribution of two gene families
III, use benzoyl CoA as a starter, and initiate the carbon skel- in the assembly of this mixed polyester. Both the cytochrome
eton formation of the benzophenones, xanthones, biphenyls, P450 and the BAHD-like HC-tranferases (Soler et al., 2007). The
and dibenzofurans (Beerhues and Liu, 2009). The compounds exact mechanism, namely whether there is either coordinated
are considered as phytoalexins and their occurrence is limited or random enzymatic linkage of the phenylpropanoid-derived
to a few known plant taxa like the Hypericaceae (benzophe- aromatic blocks and the aliphatic esters in the apoplast, either
nones and xanthones) and the Maloideae (benzophenones catalyzed by peroxidases or by laccases, remains to be estab-
and biphenyls) (Beerhues and Liu, 2009). A new kind of type lished. Recent data demonstrate that a sterol glycosyltransfer-
III PKS was also suggested to be involved in the biosynthesis ase determines efficient trafficking of cutin and suberin to the
of cyclic diarylheptanoids and phenylphenalenons from Musa- Arabidopsis seed coat (De Bolt et al., 2009). Much less is known
ceae (Brand et al., 2006). The general phenylpropanoid path- about the regulatory networks affecting composition, trans-
way is not different from the pathway from established port, and assembly of both polymers. Candidates of R2R3s
species. This was demonstrated for xanthone biosynthesis and WRKY type transcription factors have been identified,
from the pharmaceutically relevant Hypericum perforatum but await further structural and functional characterization
(Franklin et al., 2009). Nevertheless, investigations of the ge- (Soler et al., 2007).
14 | Vogt d Phenylpropanoid Biosynthesis

Sporopollenin, the major part of the pollen exine, appar- and function should also consider in more detail the following
ently contains similar building blocks compared to cutin and key aspects:
suberin, namely phenylpropanoids and hydroxylated fatty
(1) A more detailed investigation of transport and shuttling of
acids. In contrast to cutin and suberin this polymer is even more
metabolites across membranes and cells.
resistant to chemical degradation, although advances in chem-
(2) Elucidation and identification of enzyme complexes and
ical analysis have enabled the complete identification and sep-
metabolon formation to understand how several pathways
aration of its constituents (Dominguez et al., 1999). Based on
may co-exist but are independently regulated.
labeling experiments, the presence of phenylpropanoid resi-
(3) A thorough understanding of the molecular mechanisms
dues is generally accepted (Schulze-Osthoff and Wiermann,
resulting in the formation of unusual phenylpropanoids in me-
1987). The mechanical strength and durability of the pollen
dicinal and exotic plants in a continuous search for new metab-
exine are probably due to a substantial amount of ether-
olites and new pathways.
linkages between fatty acid and/or phenylpropanoids. Conclu-
sive data on the assembly and the exact contribution of New analytical and biophysical methods, such as the implemen-
the phenylpropanoid conjugates are virtually non-existent. tation and application of nanotechnology to solve biological
Whether the variable phenylpropanoid pattern deposited as problems, are currently developed for new approaches towards
soluble pollenkit serves as a fingerprint of individual units in- resolving a most complex scenario of biological systems (Borch
corporated into the polymer is also unknown. Based on func- and Hamann, 2009; Zhong, 2009). Combining traditional meth-
tional characterization of Arabidopsis knockout mutants, ods with these new fascinating molecular and chemistry tools
initial progress has been made recently at least on the aliphatic will not only contribute to a better understanding of plant nat-
components of the polymer. Tissue-specific and developmen- ural product and phenylpropanoid biosynthesis, but encourage
tally controlled members of the cytochrome P450 family con- farmers and consumers to benefit from the diversity and plastic-
trol the hydroxylation pattern of the aliphatic acids, with ity of plant natural product formation and modification.
CYP703 and CYP704 unequivocally involved in hydroxylation
of lauric acid and omega-hydroxylation of longer chain fatty FUNDING
acids, respectively (Morant et al., 2007; Dobritsaa et al., 2009).
Financial support by the Deutsche Forschungsgemeinschaft (VO
The coordinate expression of several genes of the general phe-
719/8-1) is gratefully acknowledged.
nylpropanoid metabolism with CYP703 indicates the tight as-
sociation of fatty acid biosynthesis and phenylpropanoid
biosynthesis structures in the establishment of the exine bar-
ACKNOWLEDGMENTS
rier. The precise participation of phenylpropanoids in the pro- I would like to thank Felix Stehle (IPB Halle, Germany) for the contri-
posed ether and ester linkages remains to be established, bution of Figure 3. I am also grateful to Christin Fellenberg, Dieter
although it is already known that structures with vicinal dihy- Strack (both IPB Halle, Germany), Jörg Ziegler (University of Calgary,
droxy groups are required to meet the current perceptions of Canada),andtwoanonymousreviewersforcriticalreadingoftheman-
polymer organization (Morant et al., 2007). The functional uscript and many helpful suggestions. No conflict of interest declared.
aspects of phenylpropanoids in sporopollenin and exine for-
mation besides the contribution to UV-light protection may re- REFERENCES
side in protection against various pathogens. Biosynthesis and
transport of aliphatic and aromatic building blocks are Abd El Mawla, A.M., and Beerhues, L. (2002). Benzoic acid biosyn-
thesis in cell cultures of Hypericum androsaemum. Planta. 214,
performed by the tapetum in specific organelles, termed tape-
727–733.
tosoms (Hsieh and Huang, 2007), whereas the factors contrib-
Abdulrazzak, N., et al. (2006). A coumaroyl-ester-3-hydroxylase
uting to transport and polymerization are still unknown.
insertion mutant reveals the existence of nonredundant
meta-hydroxylation pathways and essential roles for phenolic
CONCLUDING REMARKS precursors in cell expansion and plant growth. Plant Physiol.
140, 30–48.
The diversity within the structural and regulatory genes of Allan, A.C., Hellens, R.P., and Laing, W.A. (2008). MYB transcription
several superfamilies of enzymes involved in secondary me- factors that colour our food. Trends Pl. Sci. 13, 99–102.
tabolism are the direct consequence of a complex plant– Alonso, J.M., et al. (2003). Genome-wide insertional mutagenesis of
environment interaction mediated in part by soluble and Arabidopsis thaliana. Science. 301, 653–657.
insoluble secondary metabolites, including phenylpropanoids. Alves-Ferreira, M., Wellmer, F., Banhara, A., Kumar, V.,
The rapid progress in sequencing, structural elucidation, and Riechmann, J.L., and Meyerowitz, E.M. (2007). Global expression
analytical tools will ultimately result in the desired systems bi- profiling applied to the analysis of Arabidopsis stamen develop-
ology approach, which may start with a few model plants and ment. Plant Physiol. 145, 747–762.
crops, like Arabidopsis, alfalfa, poplar, and rice at the begin- Aoki, T., Akashi, T., and Ayabe, S. (2000). Flavonoids of leguminous
ning. A thorough understanding of the biosynthetic pathway plants: structure, biological activity, and biosynthesis. J. Plant
leading to the various branches of phenylpropanoid formation Res. 113, 475–488.
Vogt d Phenylpropanoid Biosynthesis | 15

Anterola, A.M., and Lewis, N.G. (2002). Trends in lignin modifica- Böttcher, C., vRöpenack-Lahaye, E., Schmidt, J., Schmotz, C.,
tion: a comprehensive analysis of the effect of genetic manipu- Neumann, S., Scheel, D., and Clemens, S. (2008). Metabolome
lations/mutations on lignification and vascular integrity. analysis of biosynthetic mutants reveals diversity of metabolic
Phytochemistry. 61, 221–292. changes and allows identification of a large number of new com-
Bais, H.P., Vepachedu, R., Gilroy, S., Callaway, R.M., and pounds in A. thaliana. Plant Physiol. 147, 2107–2120.
Vivanco, J.M. (2003). Allelopathy and exotic plant invasion: from Boudet, A.M. (2007). Evolution and current status of phenolic com-
molecules and genes in species interaction. Science. 301, pounds. Phytochemistry. 68, 2722–2735.
1377–1380. Boudet, A.M., Kajita, S., Grima-Pettenatti, J., and Goffner, D. (2003).
Barbehenn, R.V., Jones, C.-P., Hagerman, A.K., Karonen, M., and Lignin and lignocellulosics: a better control of synthesis for few
Salminen, J.-P. (2006). Ellagitannins have greater oxidative activ- and improved uses. Trends Pl. Sci. 8, 576–581.
ities than condensed tannins and galloyl glucose at high pH: po- Bowles, D., Lim, E.K., Poppenberger, B., and Vaijstij, F.E. (2006). Gly-
tential impact on caterpillars. J. Chem. Ecol. 32, 2253–2267. cosyltransferases of small lipophilic molecules. Ann. Rev. Plant
Barkman, T.J., Martins, T.R., Sutton, E., and Stout, J.T. (2007). Pos- Biol. 57, 567–597.
itive selection for single amino acid change promotes substrate Brand, S., Hölscher, D., Schierhorn, A., Svatos, A., Schröter, J., and
discrimination of a plant volatile-producing enzyme. Mol. Biol. Schneider, B. (2006). A type III polyketide synthase from Wachen-
Evol. 24, 1320–1329. dorfia thyrsiflora and its role in diarylheptanoid and phenylphe-
Barnes, S., and Prasain, J. (2005). Current progress in the use of tra- nalenone biosynthesis. Planta. 224, 413–428.
ditional medicines and nutraceuticals. Curr. Opin Plant Biol. 8, Brazier-Hicks, M., Evans, K.M., Gershater, M.C., Puschmann, H.,
324–328. Steel, P.G., and Edwards, R. (2009). The C-glycosylation in cereals.
J. Biol. Chem. 284, 17926–17934.
Baudry, A., Helm, M.A., Dubreuce, B., Caboche, M., Weisshaar, B.,
and Lepiniec, L. (2004). TT2, TT8, and TTG1 synergistically specify Briggs, G.C., Osmont, K.S., Shindo, C., Sibout, R., and Hardtke, C.S.
the expression of BANYULS and proanthocyanidin bioynthesis in (2006). Unequal genetic redundancies in Arabidopsis:
Arabidopsis thaliana. Plant J. 39, 366–380. a neglected phenomenon? Trends Pl. Sci. 11, 492–498.
Beerhues, L., and Liu, B. (2009). Biosynthesis of biphenyl and ben- Broun, R. (2005). Transcriptional control of flavonoid biosynthesis:
zophenones: evolution of benzoic acid-specific type III polyke- a complex retwork of conserved regulators involved in multiple
tide synthases in plants. Phytochemistry. 70, 1719–1727. aspects of differentiation in Arabidopsis. Curr. Opin. Plant Biol. 8,
272–279.
Ben Zvi, M.M., Negre-Zakharov, F., Masci, T., Ovadis, M., Shklarman, E.,
Ben-Meir, H., Tzfira, T., Dudareva, N., and Vainstain, A. (2008). Inter- Burr, S.J., and Fry, S.C. (2009). Reluloylated arabinoxylans are oxi-
linking showy traits: co-engineering of scent and colour biosynthesis datively crosslinked by extracellular maize peroxidase but not
by horseradish peroxidase. Mol. Plant. 2, 883–892.
in flowers. Plant Biotechnol. J. 6, 403–415.
Catinot, J., Buchala, A., Abou-Mansour, E., and Metroux, J.-P.
Bengmark, S., Mesa, M.D., and Gil, A. (2009). Plant derived health:
(2008). Salicylic acid production in response to biotic and abiotic
the effect of turmeric and curcuminoids. Nutr. Hosp. 24, 273–281.
stress depends on isochorismate in Nicotiana benthamiana. FEBS
Bernards, M.A. (2002). Demystifying suberin. Can. J. Bot. 80,
Lett. 582, 472–478.
227–240.
Chang, A., Lim, M.-H., Lee, S.-W., Robb, E.J., and Nazar, R.N. (2008).
Bhalerao, R., Nilsson, O., and Sandberg, G. (2003). Out of the
Tomato phenylalanine ammonia-lyase gene family, highly redun-
woods; forest biotechnology enters the genomic area. Curr. dant but strongly underutilized. J. Biol. Chem. 283, 33591–33601.
Opin. Biotechnol. 14, 206–213.
Conrado, R.J., varner, J.D., and De Lisa, M.P. (2008). Engineering the
Bhinu, V.S., Schäfer, U.A., Li, R., Huang, J., and Hanoufa, A. (2009). spatial organization of metabolic enzymes: mimicking nature’s
Targeted reduction of sinapine biosynthesis pathway for seed synergy. Curr. Opin. Plant Biotech. 19, 492–499.
quality improvement in Brassica napus. Transgenic Res. 18,
Crozier, A., Jaganath, I.B., and Clifford, M.N. (2009). Dietary phe-
31–44.
nolics: chemistry, bioavailability and effects on health. Nat. Prod.
Bhuiya, M.W., and Liu, C.-J. (2009). Engineering monolignol 4-O- Rep. 26, 1001–1043.
methyltransferases to modulate lignin biosynthesi. J. Biol. Chem.
D’Auria, J.C. (2006). Acyltransferases in plants: a good time to be
doi/10.1074/jbc.M109.036673.
BAHD. Curr. Opin. Plant Biol. 9, 331–340.
Bhuiyan, N.H., Selvaraj, G., Wei, Y., and King, J. (2009). Gene expres-
D’Auria, J.C., and Gershenzon, J. (2005). The secondary metabolism
sion profiling and silencing reveal that monolignol biosynthesis of Arabidopsis thaliana: growing like a weed. Curr. Opin. Plant
plays a critical role in penetration defence in wheat against pow- Biol. 8, 308–316.
dery mildew invasion. J. Exp. Bot. 60, 509–521.
Davin, L.B., and Lewis, N. (2003). An historical perspective on lignan
Boatright, J., Negre, F., Chen, X., Kish, C.M., Wood, B., Pel, G., biosynthesis: monolignol, allylphenol, and hydroxycinnamic acid
Orlova, I., Gang, D., Rhodes, D., and Dudareva, N. (2004). Under- coupling and downstream mechanisms. Phytochem. Rev. 2,
standing in vivo benzenoid metabolism in Petunia petal tissue. 257–288.
Plant Physiol. 135, 1993–2011.
Davin, L.B., Jourdes, M., Patte, A.M., Kim, K.-K., Vassao, D.G., and
Boerjan, W., Ralph, J., and Baucher, M. (2003). Lignin biosynthesis. Lewis, N.G. (2008). Dissection of lignin macromolecular config-
Ann. Rev. Plant Biol. 54, 519–546. uration and assembly: comparision to related biochemical pro-
Borch, J., and Hamann, T. (2009). The nanodisc: a novel tool for cesses in allyl/prpopenyl phenol and lignan. Nat. Prod. Rep.
membrane protein studies. Biol. Chem. 390, 805–814. 25, 1015–1090.
16 | Vogt d Phenylpropanoid Biosynthesis

De Bolt, S., et al. (2009). Mutations in UDP-glucose sterol glucosyl- Franke, R., and Schreiber, L. (2007). Suberin: a biopolyester forming
transferase cause transparent testa phenotype and suberization apoplastic plant interfaces. Curr. Opin. Pl. Biol. 10, 252–259.
defects in seeds. Plant Physiol. 151, 78–87. o, L.F.R., Kombrink, E., and Dias, A.C.P. (2009).
Franklin, G., Conceica
Dexter, R., Qualley, A., Kish, C.M., Ma, C.J., Koeduka, T., Xanthone production in Hypericum performation cells provides
Nagegowda, D.A., Dudareva, N., Pichersky, E., and Clark, D. antioxidant and antimicrobial protection upon biotic stress. Phy-
(2007). Characterization of a Petunia acetyltransferase involved tochemistry. 70, 60–68.
in the biosynthesis of the floral volatile isoeugenol. Plant J. 49, Fraser, C.M., Thompson, M.G., Shirley, A.M., Ralph, J.,
265–275. Schönherr, J.A., Sinlapadech, T., Hall, M.C., and Chapple, C.
Dixit, R., Cyr, R., and Gilroy, S. (2006). Use of intrinsic fluorescent (2007). Related Arabidopsis serine carboxypeptidase-like sina-
proteins for plant cell imaging. Plant J. 45, 599–615. poylglucose: acyltransferases display distinct but overlapping
Dixon, R. (1999). The isoflavonoids: biochemsitry, molecular biology substrate specificities. Plant Physiol. 144, 1986–1999.
and biological functions. In Comprehensive Natural Products Gális, I., Šimek, P., Narisawa, T., Sasaki, M., Horiguchi, T., Fukuda, H.,
Chemistry, Sankawa, U., ed. (Amsterdam: Elsevier), pp. 773–823. and Matsuoka, K. (2006). A novel R2R3 MYB transcription factor
Dixon, R.A., and Reddy, M.S.S. (2003). Biosynthesis of monolignols: NtMYBJS1 is a methyl jasmonate-dependent regulator of
genomic and reverse genetics approaches. Phytochem. Rev. 2, phenylpropanoid-conjugate biosynthesis in tobacco. Plant J.
289–306. 46, 573–592.
Dixon, R.A., Xie, D.Y., and Sharma, S.B. (2005). Proanthocyanidins: Giovinazzo, G., D’Amico, I., Paradiso, A., Bollini, R., Sparvoli, F., and
a final fronter in flavonoid research? New Phytol. 165, 9–28. DeGara, L. (2005). Antioxidative metabolite profiles in tomato
Do, C.T., Pollet, B., Thévenin, J., Silbout, R., Denoue, D., Barriere, Y., constitutively expressing the grapevine stilbene synthase gene.
Lapierre, C., and Jouanin, L. (2007). Both caffeoyl coenzyme A Plant Biotechnol. J. 3, 57–68.
3-O-methyltransferase and caffeic acid O-methyltransferase 1 Gonzales, A., Zhao, M., Leavitt, L.M., and Lloyd, A.M. (2008). Reg-
are involved in redundant functions for lignin, flavonoids, and ulation of the anthocyanin biosynthetic pathway by the TTG1/
sinapoyl malate biosynthesis in A. thaliana. Planta. 226, bHLH/Myb transcriptional complex in Arabidopsis seedlings.
1117–1129. Plant J. 53, 814–827.
Dobritsaa, A.A., Shrestha, J., Morant, M., Pinot, F., Swanson, R., Grienenberger, E., Besseau, S., Geoffrey, P., Debayle, D., Heintz, D.,
Moeller, B.L., and Preuss, D. (2009). CYP704B1 is a long-chain Lapierre, C., Pollet, B., Heitz, T., and Legrand, M. (2009). A BAHD
fatty acid omega-hydroxylase essential for sporopollenin synthe- acyltransferase is expressed in the tapetum of Arabidopsis
sis in pollen of Arabidopsis. Plant Physiol. 151, 574–589. anthers and is involved in the synthesis of hydroxycinnamoyl
Dohleman, F.G., and Long, S.P. (2009). More productive than maize spermidines. Plant J. 58, 246–259.
in the mid-west: how does Miscanthus do it. Plant Physiol. 150, Gross, G.G. (2008). From lignins to tannins: forty years of enzyme
2104–2115. studies on the biosynthesis phenolic compounds. Phytochemis-
Dominguez, E., Mercado, J.A., Quesada, M., and Heredia, A. (1999). try. 69, 3018–3031.
Pollen sporopollenin: degradation and structural elucidation. Halbwirth, H., and Stich, K. (2006). An NADPH and FAD-dependent
Sex. Plant Reprod. 12, 171–178. enzyme catalyzes hydroxylation of flavonoids in position 8. Phy-
Dudareva, N., and Pichersky, E. (2008). Metabolic engineering of tochemistry. 67, 1080–1087.
plant volatiles. Curr. Opin. Biotechnol. 19, 181–189. Hamada, K., Nishida, T., Yamauchi, K., Fukushima, K., Kondo, R.,
Dudareva, N., Pichersky, E., and Gershenzon, J. (2004). Biochemistry and Tsutsumi, Y. (2004). 4-coumarate:coenzyme A ligase in black
of plant volatiles. Plant Physiol. 135, 1893–1902. locust (Robinia pseudacacia) catalyses the conversion of sinapate
Ehlting, J., et al. (2005). Global transcript profiling of primary stems to sinapoyl–CoA. J. Plant Res. 117, 303–310.
from Arabidopsis thaliana identifies candidate genes for missing Hamberger, B., Ellis, M., Friedmann, M., de Azevedo Sousa, C.,
links in lignin biosynthesis and transcriptional regulators of fiber Barbazuk, B., and Douglas, C. (2007). Genome-wide analyses
differentiation. Plant J. 42, 618–640. of phenylpropanoid-related genes in Populus trichocarpa, Ara-
Ehlting, J., Sauveplane, V., Olry, A., Ginglinger, J.F., Provart, N.J., bidopsis thaliana, and Oryza sativa: the Populus lignin toolbox
and Werck-Reichhart, D. (2008). An extensive (co-)expression and conservation and diversification of angiosperm gene fami-
analysis tool for the cytochrome P450 superfamily in Arabidopsis lies. Can J. Bot. 85, 1182–1201.
thaliana. BMC Plant Biology. 8, 1–19. Hartmann, T. (2007). From waste products to ecochemicals: 50 years
Fellenberg, C., Böttcher, C., and Vogt, T. (2009). Phenylpropanoid research on plant secondary metabolism. Phytochemistry. 68,
conjugate biosynthesis in flower buds of Arabidopsis thaliana. 2831–2846.
Phytochemistry. 70, 1392–1400. Henning, S.M., Choo, J.J., and Heber, D. (2008). Nongallated com-
Fellenberg, C., Milkowski, C., Hause, B., Lange, P.-R., Böttcher, C., pared with gallated flavan-3-ols in green and black tea are more
and Vogt, T. (2008). Tapetum specific location of a cation- bioavailable. J. Nutr. 138, 1529–1534.
dependent O-methyltransferase in A. thaliana. Plant J. 56, Herbst, K.J., Ni, Q., and Zhang, J. (2009). Dynamic visualization of
132–145. signal transduction in living cells: from second messengers to
Forouhar, F., et al. (2005). Structural and biochemical studies iden- kinases. IUBMB Life. 61, 902–908.
tify tobacco SABP2 as a methyl salicylate esterase and implicate it Hernandez, J.M., Feller, A., Morohashi, K., Frame, K., and
in plant innate immunity. Proc. Natl Acad. Sci. U S A. 102, Grotewold, E. (2007). A basic helix-loop-helix domain of maize
1773–1778. R links transcriptional regulation and histone modifications by
Vogt d Phenylpropanoid Biosynthesis | 17

recruitment of an EMSY-related factor. Proc. Natl Acad. Sci. U S A. Katsumoto, Y., et al. (2007). Engineering of the rose flavonoid bio-
104, 17222–17227. synthetic pathway successfully generated blue-hued flowers ac-
Herrmann, K. (1995). The shikimate pathway as an entry to aromatic cumulating delphinidin. Plant Cell. Rep. 48, 1589–1600.
secondary metabolism. Plant Physiol. 107, 7–12. Katsuyama, Y., Kita, T., and Horinouchi, S. (2009). Identification and
Herrmann, K., and Weaver, L.M. (1999). The shikimate pathway. characterization of multiple curcumin synthases from the herb
Ann. Rev. Plant Physiol. Plant Mol Biol. 50, 472–503. Curcuma longa. FEBS Lett. 583, 2799–2803.
Hoffmann, L., Besseau, S., Geoffrey, P., Ritzenthaler, C., Meyer, D., Katsuyama, Y., Matsuzawa, M., Funa, N., and Horinuchi, S. (2007).
Lapierre, C., Pollet, B., and Legrand, M. (2004). Silencing of In vitro synthesis of curcuminoids by type III polyketide synthase
hydroxycinnamoyl-coenzyme A shikikmate/quinate hydroxycin- from rice. J. Biol Chem. 282, 37702–37709.
namoyltransferase affects phenylpropanoid biosynthesis. Plant Kim, S.J., Kim, M.R., Bedgar, D.L., Moinuddin, S.G.A., Cardenas, C.L.,
Cell. 16, 1446–1465. Davin, L.B., and Lewis, N. (2004). Functional reclassification of
Horiuchi, J., Badri, D.V., Kimball, B.A., Negre, F., Dudareva, N., the putative cinnamyl alcohol dehydrogenase multigene family
Paschke, M.W., and Vivanca, J.M. (2007). The floral volatile in Arabidopsis. Proc. Natl Acad. Sci. U S A. 101, 1455–1460.
methyl benzoate from snapdragon (Anthirrhinum majus) Kita, T., Imai, S., Sawasa, H., Kumagai, H., and Seto, H. (2008). The
triggers phytotoxic effects in Arabidopsis thaliana. Planta. biosynthetic pathway in tumeric (Curcumis longa) as revealed by
13
226, 1–10. C-labelled precursors. Biosc. Biotechn. Biochem. 72, 1789–1798.
Hsieh, K., and Huang, A.H. (2007). Tapetosomes in Brassica tapetum Koeduka, T., Baiga, T.J., Noel, J.P., and Pichersky, E. (2009). Biosyn-
accumulate endoplasmic reticulum-derived flavonoids and thesis if t-anthole in anise: caracterisation of t-anol/isoeugenol
alkanes for delivery to the pollen surface. Plant Physiol. 19, synthase and an O-methyltransferase specific for a C7–C8 pro-
582–596. penyl side chain. Plant Physiol. 149, 384–394.
Huang, J., Bhinu, V.-S., Li, X., Dallal-Bashi, Z., Zhou, R., and Koeduka, T., et al. (2006). Eugenol and isoeugenol, characteristic
Hanoufa, A. (2009). Pleiotropic changes in Arabidopsis f5h aromatic constitutents of spices, are biosynthesized via reduc-
and sct mutants revealed by large-scale gene expression and me- tion of a coniferyl alcohol ester. Proc. Natl Acad. Sci. U S A.
tabolite analysis. Planta. 230, 1057–1069. 103, 10128–10133.
Hugueney, P., Provenzano, S., Verriès, C., Ferrandino, A., Kolodziej, H. (2007). Fascinating metabolic pools of Pelargonium
Meudec, E., Batelli, G., Merdinoglu, G., Cheynier, V., sidoites and Pelargonium reniforme, traditional and phytomed-
Schubert, A., and Ageorges, G. (2009). A novel cation-dependent ical sources of the herbal medicine Umckaloabo. Phytomedicine.
O-methyltransferase involved in anthocyanin methylation in 14, 9–17.
grapevine. Plant Physiol. 150, 2057–2070. La Camera, S., Gouzerh, G., Dhondt, S., Hoffmann, L., Frittig, B.,
Humphreys, J.M., and Chapple, C. (2002). Rewriting the lignin road- Legrand, M., and Heitz, T. (2004). Metabolic reprogramming
map. Curr. Opin. Plant Biol. 5, 224–229. in plant innate immunity: the contributions of phenylpropanoid
Ibdah, M., Chen, Y.t, Wilkerson, C.G., and Pichersky, E. (2009). An and oxylipin pathways. Immunol. Rev. 198, 267–284.
aldehyde oxidase in developing seeds of Arabidopsis converts Larbat, R., Kellner, S., Specker, S., Hehn, A., Gontier, E., Hans, J.,
benzaldehyde to benzoic acid. Plant Physiol. 150, 416–423. Bourgaud, F., and Matern, U. (2007). Molecular cloning and func-
Jørgensen, K., Rasmussen, A.V., Morant, M., Nielson, A.H., tional characterization of psoralene synthase, the first commit-
Bjarnholt, N., Zagrobelny, M., Bak, S., and Møller, B.L. (2005). ted monooxygenase in furanocoumarin biosynthesis. J. Biol.
Metabolon formation and metabolic channeling on the biosyn- Chem. 282, 542–554.
thesis of plant natural products. Curr. Opin. Plant. Biol. 8, Lee, H.-I., and Raskin, I. (1999). Purification, cloning, and expression
280–291. of a pathogen-inducible UDP-glucose:salicylic acid glucosyltrans-
Kai, K., Mizutani, M., Kawamura, N., Yamamoto, R., Tamai, M., ferase from tobacco. J. Biol. Chem. 274, 36637–36642.
Yamaguchi, H., Skata, K., and Shimizu, B. (2008). Scopoletin is León, J., Shulaev, V., Yalpani, N., Lawton, M.A., and Raskin, I. (1995).
biosynthesized via ortho-hydroxylation of feruloyl CoA by a 2- Benzoid acid 2-hydroxylase, a soluble oxygenase from tobacco
oxolutarate-dependent dioxygenase in Arabidopsis thaliana. catalyzes salicylic acid biosynthesis. Proc. Natl Acad. Sci. U S A.
Plant J. 55, 989–999. 92, 10413–10417.
Kakkar, R.K., and Rai, V.K. (1993). Plant polyamines in flowering and Li, X., Weng, J.K., and Chapple, C. (2008). Improvement of biomass
fruit development. Phytochemistry. 33, 1281–1288. through lignin modification. Plant J. 54, 569–581.
Kaneda, M., Rensing, K.H., Wong, J.C.T., Banno, B., Mansfield, S.D., Lillo, C., Lea, U.S., and Ruoff, P. (2008). Nutrient depletion as a key
and Samuels, A.L. (2008). Tracking monolignol during wood de- factor for manipulating gene expression and product formation
velopment in lodgepole pine. Plant Physiol. 147, 1750–1760. in different branches of the flavonoid pathway. Plant Cell Envi-
Kao, Y.Y., Harding, S.A., and Tsai, C.-J. (2002). Differential expres- ronm. 31, 582–601.
sion of two distinct phenylalanine ammonia-lyase genes in con- Lim, E.-K., Doucet, C.J., Li, Y., Elias, L., Worrall, D., Spencer, S.P.,
densed tannin-accumulating and lignifying cells of quaking Ross, J., and Bowles, D.J. (2002). The activity of Arabidopsis gly-
aspen. Plant Physiol. 130, 796–807. cosyltransferases toward salicylic acid, 4-hydroxybenzoic acids
Kapteyn, J., Qualley, A.V., Xie, Z., Fridman, E., Dudareva, N., and and other benzoates. J. Biol. Chem. 277, 586–592.
Gang, D.R. (2007). Evolution of cinnamate/p-coumarate carboxyl Liu, C.-J., Deavours, B.E., Richard, S.B., Ferrer, J.-L., Blount, J.W.,
methyltransferases and their role in the biosynthesis of methyl- Huhman, D., Dixon, R.A., and Noel, J.P. (2006). Structural basis
cinnamate. Plant Cell. 19, 3212–2229. for dual functionality of isoflavonoid O-methyltransferase in
18 | Vogt d Phenylpropanoid Biosynthesis

the evolution of plant defence response. Plant Cell. 18, direction of the phenylpropanoid pathway to feruloylmalate in
3656–3669. Arabidopsis mutants defect in cinnamoyl-CoA reductase I.
Li-Weber, M. (2009). New therapeutic aspects of flavones: the an- Planta. 227, 943–956.
ticancer properties of Scutellaria and its main active constituents Mock, H.P., and Strack, D. (1992). Energetics of uridine 5-diphopho-
Wogonin, Baicalein and Baicalin. Cancer Treat. Rev. 35, 57–68. glucose: hydroxycinnamic acid glucosyltransferase reaction. Phy-
Long, M.C., Nagegowda, D.A., Kaminaga, Y., Ho, K.K., Kish, C.M., tochemistry. 32, 575–579.
Schnepp, J., Sherman, D., Weiner, H., Rhodes, D., and Morant, M., Jørgenson, K., Schaller, H., Pinot, F., Møller, B.L.,
Dudareva, N. (2009). Involvement of snapdragon benzaldehyde Werck-Reichardt, D., and Bak, S. (2007). CYP703 is an an-
dehydrogenase in benzoic acid biosynthesis. Plant J. 59, 256–265. cient cytochrome P450 in land plants catalyzing in-chain hy-
Louie, G.V., Baiga, T.J., Bowman, M.E., Koeduka, T., Taylor, J.H., droxylation of lauric acid to provide building blocks for
Spassova, S.M., Pichersky, E., and Noel, J.P. (2007). Structure sporopollenin synthesis in pollen. Plant Cell. 19, 1473–1487.
and reaction mechanism of basil eugenol synthase. PLoS ONE. Nagel, J., Culley, L.K., Lu, Y., Liu, E., Matthews, P.D., Stevens, J.F.,
2 (10), e993. doi:10.1371/journal.pone.0000993. and Page, J.E. (2008). EST analysis of hop glandular trichomes
Louie, G.V., Bowman, M.E., Moffitt, M.C., Baiga, T.J., Moore, B.S., identifies an O-methyltransferase that catalyzes the biosynthesis
and Noel, J.P. (2006). Structural determinants and modulation of xanthohumol. Plant Cell. 20, 186–200.
of substrate specificity in phenylalanine ammonia lyases. Chem. Nakatsubo, T., Mizutani, M., Suzuji, S., Hattori, T., and Umezawa, T.
Biol. 13, 1327–1338. (2008). Characterization of Arabidopsis thaliana pinoresinol re-
Luo, J., Fuell, C., Parr, A., Hill, L., Bailey, P., Elliott, K., Fairhurst, S.A., ductase, a new type of enzyme involved in lignan biosynthesis. J.
Martin, C., and Michael, A.J. (2009). A novel polyamine acyltrans- Biol. Chem. 283, 15550–15557.
ferase responsible for the accumulation of spermidine conju- Nawrath, C., and Métreux, J.-P. (1999). Salicylic acid induction-
gates in Arabidopsis seed. Plant Cell. 21, 318–333.
deficient mutants of Arabidopsis thaliana express PR-2 and
Marinova, K., Pourcel, L., Weder, B., Schwarz, M., Barron, D., PR-5 and accumulate high levels of camalexin after pathogen
Routaboul, J.M., Debeaujom, I., and Klein, M. (2007). The Arabi- inoculation. Plant Cell. 11, 1393–1404.
dopsis MATE transporter TT12 acts as a flavonoid/H+-antiporter
Neish, A.C. (1960). Biosynthetic pathways of aromatic compounds.
active in proanthocyanidin accumulating cells in the seed coat.
Ann Rev. Plant Phys. 11, 55–80.
Plant Cell. 19, 2023–2038.
Niggeweg, R., Michel, A.J., and Martin, C. (2004). Engineering
Matsuda, F., Yonekura-Sakakibara, K., Niida, R., Kuromori, T.,
plants with increased levels of chlorogenic acid. Nature Biotech.
Shinozaki, K., and Saito, K. (2008). MS/MS spectral tag-based an-
22, 746–754.
notation of non-targeted profile of plant secondary metablites.
Plant J. 57, 555–577. Ober, D. (2005). Seeing double: gene duplication and diversification
in plant secondary metabolism. Trends Pl. Sci. 10, 445–449.
Matsui, K., Umemura, Y., and Ohme-Tkagi, M. (2008). AtMYBL2,
a protein with a single MYP domain, acts as a negative regulator Ogata, J., Kanno, Y., Itoh, Y., Tsugawa, H., and Suzuki, M. (2005).
of anthocyanin biosynthesis in Arabidopsis. Plant J. 55, 954–967. Plant biochemistry: anthocyanin biosynthesis in roses. Nature.
435, 757–758.
Matsuno, M., et al. (2009). Evolution of a novel phenolic pathway
for pollen development. Science. 326, 1688–1692. Ogawa, D., Nakajima, N., Seo, S., Mitsuhara, I., Kamada, H., and
Ohashi, Y. (2006). The phenylalanine pathway is the main route
Meissner, D., Albert, A., Böttcher, C., Strack, D., and Milkowski, C.
to salicylic acid biosynthesis in tobacco mosaic virus-infected to-
(2008). The role of UDP-glucose: hydroxycinnamic acid glucosyl-
transferase in phenylpropnaoid metabolism and the response to bacco leaves. Plant Biotechnol. 23, 395–398.
UV-Bradiation in Arabidopsis thaliana. Planta. 228, 663–674. Ono, E., Fukuchi-Mizutani, M., Nakamura, N., Fukui, Y., Yonekura-
Mellway, R.D., Tran, L.T., Prouse, M.B., Campbell, M.M., and Sakakibara, K., Yamaguchi, M., Nakayama, T., Tanaka, T.,
Constabel, P. (2009). The wound-, pathogen-, and ultraviolet Kusumi, T., and Tanaka, Y. (2006). Yellow flowers generated
B-responsive MYB134 gene encodes an R2R3 MYB transcription by expression of the aurone biosynthetic pathway. Proc. Natl
factor that regulates proanthocyanidin synthesis in poplar. Plant Acad. Sci. U S A. 103, 11075–11080.
Physiol. 150, 924–941. Orallo, F. (2008). Trans-resveratrol: a magical elixir for eternal
Messiano, D.B., Da Silva, T., Nascimento, I.R., and Lopez, L.M.X. youth? Curr. Med. Chem. 15, 1887–1898.
(2009). Biosynthesis of antimalarial lignans from Holostylis reni- Osmani, S.A., Bak, S., Imberty, A., Olson, C.E., and Møller, B.L.
formis. Phytochemistry. 70, 590–596. (2008). Catalytic key amino acids and UDP-sugar donor specificity
Meurer, B., Wiermann, R., and Strack, D. (1988). Phenylpropanoid of a plant glucuronosyltransferase, UGT94B1: molecular model-
pattern in Fagales pollen and their phylogenetic relevance. Phy- ling substantiated by site-specific mutagenesis and biochemical
tochemistry. 27, 823–828. analyses. Plant Physiol. 148, 1295–1308.
Milkowski, C., and Strack, D. (2004). Serine carboxypeptidase like Ossowski, S., Schwab, R., and Weigel, D. (2008). Gene silencing in
acyltransferases. Phytochemistry. 65, 517–524. plants using artificial microRNAs and other small RNAs. Plant J.
Milkowski, C., Baumert, A., and Strack, D. (2000). Identification of 53, 674–690.
four Arabidopsis genes encoding hydroxycinnamate glucosyl- Pervaiz, S. (2004). Chemotherapeutic potential of the chemopre-
transferases. FEBS Lett. 486, 183–184. ventive phytoalexin resveratrol. Drug. Resis. 7, 333–344.
Mir Derikvand, M., Berrio Sierra, J., Ruel, K., Pollet, B., Do, C.-T., Petersen, M., Abdullah, Y., Eberle, D., Gehlen, K., Hücherig, S.,
Thévenin, J., Buffard, D., Jouanin, L., and Lapierre, C. (2008). Re- Janiak, V., Kim, K.H., Sander, M., Wetzel, C., and Wolters, S.
Vogt d Phenylpropanoid Biosynthesis | 19

(2009). Evolution of rosmarinic acid biosynthesis. Phytochemis- Scalliet, G., Lionnet, C., Le Bechec, M., Dutron, L., Magnard, J.L.,
try. 70, 1663–1679. Baudino, S., Bergougnoux, V., Julien, F., Chambrier, P.,
Pichersky, E., and Gang, D. (2005). Genetics and biochemistry of sec- Vergne, P., Dumas, C., Cock, J.M., and Hugueney, P. (2006). Role
ondary metabolites in plants: an evolutionary perspective. of petal specific orcinol O-methyltransferase in the evolution of
Trends Pl. Sci. 4, 439–445. rose scent. Plant Physiol. 140, 18–29.
Pollard, M., Beisson, F., Li, Y., and Ohlrogge, J.B. (2008). Building Schijlen, E., de Vos, C.H., Jjonker, H., van den Broeck, H.,
lipid barriers: biosynthesis of cutin and suberin. Trends Pl. Sci. Molthoff, J., van Tune, A., Martens, S., and Bovy, A. (2006). Path-
13, 236–246. way engineering for healthy phytochemicalsleadingtotheproduc-
Pourcel, L., Routaboul, J.M., Kerhoas, L., Caboche, M., Lepiniec, L., tion of novel flavonoids in tomato fruit. Plant Biotech. J. 4, 433–444.
and Debeaujon, I. (2005). TRANSPARENT TESTA10 encodes a Schilmiller, A.L., Stout, J., Weng, J.K., Humphreys, J., Rueger, M.O.,
laccase-like enzyme involved in oxidative polymerisation of and Chapple, C.C.S. (2009). Mutations in the CINNAMATE
flavonoids in Arabidopsis seed coat. Plant Cell. 17, 2966–2980. 4-HYDROXYLASE gene impact metabolism, growth and
Rezai-Zadeh, K., et al. (2005). Green tea epigallocatechin-3-gallate development in Arabidopsis. Plant J. doi: 10.1111/j.1365–
(EGCG) modulates amyloid precursor protein cleavage and 313X.2009.03996.x.
reduces cerebral amyloids in Alzheimer transgenic mice. Schmid, J., and Amrhein, N. (1995). The molecular organisation of
J. Neurosci. 25, 8807–8814. the shikimate pathway in plants. Phytochemistry. 39, 739–747.
Richards, T.A., Dacks, J.B., Campbel, S.A., Blanchard, J.L., Schoch, G., Goepfert, S., Morant, M., Hen, A., Meyer, D.,
Foster, P.G., McLeod, R., and Roberts, C.W. (2006). Evolutionary Ullmann, P., and Werck-Reichart, D. (2001). CYP98A3 from Ara-
origins of the eukaryotic shikimate pathway: gene fusions, hor- bidopsis thaliana is a 3#-hydroxylase of phenolic esters, a missing
izontal gene transfer, and endosymbiotic replacement. Eukar- linking the phenylpropanoid pathway. J. Biol. Chem. 276,
yot. Cell. 5, 1517–1531. 36566–36574.
Rippert, P., Puyaubert, J., Grisollet, D., Derrier, L., and Matringe, M. Schoch, G., Morant, M., Abdulrazzak, N., Asnaghi, C., Goepfert, S.,
(2009). Tyrosine and phenylalanine are synthesized within the Petersen, M., Ullmann, P., and Werck-Reichhart, D. (2006). The
plastids in Arabidopsis. Plant Physiol. 149, 1251–1260. meta-hydroxylation step in the phenylpropanoid pathway:
Ritter, H., and Schulz, G.E. (2004). Structural basis for the entrance a new level of complexity in the pathway and its regulation. En-
into phenylpropanoid metabolism catalyzed by phenylalanine viron. Chem. Lett. 4, 127–136.
amoonia-lyase. Plant Cell. 16, 3426–3436. Schulze-Osthoff, K., and Wiermann, R. (1987). Phenols as inte-
Rohde, A., et al. (2004). Molecular phenotyping of the pal1 and pal2 grated compounds from Pinus pollen. J. Plant Physiol. 131, 5–15.
mutants in Arabidopsis thaliana reveals far-reaching consequen- Schuurink, R.C., Haring, M.A., and Clark, D.G. (2006). Regulation of
ces on phenylpropanoid, amino acid and carbohydrate metabo- volatile benzenoid biosynthesis in petunia flowers. Trends Pl. Sci.
lism. Plant Cell. 16, 2749–2771. 11, 20–25.
Rommens, C.M., Richael, C.M., Yan, H., Navarre, D.A., Ye, J., Sederoff, R.R., MacKAy, J.J., Ralph, J., and Hatfield, R.D. (1999). Un-
Krucker, M., and Swords, K. (2009). Engineered native pathways expected variation in lignin. Curr. Opin. Pl. Biol. 2, 145–152.
for high kaempferol and caffeoylquinate production in potato. Shadle, G.L., Wesley, S.V., Korth, K.L., Chen, F., Lamb, C., and
Plant Biotechnol. J. 6, 870–876. Dixon, C. (2003). Phenylpropanoid compounds and disease resis-
Ross, J.R., Nam, K.H., D’Auria, J.C., and Pichersky, E. (1999). S- tance in transgenic tobacco with altered expression of L-phenyl-
adenosyl-L-methionine: salicylic acid carboxyl methyltransferase, alanine ammonia-lyase. Phytochemistry. 64, 153–161.
an enzyme involved in floral scent production and plant defense, Shen, T., Wang, X.N., and Lou, H.X. (2009). Natural stilbenes: an
represents a class of plant methyltransferases. Arch. Biochem. overview. Nat. Prod. Rep. 26, 916–935.
Biophys. 367, 9–16. Siehl, E.L., Singh, B.K., and Conn, E.E. (1986). Tissue distribution and
Röther, D., Poppe, L., Morlock, G., Virguz, S., and Rétey, J. (2002). An subcellular localization of prephenate aminotransferase in
active site homology model of phenylalanine ammonia lyase leaves of Sorghum bicolour. Plant Physiol. 81, 711–713.
from Petroselinum crispum. Eur. J. Biochem. 269, 3065–3075. Sinlapadech, T., Stout, J., Ruegger, M.O., Deak, M., and Chapple, C.
Rubin, E.M. (2008). Genomics of cellulosic biofuels. Nature. 454, (2007). The hyperfluorescent trichome phenotype of the brt1
841–845. mutant of Arabidopsis is the result of a defect in a sinapic acid:
Samuels, A., Rensing, K., Douglas, C.J., Mansfield, S., UDPG glucosyltransferase. Plant J. 49, 655–668.
Dharmawardhana, D., and Ellis, B.E. (2002). Cellular machinery Soler, M., Serra, O., Molinas, M., Huguet, G., Fluch, S., and
of wood production: differentiation of secondary xylem in Pinus Figueras, M. (2007). A genomic approach to suberin biosynthesis
contorta var. latifolia. Planta. 216, 72–82. and cork differentiation. Plant Physiol. 144, 419–431.
Sasaki, K., Mito, K., Ohara, K., Yamamoto, H., and Yazaki, K. (2008). Stanjek, V., Piel, J., and Boland, W. (1999). Biosynthesis of furano-
Cloning and characterization of naringenin 8-prenyltransferase, coumarins: mevalonate- independent prenylation of umbellifer-
a flavonoid-specific prenyltransferase of Sophora flavescens. one in Apium graveolens (Apiaceae). Phytochemistry. 50,
Plant Physiol. 146, 1075–1084. 1141–1145.
Sawada, S., Suzuki, H., Ishimaida, F., Yamaguchi, M., Iwashita, T., Stehle, F., Brandt, W., Milkowski, C., and Strack, D. (2006).
Fukui, Y., Hemmi, H., Nishino, T., and Nakayama, T. (2005). Structure determinants and substrate recognition of serine
UDP-glucuronic acid:anthocynin glucuronosyltransferase from carboxypeptidase-like acyltransferases from plant secondary
red daisy (Bellis perennis) flowers. J. Biol. Chem. 280, 899–906. metabolism. FEBS Lett. 580, 6366–6374.
20 | Vogt d Phenylpropanoid Biosynthesis

Stehle, F., Brandt, W., Stubbs, M.T., Milkowski, C., and Strack, D. Gerbera hybrida catalyzes the conversion of (+)-catechin to cya-
(2009). Sinapoyltransferases in the light of molecular evolution. nidin and a novel procyanidin. FEBS Lett. 580, 1642–1648.
Phytochemistry. 70, 1652–1662. Werner, I., Bacher, A., and Eisenreich, W. (1997). Retrobiosynthetic
Stevens, J., and Page, J. (2004). Xanthohumol and related prenyl- NMR studies with 13C-labelled glucose. J. Biol. Chem. 272,
flavonoids from hops and beer: to your good health. Phytochem- 25474–25482.
istry. 65, 1317–1330. Werner, R.A., Rossmann, A., Schwarz, C., Bacher, A., Schmidt, H.L.,
Stewart, J.J., Akiyama, T., Chapple, C.C.S., Ralph, J., and and Eisenreich, W. (2004). Biosynthesis of gallic acid in Rhus
Mansfield, S.D. (2009). The effects on lignin structure on over- typhina: discrimination between alternative pathways from nat-
expression of ferulate 5-hydroxylase in hybrid poplar. Plant ural oxygen isotope labelling. Phytochemistry. 65, 2809–2813.
Physiol. 150, 621–635. Wildermuth, M.C., Dewdeny, J., Wu, G., and Ausubel, F.M. (2001).
Stracke, R., Ishihara, H., Huep, G., Barsch, A., Mehrtens, F., Isochorismate synthase is required to synthesize salicylic acid for
Niehaus, K., and Weisshaar, B. (2007). Differential regulation plant defence. Nature. 414, 562–565.
of closely related R2R3-MYP transcription factors controls flavo- Xie, Z., Kapteyn, J., and Gang, D.R. (2008). A systems biology inves-
nol accumulation in different parts of the Arabidopsis thaliana tigation of the MEP/terpenoid and shikimate/phenylpropanoid
seedling. Plant J. 50, 660–677. pathways points to multiple levels of metabolic control in sweet
Strawn, M.A., Marr, S.K., Inoue, K., Inada, N., Zubieta, C., and basil glandular trichomes. Plant J. 54, 349–361.
Wildermuth, M.C. (2007). Arabidopsis isochorismate synthase Yamada, T., Matsuda, F., Kasai, K., Fukuoka, S., Kitamura, K.,
functional in pathogen-induced salicylate biosynthesis exhibits Tozawa, Y., Miyagawa, H., and Wakasa, K. (2008). Mutation
properties consistent with a role in diverse stress responses. of a rice gene encoding a phenylalanine biosynthetic enzyme
J. Biol. Chem. 282, 5919–5933. results in accumulation of phenylalanine and tryptophan. Plant
Terrier, N., Torregrosa, L., Ageorges, A., Vialet, S., Verries, C., J. 20, 1316–1329.
Cheynier, V., and Romieu, C. (2009). Ectopic expression of Yonekura-Sakakibara, K., Toghe, T., Matsuda, F., Nakabayashi, R.,
VvMybPA2 promotes proanthocyanidin biosynthesis in grape- Takayama, H., Niida, R., Watanabe-Takahashi, A., Inoue, E.,
vine and suggests additional targets in the pathway. Plant Phys- and Saito, K. (2008). Comprehensive flavonol profiling and tran-
iol. 149, 1028–1041. scriptome analysis leading to decoding gene-metabolite correla-
Tian, L., Pang, Y., and Dixon, R. (2008). Biosynthesis and genetic en- tions in Arabidopsis. Plant Cell. 20, 2160–2176.
gineering of proanthocyanidins and (iso)flavonoids. Phytochem. Yonekura-Sakakibara, K., Toghe, T., Nilda, R., and Saito, K. (2007).
Rev. 7, 445–465. Identification of a flavonol 7-O-rhamnosyltransferase gene de-
Tohge, T., et al. (2005). Functional genomics by integrated trancrip- termining flavonoid pattern in Arabidopsis by transcriptome
tome and metabolome of Arabidopsis plants overexpressing coexpression analysis and reverse genetics. J. Biol. Chem. 282,
a MYB-transcription factor. Plant J. 42, 218–235. 14932–14941.
Van Moerkercke, A., Schauvinold, I., Pichersky, E., Haring, M.A., and Yoshida, K., Mori, M., and Kondo, T. (2009). Blue flower colour de-
Schuurink, R. (2009). A plant thiolase involved in benzoic acid velopment by anthocyanins: from chemical structure to cell phys-
biosynthesis and volatile benzenoid production in plants. Plant iology. Nat. Prod. Rep. 26, 884–915.
J. 60, 292–302.
Zeller, G., Henz, S.R., Widmer, C.K., Sachsenberg, T., Rätsch, G.,
Vanholme, R., Morreel, K., Ralph, J., and Boerjan, W. (2008). Lignin Weigel, D., and Laubinger, S. (2009). Stress-induced changes in
engineering. Curr. Opin. Plant. Biol. 11, 278–285. the Arabidopsis thaliana transcriptome analyzed using whole-
Veitch, N. (2009). Isoflavonoids of the Leguminosae. Nat. Prod. Rep. genome tilling arrays. Plant J. 58, 1068–1082.
26, 776–802. Zhao, J., and Dixon, R. (2009). Mate transporter facilitate vacuolar
Vlot, A.C., Klessig, D.F., and Park, S.W. (2008). Systemic aquired re- uptake of epicatechin 3#-O-glucoside for proanthocyanidin bio-
sistance, the elusive signal(s). Curr. Opin. Pl. Biol. 11, 436–442. synthesis in Medicago truncatula and Arabidopsis. Plant Cell. 21,
Wagner, A., Ralph, J., Akiyama, T., Flint, H., Phillips, L., Torr, K., 2323–2340.
Nanayakkara, B., and Te Kiri, L. (2007). Exploring lignification Zhao, N., Ferrer, J.-L., Ross, J., Guan, J., Yang, Y., Pichersky, E.,
in conifers by silencing hydroxycinnamoyl–CoA: shikimate Noel, J.P., and Chen, F. (2008). Structural, biochemical, and phy-
hydroxycinnamoyltransferase in Pinus radiata. Proc. Natl Acad. logentical analyses suggest that indole-3-acetic acid methyl-
Sci. U S A. 104, 11856–11861. transferase is an evolutionary ancient member of the SABATH
Walters, D.R. (2003). Polyamines and plant disease. Phytochemistry. family. Plant Physiol. 146, 455–467.
64, 97–107. Zhao, N., Guan, J., Forouhar, F., Tschaplinski, T.J., Cheng, Z.-M.,
Watts, K.T., Mijts, B.N., Lee, B.C., Manning, A.J., and Schmidt- Tong, L., and Chen, F. (2009). Two poplar methyl salicylate
Dannert, J. (2006). Discovery of a substrate selectivity switch in esterases display comparable biochemical properties but diver-
tyrosine ammonia lyase, a member of the aromatic amino acid gent expression patterns. Phytochemistry. 70, 32–38.
lyase family. Chem. Biol. 13, 1317–1326. Zhong, W. (2009). Nanomaterials in fluorescence-based biosensing.
Weisshaar, B., and Jenkins, G.I. (1998). Phenylpropanoid metabo- Anal. Bioanal. Chem. 394, 47–56.
lism and its regulation. Curr. Opin. Pl. Biol. 1, 251–257. Zhu, H.F., Fitzsimmons, K., Khandelwal, A., and Kranz, R.G. (2009).
Wellmann, F., Griesser, M., Schwab, W., Martens, S., Eisenreich, W., CPC, a single repeat R3 MYB, is a negative regulator of anthocy-
Matern, U., and Lukacin, R. (2006). Anthocyanidin synthase from anin biosynthesis in Arabidopsis. Mol. Plant. 2, 790–802.

You might also like