ArticlePDF Available

Vitamin K1 (phylloquinone): function, enzymes and genes

Authors:

Figures

Content may be subject to copyright.
Provided for non-commercial research and educational use only.
Not for reproduction, distribution or commercial use.
This chapter was originally published in the book Advances in Botanical Research,
Vol. 59, published by Elsevier, and the attached copy is provided by Elsevier for the
author's benefit and for the benefit of the author's institution, for non-commercial
research and educational use including without limitation use in instruction at your
institution, sending it to specific colleagues who know you, and providing a copy to
your institution’s administrator.
All other uses, reproduction and distribution, including without limitation commercial
reprints, selling or licensing copies or access, or posting on open internet sites, your
personal or institution’s website or repository, are prohibited. For exceptions,
permission may be sought for such use through Elsevier's permissions site at:
http://www.elsevier.com/locate/permissionusematerial
From: Chloë van Oostende, Joshua R. Widhalm, Fabienne Furt, Anne-Lise Ducluzeau
and Gilles J. Basset, Vitamin K1 (Phylloquinone): Function, Enzymes and Genes.
In Fabrice Rébeillé and Roland Douce, editors:
Advances in Botanical Research, Vol. 59,
Amsterdam, The Netherlands, 2011, pp. 229-261.
ISBN: 978-0-12-385853-5
© Copyright 2011 Elsevier Ltd.
Academic Press.
Vitamin K
1
(Phylloquinone): Function, Enzymes and Genes
CHLOE
¨VAN OOSTENDE, JOSHUA R. WIDHALM,
FABIENNE FURT, ANNE-LISE DUCLUZEAU
AND GILLES J. BASSET
1
Center for Plant Science Innovation, University of Nebraska-Lincoln,
Lincoln, Nebraska, USA
I. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
II. Structure and Chemistry of Vitamin K . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
III. Biochemical Roles of Vitamin K . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
A. Vertebrates... . . . . . . ...... . . . . . . ....... . . . . . . ...... . . . . . ........ . . . . . ...... . . . . . 233
B. Plants and Cyanobacteria .. .. .. .. ... .. .. .. .... .. .. .. ... .. .. .. .... .. .. .. ... .. 236
IV. Detection and Distribution of Phylloquinone in Plants . . . . . . . . . . . . . . . . . . 238
A. Detection....... . . . . . . ...... . . . . . . ....... . . . . . . ...... . . . . . ........ . . . . . ...... . . . 238
B. Tissular Distribution . .. . .. .. .. .. .. . .. .. .. .. .. . .. .. .. .. .. .. . .. .. .. .. .. . .. .. .. . 238
C. Subcellular Distribution . .. ..... .. .. .... .. .. .... .. .. ... .. .. .... .. .. .... .. .. .. 238
V. Phylloquinone Biosynthesis in Plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
A. Early Work . .. ..... .. .. .. ... .. .. .... .. .. ...... .. .. .... .. .. ..... .. .. ..... .. .. ... 240
B. Isochorismate Synthase/PHYLLO (Reactions 1–4) ....... . . . . . ....... . . 242
C. OSB-CoA Ligase (Reaction 5) . . . . . . ...... . . . . . . ....... . . . . . . ...... . . . . . ... 244
D. DHNA-CoA Synthase/DHNA-CoA Thioesterase (Reactions 6/7)... 244
E. DHNA Phytyl Transferase (Reaction 8)... . . . . . ........ . . . . . ...... . . . . . . . 246
F. Demethylphylloquinone Methyltransferase (Reaction 9) .. . . . . . ...... . 246
G. Mutant Phenotype.... . . . . . ....... . . . . . . ....... . . . . . ....... . . . . . ....... . . . . . . . 247
H. Subcellular Localization of Phylloquinone Biosynthetic Enzymes ... 247
VI. Evolution of Naphthoquinone Biosynthesis in Photosynthetic
Eukaryotes ................................................................ 250
VII. Phylloquinone Turnover . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
1
Corresponding author: E-mail: gbasset2@unl.edu
Advances in Botanical Research, Vol. 59 0065-2296/11 $35.00
Copyright 2011, Elsevier Ltd. All rights reserved. DOI: 10.1016/B978-0-12-385853-5.00001-5
Author's personal copy
VIII. Engineering of Phylloquinone in Plants. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
IX. Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
ABSTRACT
Phylloquinone (2-methyl-3-phytyl-1,4-naphthoquinone) is a conjugated isoprenoid
that serves as a cardinal redox cofactor in plants and some cyanobacteria. In humans
and other mammals, it is required as a vitamin (vitamin K
1
) for blood coagulation
and bone metabolism. Until recently, the biosynthesis of phylloquinone in plants was
considered identical to that of menaquinone (vitamin K
2
) in facultative anaerobic
bacteria. It resulted that most of the plant research on phylloquinone focused histori-
cally on the study of its function, while very little was done on its metabolism per se.
There is today, tough, compelling evidence that plants have evolved an unprecedented
metabolic architecture to synthesize phylloquinone, including extraordinary events of
gene fusion, highly divergent enzymes and a separated compartmentalization in
chloroplasts and peroxisomes. Phylogenetic reconstructions also demonstrate that
the plant genes involved in the formation of phylloquinone display a high degree of
evolutionary chimerism owing to multiple events of horizontal gene transfer and gene
losses. Plant phylloquinone biosynthesis is also connected via shared intermediates to
the metabolism of salicylate, tocopherols, chlorophylls, and in some species to
anthraquinones.
I. INTRODUCTION
The discovery of vitamin K arose from the observation in the late 1920s and
early 1930s that chicks reared on a reconstituted ‘sterol-free’ diet developed a
hemorrhagic disease characterized by a severe impairment in blood coagulation
(Almquist and Stokstad, 1935; Dam, 1929, 1935; Dam and Schønheyder, 1934;
Holst and Halbrook, 1933; McFarlane et al., 1931; Schønheyder, 1935). The
lack of sterols or fat in the diet asthe cause of the disease was quickly ruled out,
as haemorrhages still appeared in chicks receiving a daily supplement of choles-
terol and oil from cod-liver or flax seeds. Nor did it appear that the haemor-
rhages were caused by a lack of any of the known vitamins.Feeding experiments
with supplements obtained from various fractionation procedures showed,
however, that the protecting factor resembled vitamin E, being thermostable,
fat-soluble and non-saponifiable. (Almquist and Stokstad, 1935; Dam, 1935;
Dam and Schønheyder, 1934). The disease could be prevented or cured by
supplementing the chicks’ diet with various plant or animal products such as
fresh cabbage, dried alfalfa, tomatoes, hemp seeds, putrefied fish meal—but not
fresh—or hog liver fat (Almquist and Stokstad, 1935; Dam, 1935; Dam and
Schønheyder, 1934). Almquist and Stokstad (1935)at the College of Agriculture
of the University of California-Berkeley established early on that the green parts
230 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
of the plant ingredients were the sources of the antihemorrhagic factor, and that
it was distinct from chlorophyll and xanthophyll. They also understood that, in
the case of putrefied fish meal, the antihemorrhagic factor originated from the
development of microorganisms (Almquist and Stokstad, 1935). Henrik Dam
at the Biochemical Institute of the University of Copenhagen recognized this
antihemorrhagic factor as a vitamin; he named it ‘vitamin K’, for K was the first
letter in the alphabet that had not been used to designate other vitamins, and
coincidently happened to correspond to the first letter in the word ‘koagulation’
as spelled in Scandinavian (Dam, 1935). A few years later, Edward A. Doisy’s
group at the Laboratory of Biological Chemistry from St. Louis University
School of Medicine purified vitamin K
1
(phylloquinone) from alfalfa, deter-
mined its structure and achieved its chemical synthesis (Binkley et al.,1939;
MacCorquodale et al.,1939a,b;McKeeet al.,1939). Shortly after, vitamin K
2
(menaquinone) was isolated from putrefied fish meal and characterized (Doisy
et al.,1941). The 1943 Nobel Prize in Physiology or Medicine was co-awarded to
Henrik Dam ‘for his discovery of vitamin K’ and to Edward A. Doisy ‘for his
discovery of the chemical nature of vitamin K’. The award did not acknowledge
the pioneering work of Almquist and Stokstad, who co-discovered vitamin K
independently from Dam, or that of Schønheyder, who demonstrated that
vitamin K deficiency impaired blood coagulation. As for plants, one had to
wait until the mid-1980s to find out that phylloquinone participates in the
photosynthetic electron transfer chain and until the past couple of years to
discover that it doubles as an electron acceptor linked to the formation of
disulfide bridges in proteins.
Some readers may also be surprised to learn that until the middle of this
decade—and despite the cardinal role played by phylloquinone in photosyn-
thesis and human nutrition—not much was known about the biosynthesis of
this vitamin. As we will see later, plant biochemists were among the leaders in
the early studies of vitamin K biosynthesis. Unfortunately, as emerged a
general assumption that the biosynthesis of phylloquinone in photosynthetic
organisms was identical to that of menaquinone in facultative anaerobic
bacteria, research on the metabolism of vitamin K in plants virtually ceased
for decades. If it is indeed correct to view the individual steps of phylloqui-
none and menaquinone biosynthesis as similar, the most recent investiga-
tions showed that plants evolved an unprecedented architecture to synthesize
phylloquinone, including extraordinary events of gene fusion and horizontal
gene transfer, split of the pathway between plastids and peroxisomes and
multiple metabolic branch points that link the biosynthesis of phylloquinone
to that of salicylate, tocopherols and chlorophylls. The study of phylloqui-
none biosynthesis in cyanobacteria even led a couple of years ago to the
discovery of a ‘missing’ enzyme of the vitamin K biosynthetic pathway.
VITAMIN K
1
(PHYLLOQUINONE) 231
Author's personal copy
This review aims to summarize the current knowledge concerning
the function of vitamin K in vertebrates and oxygenic photosynthetic organ-
isms and its metabolism—emphasizing the most recent advances in under-
standing the phylloquinone biosynthetic pathway and its evolution
in plants—and to point out areas that are still obscure. We refer the reader
to the reviews of Sakuragi and Bryant (2006), Fromme and Grotjohann
(2006) and van der Est (2006) for a detailed coverage of the biosynthesis
and function of phylloquinone in cyanobacteria. We will nonetheless discuss
on occasion recent works concerning the biosynthesis of isoprenoid naphtho-
quinones in cyanobacteria and facultative anaerobic bacteria, as some of the
findings in these organisms could represent paradigms for plants.
II. STRUCTURE AND CHEMISTRY OF VITAMIN K
The term vitamin K encompasses a class of fat-soluble compounds formed
from a naphthoquinone ring attached to a poly-isoprenyl side chain of
variable length and saturation (Fig. 1A). Its main natural forms are vitamin
K
1
(phylloquinone; 2-methyl-3-phytyl-1,4-naphthoquinone) that is found in
plants (Oostende et al., 2008), green algae (Lefebvre-Legendre et al., 2007)
and certain cyanobacteria (Collins and Jones, 1981) and vitamin K
2
(mena-
quinones; 2-methyl-3-(all-trans-polyprenyl)-1,4-naphthoquinone) that is
found in most groups of archaea and bacteria (Collins and Jones, 1981),
the cyanobacterium Gloeobacter violaceus (Mimuro et al., 2005), red algae
(Yoshida et al., 2003) and diatoms (Ikeda et al., 2008). Vitamin K-synthesiz-
ing organisms appear to contain either phylloquinone or menaquinones but
not both. Phylloquinone has a partially unsaturated side chain formed of one
isopentenyl followed by three isopentyl units, while menaquinones have a
fully unsaturated side chain composed of 2–13 isopentenyl units (Fig. 1A).
Menaquinones are often designated as menaquinone-n(MK-n), where n
refers to the number of isopentenyl units in the side chain.
The naphthoquinone ring of vitamin K can exist at different levels of
oxidation, varying from epoxide (the most oxidized) to quinol (the most
reduced) through the intermediate quinone and semi-quinone (Fig. 1B).
The epoxide form is the product of an enzymatic reaction that so far
has been identified only in animal cells. Most of plant and cyanobacterial
phylloquinone is in the quinone form (Oostende et al., 2008; Widhalm et al.,
2009).
While in facultative anaerobic bacteria, menaquinones are often found
lacking the methyl group at position 2 of the naphthoquinone ring—for
instance, up to 90% of the menaquinone pool in aerobically grown
232 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
Escherichia coli is demethylated (Unden, 1988)—in plants and cyanobac-
teria, phylloquinone is virtually all in the methylated form [there is a single
report of the presence of trace levels of demethylphylloquinone in spinach
chloroplasts (McKenna et al., 1964)].
III. BIOCHEMICAL ROLES OF VITAMIN K
A. VERTEBRATES
The major known function of vitamin K in vertebrates is that of a cofactor
for the -carboxylation of specific glutamate residues, thus conferring strong
chelating properties to certain proteins whose activity depends on calcium
binding. Among such -carboxylglutamate (Gla)-containing proteins are
Fig. 1. (A) Structures of phylloquinone (vitamin K
1
) and menaquinones
(vitamin K
2
). The phytyl side chain of phylloquinone contains one isopentenyl unit
and three isopentyl units, while that of menaquinones are made exclusively of iso-
pentyl units. (B) Interconversion of the different redox forms of the naphthoquinone
ring. R: poly-isoprenyl moiety.
VITAMIN K
1
(PHYLLOQUINONE) 233
Author's personal copy
blood coagulation factors (prothrombin, factors VII, IX and X), proteins
that participate in bone metabolism (osteocalcin, Matrix Gla Protein) and
cell signalling (Gas6).
In order to fulfil its role as cofactor for the -carboxylase, vitamin K must
be in the quinol form. As a by-product of the -carboxylation, the bireduced
naphthalenoid ring of vitamin K is converted to the fully oxidized epoxide
form (Fig. 2). It is salvaged by an integral enzyme complex, named Vitamin
K epoxide reductase (VKOR), whose catalytic subunit (VKORC1; EC
1.1.4.1) reduces the epoxide back to quinone and the quinone back to quinol
(Chu et al., 2006;Fig. 2). This enzyme is the target of the vitamin K antago-
nist warfarin, used as an anticoagulant drug and rodenticide. Besides its role
as a cofactor, studies on mammalian cell cultures indicate that vitamin K acts
as a transcriptional regulator (Ichikawa et al., 2006) and as an antioxidant
(Li et al., 2003).
The lack of vitamin K results in non-functional Gla-containing proteins,
which, in turn, can lead to impaired blood clotting and bone mineralization.
Severe vitamin K deficiency, which can lead to easy bruising and bleeding, is,
however, rare in healthy adults because vitamin K is widespread in foods,
Fig. 2. Scheme of the vitamin K-dependent -carboxylation of glutamyl residues
in vertebrates. Vitamin K quinol is converted to an oxygenated intermediate that
abstracts a proton from the -carbon of the glutamyl residue of the carboxylase
substrate, followed by the addition of carbon dioxide. The concomitant oxidation
of vitamin K quinol into vitamin K epoxide, and its subsequent salvaging by the
enzyme complex vitamin K epoxide reductase (VKOR), is called the vitamin K cycle.
234 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
and the gut flora produces basal levels of menaquinones (Suttie, 1995). Only
individuals having chronicle hepatic and pancreatic disorders (Savage and
Lindenbaum, 1983), those receiving long-term antibiotic (Savage and
Lindenbaum, 1983; Shevchuk and Conly, 1990) or vitamin K antagonist
treatments (Bach et al., 1996), appear to be at risk of acute vitamin K
deficiency. Newborns, whose intestinal flora is not yet established, stand
apart and are naturally exposed to an increased risk of vitamin K deficien-
cy—often leading to dramatic haemorrhage of the central nervous system.
The risk is actually higher for infants who are exclusively breast-fed because
the human milk contains only traces of this vitamin (American Academy of
Pediatrics, 2003). It is therefore routine—and often mandatory—in many
countries to administer intramuscular or oral vitamin K at birth as a pro-
phylactic measure (American Academy of Pediatrics, 2003). The incidence of
unexpected bleeding during the first week of life in previously healthy neo-
nates ranges from 250 to 1700 per 100,000 births, and these numbers rise
to 4400–7200 per 100,000 births in infants 2–12 weeks of age who have
received no or inadequate vitamin K prophylaxis (American Academy of
Pediatrics, 2003).
The adequate intake values for vitamin K in the United States are current-
ly set at 120 g/day for men and 90 g/day for women (Food and Nutrition
Board, 2001). Specific levels have not yet been established in the European
Union, but the Committee on Medical Aspects of Food and Nutrition Policy
in the United Kingdom considered that an intake of 1 g/kg of body weight/
day is likely adequate for the proper carboxylation of blood coagulation
factors. In a typical western diet, phylloquinone is the main contributor of
vitamin K intake; about half of it comes from green leafy vegetables, fol-
lowed by soybean, olive, canola and cottonseed oils (Booth and Suttie, 1998).
American and British studies reported average values for dietary vitamin K
intake ranging from 60 to 70 g/day and suggested that one-half of the
populations investigated had vitamin K intakes below the present guidelines
(Vermeer et al., 2004). The impact on bone health of such suboptimal intakes
is currently debated. There is evidence that the level of circulating under-
carboxylated osteocalcin increases after menopause, and that it correlates
with an increased risk of hip fracture (Szulc et al., 1996). Some epidemiologi-
cal studies also reported that individuals with the highest vitamin K intakes
have lower risk of hip fracture than those with the lowest intakes (Booth
et al., 2000; Feskanich et al., 1999), but others found no correlations
(McLean et al., 2006; Rejnmark et al., 2006). None of these studies could
establish a relationship between vitamin K intake and bone mineral density.
Clinical trials indicated a possible increase in bone mineral density and bone
strength in postmenopausal women receiving vitamin K supplementation,
VITAMIN K
1
(PHYLLOQUINONE) 235
Author's personal copy
but the doses used were several orders of magnitude higher than those
commonly found in the diet (Iwamoto et al., 2001; Knapen et al., 2007).
B. PLANTS AND CYANOBACTERIA
Until recently, the sole firmly established function of phylloquinone in pho-
tosynthetic organisms was that of a light-dependent electron carrier—the A
1
acceptor—in photosystem I (Brettel et al., 1987; Petersen et al., 1987;
Sigfridsson et al., 1995). The process is a one-electron transfer, that is, a
quinone/semi-quinone turnover—from chlorophyll ato the iron–sulphur
cluster of ferredoxin reductase (Boudreaux et al., 2001; Sigfridsson et al.,
1995;Fig. 3). There are two molecules of phylloquinone, called Q
K
A and
Q
K
B, that are bound to the PsaA and PsaB subunits, respectively, at the
stromal side of each photosystem I monomer (Ben-Shem et al., 2003;
Boudreaux et al., 2001; Jordan et al., 2001). Both molecules are active in
electron transport, but the transfer rate through Q
K
B appears to be 50 times
higher than that through Q
K
A(Guergova-Kuras et al., 2001). The reasons
for such a difference in kinetics between the two branches are not yet fully
understood (Fromme and Grotjohann, 2006).
Fig. 3. Scheme of the electron transfer in photosystem I and approximate mid-
point potentials of the cognate electron carriers in plants and cyanobacteria. Phyllo-
quinone is located at the A
1
site of the PsaA and PsaB subunits of photosystem I to
serve as a one-electron carrier from chlorophyll aA
0
) to the Fe-S cluster (F
X
,F
A
/F
B
).
P700, photosystem I reaction center; P700*, excited photosystem I reaction center.
236 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
Reports that at least half of phylloquinone is not bound to photosystem I
(Gross et al., 2006; Lohmann et al., 2006), together with the detection of the
quinol form of phylloquinone in multiple plant species (Oostende et al., 2008)
and the cyanobacterium Synechocystis (Widhalm et al., 2009), have recently
indicated that in photosynthetic organisms, phylloquinone is probably
involved in redox reactions distinct from that of the one-electron transfer
in photosystem I. Further evidence for such an additional role came with the
discovery of phylogenetic relationships between mammalian VKOR and
predicted oxidoreductases in photosynthetic organisms. Indeed, although
there is no genomic or biochemical indication that the -carboxylation of
glutamic acid residues and the resulting generation of vitamin K epoxide
occur outside of the metazoan lineage, homology searches detect cyanobac-
terial and plant proteins that are similar to mammalian VKORC1. Remark-
ably, these VKORC1 homologues display a C-terminal fusion with a soluble
thioredoxin-like domain having the hallmarks of a protein disulfide isomer-
ase (Furt et al., 2010; Goodstadt and Ponting, 2004; Li et al., 2010; Singh
et al., 2008). In vitro assays demonstrated that the Synechococcus-fused
enzyme could couple the formation of disulfide bonds in an artificial protein
substrate to the reduction of phylloquinone (Li et al., 2010). Similarly, the
Arabidopsis orthologue—the At4g35760 gene product, which is localized in
plastids—was shown to catalyze the conversion of conjugated naphthoqui-
none species into their quinol forms using either dithiotreitol or its protein
disulfide isomerase moiety as electron donors. However, unlike mammalian
VKORC1, the plant and cyanobacterial enzymes lack phylloquinone epoxide
reductase activity and are resistant to warfarin (Furt et al., 2010). The
Arabidopsis enzyme also appears to be inactive on conjugated benzoqui-
nones such as plastoquinone and ubiquinone. Such a substrate stringency
might be unique to plants, for there is evidence that the recombinant Syne-
chococcus enzyme binds ubiquinone (Li et al., 2010), so as does the E. coli
DsbB protein (quinone oxidoreductase), which features similarities in se-
quence, structure and mode of action with VKORC1, and in fact uses
menaquinone or ubiquinone as oxidant molecules (Inaba et al., 2004;
Takahashi et al., 2004).
Along this line, it is noteworthy that in Synechocystis sp. PCC 6803, the
phenotype of the VKORC1-like (slr0565) knockout does not parallel that of
mutant strains lacking phylloquinone. Indeed, deletion of slr0565 causes
lethality or severe growth retardation depending on the presence or absence
of glucose in the culture medium, respectively (Singh et al., 2008), whereas in
similar conditions, Synechocystis mutants blocked in phylloquinone biosyn-
thesisdisplay either no or moderate growth defects (see Section V.G). It is
therefore conceivable that in phylloquinone-deficient cyanobacteria, the
VITAMIN K
1
(PHYLLOQUINONE) 237
Author's personal copy
VKORC1-like enzyme functions using an alternate substrate, possibly plas-
toquinone. In plants, however, the lack of phylloquinone causes complete
loss of photoautotrophy (see Section V.G), and failure to obtain Arabidopsis
transgenics, whose VKORC1-like expression is deregulated, indicating that
this enzyme fulfils a core and vital role (Furt et al., 2010).
IV. DETECTION AND DISTRIBUTION OF
PHYLLOQUINONE IN PLANTS
A. DETECTION
Early methods for the determination of vitamin K in food or biological
extracts relied on a chick bioassay. Analyses were tedious, entailing large
extraction volumes—especially in samples of animal origin due to their
extremely low vitamin K content—and provided only semiquantitative
data (Dam and Schønheyder, 1936). Subsequent quantitative methods were
developed using thin-layer chromatography, gas chromatography and high-
performance liquid chromatography (HPLC); the latter coupled either to
fluorometric or to electrochemical detection (Davidson and Sadowski, 1997;
McCarthy et al., 1997). HPLC methods based on the reduction of the
naphthoquinone ring to its fluorescent quinol form prior to its detection by
fluorometry have proven to combine high sensitivity and selectivity and are
today the preferred applications for the routine quantification of vitamin K
in complex extracts (Booth and Sadowski, 1997; Davidson and Sadowski,
1997). On an additional technical note, let us mention that in green plant
tissues, phylloquinone is often sufficiently abundant to be detected and
quantified using HPLC–spectrophotometry (Fraser et al., 2000).
B. TISSULAR DISTRIBUTION
The level of phylloquinone varies greatly between different plant species and
tissues (Table I). Leaves usually have the highest levels, while most fruits,
tubers and seeds contain several-fold less. It is noteworthy that staple crops
(e.g. grains and tubercles) are among the poorest plant sources of
phylloquinone.
C. SUBCELLULAR DISTRIBUTION
At the subcellular level, plastids account for most if not all of the phylloquinone
content of plant tissues (Lohmann et al.,2006;Oostendeet al.,2008). Subplas-
tidial fractionation experiments demonstrated that about a third of
238 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
phylloquinone in Arabidopsis chloroplasts is deposited in plastoglobules, and
therefore, a significant amount of phylloquinone is not bound to photosystem I
(Lohmann et al.,2006). A similar conclusion was inferred from the observation
that Arabidopsis mutants containing less than a quarter of wild-type levels of
phylloquinone retained most of their photosystem I activity (Gross et al.,2006).
The enrichment of naphthoquinone oxidoreductase activities in plasma
membrane preparations of corn roots (Lu
¨thje et al., 1998) and soybean
hypocotyls (Bridge et al., 2000; Schopfer et al., 2008) suggests that small
pools of phylloquinone may occur outside plastids. One of these studies
reported the direct detection of phylloquinone in the plasma membrane
(Lu
¨thje et al., 1998), but the possibility of proplastid breakage (e.g. using
galactolipids as markers) was not investigated.
TABLE I
Phylloquinone Content of Some Plant Species and Plant Food-Products
Phylloquinone (g/100 g)
A. thaliana (green leaf) 365
(a)
Brassica oleracea (canola oil) 127
(b)
Brassica oleracea (collard greens) 440
(b)
Brassica oleracea (broccoli) 180
(b)
Brassica oleracea (brussel sprouts) 177
(b)
Brassica oleracea (cauliflower) 20
(b)
Cicer arietinum (chickpeas) 9
(c)
Daucus carota (tuber) 2.7
(a)
Lactuca sativa (green leaf) 126
(c)
Lactuca sativa (‘iceberg’ lettuce) 35
(b)
Manihot esculenta (cassava) 1.9
(c)
Olea europaea (olive oil) 55
Oryza sativa (grain) 0.1
(c)
Oryza sativa (green leaf) 662
(a)
Phaseolus vulgaris (dry bean) 5.6
(c)
Phaseolus vulgaris (green beans) 33
(b)
Solanum. lycopersicon (green leaf) 1217
(a)
Solanum. lycopersicon (green fruit) 19
(a)
Solanum. lycopersicon (red ripe fruit) 8
(a)
Solanum tuberosum (tuber) 1.3
(a)
Glycine max (soybean oil) 193
(b)
Glycine max (‘Edamame’ seed) 31
(c)
Triticum spp. (whole grain flour) 1.9
(c)
Vicia faba (fava bean) 9
(c)
Zea mays (grain) 0.3
(c)
Zea mays (green leaf) 1514
(a)
Zea mays (oil) 3
(b)
Data are compiled from Oostende et al. (2008)
(a)
;Booth and Suttie (1998)
(b)
; USDA National
Nutrient Database for Standard Reference (http://www.nal.usda.gov/fnic/foodcomp/search/)
(c)
.
Staple crops are shown in bold.
VITAMIN K
1
(PHYLLOQUINONE) 239
Author's personal copy
V. PHYLLOQUINONE BIOSYNTHESIS IN PLANTS
In essence, the phylloquinone biosynthetic pathway of plants consists of two
separated metabolic branches: one for the naphthoquinone ring and the
other for the phytyl moiety, which is also used for the biosyntheses of
tocopherols and chlorophylls. We focus hereafter on the enzymatic steps
leading to the formation of the naphthoquinone ring, as the biosynthesis of
the phytyl-diphosphate precursor from the methylerythritol-phosphate path-
way in plastids is not specific to phylloquinone and has been previously
covered (see, for instance, Lichtenthaler, 1999; Rohmer, 2003).
The biosynthesis of the naphthoquinone ring entails seven enzymatic steps.
The immediate precursor chorismate is first converted into isochorismate, to
which a succinyl side chain is added at the C2 position (Fig. 4). After
elimination of pyruvate and aromatization of the cyclohexadiene ring, the
succinyl chain is activated by ligation with CoA and then cyclized, yielding
1,4-dihydroxynaphthoyl-CoA (DHNA-CoA). The CoA moiety is then
removed, DHNA is conjugated to its phytyl partner and then methylated.
An alternative pathway, in which the naphthoquinone backbone originates
from a chorismate–inosine conjugate termed futalosine, was recently
described in some species of Deinococcus-Thermus, Actinobacteria and
E-Proteobacteria (Hiratsuka et al., 2008). There is currently no genomic or
biochemical evidence that this biosynthetic route occurs in phylloquinone-
synthesizing eukaryotes.
A. EARLY WORK
The basic architecture of isoprenyl naphthoquinone biosynthesis was estab-
lished in the 1970s and 1980s simultaneously in plants and facultative anaer-
obic bacteria using radiolabelling experiments. It quickly emerged from these
studies that the individual steps of the phylloquinone and menaquinone
biosynthetic pathways were virtually identical. In plants, shikimate was
identified as the precursor of the naphthoquinone moiety via the formation
of o-succinylbenzoate (OSB), OSB-CoA and DHNA (Dansette and Azerad,
1970; Heide et al., 1982; Hutson and Threlfall, 1980; Thomas and Threlfall,
1974), while the ring prenylation and methylation steps were found to use
phytyl-diphosphate and s-adenosylmethionine as substrates, respectively
(Gaudillie
`re et al., 1984; Schultz et al., 1981). Interestingly, some of these
early works revealed that in Rubiaceae, OSB doubles as an intermediate in
the biosynthesis of anthraquinone species—of which still today not much is
240 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
known—and suggested that anthraquinone and naphthoquinone biosynth-
eses intersect at the level of DHNA.
The plant genes of phylloquinone biosynthesis have been identified only in
the past decade; for most of them using the genomic and genetic resources of
Arabidopsis thaliana and homology searches with the bacterial men genes as
query. Table II lists the cognate Arabidopsis proteins with their orthologues
in Synechocystis sp. PCC 6803 and E. coli. Two steps (reactions 6 and 7,
Fig. 2) remain to be characterized: for the first one, the gene is predicted but
has not been functionally confirmed; for the second one, orthology appears
to be missing (see below).
Fig. 4. The biosynthesis pathway of phylloquinone. 1, isochorismate synthase;
2, SEPHCHC synthase; 3, SHCHC synthase; 4, OSB synthase; 5, OSB-CoA ligase;
6, DHNA-CoA synthase; 7, DHNA-CoA thioesterase; 8, DHNA prenyltransferase;
9, demethylphylloquinone methyltransferase. SAH, s-adenosylhomocysteine;
SAM, s-adenosylmethionine; SEPHCHC, 2-succinyl-5-enolpyruvyl-6-hydroxy-3-
cyclohexene-1-carboxylic acid; SHCHC, (1R,6R)-2-succinyl-6-hydroxy-2,4-cyclohex-
adiene-1-carboxylic acid. EC numbers are indicated under each corresponding
reaction.
VITAMIN K
1
(PHYLLOQUINONE) 241
Author's personal copy
B. ISOCHORISMATE SYNTHASE/PHYLLO (REACTIONS 1–4)
Genetic approaches identified an Arabidopsis gene, termed PHYLLO
(At1g68890), that encodes an extraordinary protein composed of four mod-
ules homologous to the bacterial MenF (5.4.4.2), MenD (2.2.1.9), MenC
(4.2.1.113) and MenH (4.2.99.20) proteins, respectively (Fig. 5). The MenF
module lacks a complete chorismate binding domain, suggesting that it
cannot catalyze the conversion of chorismate into isochorismate (this
truncated domain appears nevertheless to be conserved from monocots to
dicots, pointing to a selective driving force for its maintenance. What this is,
however, is still an enigma). The Arabidopsis genome encodes in fact two
separated and catalytically active isochorismate synthases, ICS1 (At1g74710)
and ICS2 (At1g18870), that share about 80% identity (Garcion et al., 2008;
Strawn et al., 2007; Wildermuth et al., 2001). The ics1/ics2 double knockout
is devoid of phylloquinone (Garcion et al., 2008; Gross et al., 2006), thus
providing genetic evidence that PHYLLO is not sufficient for the de novo
synthesis of isochorismate, and that phylloquinone biosynthesis is dependent
upon a pool of isochorismate that is produced by separated isochorismate
synthases. As ICS1 bears most of the flux of isochorismate biosynthesis
(Garcion et al., 2008; Gross et al., 2006) and several plant genomes encode
for a single ICS copy, it is unclear if Arabidopsis ICS1 and ICS2 have
dedicated functions, or simply originate from recent duplication events and
are evolving separately. Whatever the case, isochorismate represents a meta-
bolic branch point where plant phylloquinone biosynthesis is likely to
TABLE II
Correspondence Between the Phylloquinone Biosynthesis Enzymes in Arabidopsis
and Synechocystis and Their Orthologues Involved in the Biosynthesis of
Menaquinone-8 in E. coli
A. thaliana
Synechocystis sp.
PCC6803 E. coli
Isochorismate synthase At1g74710 (ICS1)
At1g18870 (ICS2)
Slr0817 MenF
SEPHCHC synthase At1g68890 (PHYLLO) Sll0603 MenD
SHCHC synthase At1g68890 (PHYLLO) Slr1916 MenH
OSB synthase At1g68890 (PHYLLO) Sll0409 MenC
OSB-CoA ligase At1g30520 (AAE14) Slr0492 MenE
DHNA-CoA synthase At1g60550 (putative) Sll1127 MenB
DHNA-CoA thioesterase Unknown Slr0204 Unknown
DHNA phytyltransferase At1g60600 (ABC4) Slr1518 MenA
Demethylphylloquinone
methyltransferase
At1g23360 Sll1653 UbiE
242 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
compete with that of the plastid-produced hormone, salicylate, which also
uses isochorismate as a precursor (Wildermuth et al., 2001). Confirming this
view, the constitutive expression in tobacco chloroplasts of a bacterial iso-
chorismate pyruvate lyase, which converts isochorismate to salicylate, was
shown to result in an increase in salicylate level at the expense of phylloqui-
none (Verberne et al., 2007). At the time of the identification of PHYLLO,
MenH was still thought to correspond to DHNA-CoA thioesterase. The
menC and menH fused modules were therefore viewed as encoding for
domains that catalyzed reactions two steps apart from each other. It was
hypothesized that such an arrangement could indicate the existence of physi-
cal associations between PHYLLO, OSB-CoA ligase and DHNA synthase
(Gross et al., 2006). It is now evident that the PHYLLO MenDCH modules
catalyze consecutive reactions that lead to the synthesis of OSB. This does
not actually rule out that PHYLLO interacts with other enzymes in the
pathway, particularly because the fused structure of PHYLLO itself is sug-
gestive of a metabolon where biosynthetic intermediates are channelled from
one catalytic domain to the other.
Homology searches point to the existence of clusters of menF, menD,
menC and menH orthologues in green algae, mosses, diatoms and rhodo-
phytes. The menF domain of such clusters, in contrast to that of flowering
plants, features a full chorismate binding domain and is a priori functional.
AChlamydomonas reinhardtii cDNA corresponding to the menD orthologue
was shown to contain in-frame stop codons both upstream of the initiation
codon and at the end of the coding sequence, suggesting that in this species,
the menF, menD and menC modules are translated as separated polypeptides
(Lefebvre-Legendre et al., 2007). It remains therefore to establish if fused and
multifunctional enzymes equivalent to PHYLLO occur outside of the flower-
ing plant lineage.
Fig. 5. Arrangement of the functional domains of the Arabidopsis PHYLLO
protein and their approximate percentage of identity with their separated Men
orthologues in E. coli and with Arabidopsis isochorismate synthases 1 and
2 (AtICS1 and AtICS2). CTP, chloroplast transit peptide.
VITAMIN K
1
(PHYLLOQUINONE) 243
Author's personal copy
C. OSB-COA LIGASE (REACTION 5)
OSB-CoA ligase (6.2.1.26) activates the carboxyl group on the succinyl side
chain of OSB by creating a high-energy bond with the pantetheine moiety of
CoA (Kolkmann and Leistner, 1987;Fig. 2). Plant OSB-CoA ligase was
identified as part of a general characterization effort of the CoA ligase family
in Arabidopsis (Kim et al., 2008). A putative CoA ligase termed AAE14 (acyl
activating enzyme 14; the product of gene At1g30520) was singled out as one
of the top coexpressors of some previously identified phylloquinone biosyn-
thetic genes (At1g60600, DHNA phytyl transferase; At1g23360, demethyl-
phylloquinone methyltransferase; At1g68890, PHYLLO) and of the
predicted DHNA-CoA synthase (At1g60550; see below). Direct evidence
for the involvement of AAE14 in phylloquinone biosynthesis came from
the isolation of three independent T-DNA mutant lines corresponding to
insertions in the first intron, and fourth and ninth exons of At1g30520,
respectively; all of which lacked phylloquinone (Kim et al., 2008). The
T-DNA mutants were also found to accumulate OSB and could be partially
rescued by exogenous applications of DHNA (Kim et al., 2008). Expression
of At1g30520 cDNA was shown to fully restore menaquinone biosynthesis in
the E. coli menE knockout, thus verifying that AAE14 bore OSB-CoA ligase
activity (Kim et al., 2008).
D. DHNA-COA SYNTHASE/DHNA-COA THIOESTERASE (REACTIONS 6/7)
DHNA-CoA synthase (4.1.3.36) catalyzes the cyclization of OSB-CoA
(Fig. 2). The enzyme, which belongs to the crotonase-fold family, is often
termed in the literature and numerous databases as DHNA synthase or
naphthoate synthase, but it is clear that its reaction product is DHNA-
CoA, not DHNA (Jiang et al., 2010; Truglio et al., 2003). The enzyme’s
substrate, OSB-CoA, is highly unstable at physiological pH and has been
shown to spontaneously decompose in vitro into the spirodilactone form of
OSB (Fig. 6)(Heide et al., 1982; Meganathan and Bentley, 1979). Should
such a decomposition occur in vivo, it is not known if or how OSB spirodi-
lactone is recycled.
Bacterial DHNA-CoA synthases appear to fall into two catalytic classes.
Type I enzymes use a bound bicarbonate anion as a catalytic base, while type
II enzymes use the side-chain carboxylate of one of their acidic residues
(Jiang et al., 2010). Type I enzymes are consequently deemed bicarbonate
dependent and their type II counterparts bicarbonate independent (Jiang
et al., 2010). Sequence comparisons and phylogenetic reconstructions indi-
cate that cyanobacterial DHNA-CoA synthase and its predicted Arabidopsis
244 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
homologue belong to the type I category, suggesting that phylloquinone
biosynthesis is regulated by the intracellular level of bicarbonate.
The subsequent removal of CoA from DHNA, catalyzed by DHNA-CoA
thioesterase (3.1.2.-), has long puzzled the elucidation of the vitamin K
biosynthesis pathway. After being misattributed to DHNA-CoA synthase
(MenB), then to SHCHC synthase (MenH) in E. coli, it was later proposed
that the hydrolysis of DHNA-CoA, which like its OSB-CoA precursor
spontaneously decomposes at physiological pH, could be merely chemical
(Sakuragi and Bryant, 2006). But recent phylogenomics approaches in cya-
nobacteria detected putative CoA thioesterases, whose encoding genes were
arranged in clusters with known phylloquinone biosynthetic genes (Widhalm
et al., 2009). Deletion of the Synechocystis orthologue—gene slr0204
resulted in a dramatic decrease of the phylloquinone content in the knockout
cells, thus verifying the existence of a functional linkage between the putative
CoA thioesterase and phylloquinone biosynthesis (Widhalm et al., 2009).
Further investigations demonstrated that the knockout mutant accumulated
DHNA-CoA and could be chemically rescued with DHNA, but not with
OSB, thus pointing to the location of the blockage in the pathway (Widhalm
et al., 2009). The purified recombinant Slr0204 was shown to catalyze the
hydrolysis of DHNA-CoA and to display absolute preference for this sub-
strate. It is thought that such a substrate stringency may reflect the presence
of OSB-CoA upstream in the pathway, and whose enzymatic hydrolysis
would create a futile cycle in the phylloquinone biosynthesis pathway
(Widhalm et al., 2009). Although the Synechocystis slr0204 knockout
completely lacked DHNA-CoA thioesterease activity, low levels of phyllo-
quinone could still be detected in this mutant revealing the occurrence of a
basal chemical hydrolysis of DHNA-CoA in vivo (Widhalm et al., 2009).
Such a background decomposition likely explains why DHNA-CoA
Fig. 6. Hydrolysis and lactonization of OSB-CoA.
VITAMIN K
1
(PHYLLOQUINONE) 245
Author's personal copy
thioesterase did not show up in forward genetic screens aimed at identifying
men genes in bacteria, and one can therefore grasp through this illustrative
case the power of phylogenomics in predicting gene function based on the
detection of conserved physical associations of genes in genomes.
Except for the extremophilic rhodophytes Cyanidiales (Cyanidioschyzon
merolae and Cyanidium caldarium) and the cercozoan (Paulinella chromato-
phora), the plastid or chromatophore of which encode homologues of cya-
nobacterial DHNA-CoA thioesterase arranged in clusters with
phylloquinone biosynthetic genes (see Section VI), DHNA-CoA thioesterase
remains elusive in phylloquinone-synthesizing eukaryotes. Homology
searches do detect two pairs of Arabidopsis paralogs (At1g68260/
At1g68280, At1g35250/At1g35290) that share 17–28% of identity with Syne-
chocystis Slr0204, but these Arabidopsis genes have recently been shown to
encode orthologues of solanaceous methyl ketone synthases (Yu et al., 2010).
E. DHNA PHYTYL TRANSFERASE (REACTION 8)
DHNA phytyl transferase (2.5.1.-), an integral membrane protein, couples
the naphthoquinone ring to the phytyl side chain (Fig. 2). It was the first
enzyme specific to phylloquinone biosynthesis to be described in plants and
was initially identified in the Arabidopsis abc (aberrant chloroplast develop-
ment) T-DNA mutant series (Shimada et al., 2005). The cognate mutant-
designated abc4was shown to correspond to an insertion in the ninth exon of
gene At1g60600 and to lack phylloquinone (Shimada et al., 2005). Function-
al assignment was based on homology with Synechocystis sp. PCC 6803
DHNA phytyl transferase, which shares 41% identity with the At1g60600
protein (Shimada et al., 2005).
F. DEMETHYLPHYLLOQUINONE METHYLTRANSFERASE (REACTION 9)
Demethylphylloquinone methyltransferase (2.1.1.-) catalyzes the methyla-
tion of 2-phytyl-1,4-naphthoquinone (demethylphylloquinone) and corre-
sponds to the last step of the phylloquinone biosynthetic pathway (Fig. 2).
Mining the Arabidopsis genome with Synechocystis demethylphylloquinone
methyltransferase—the E. coli UbiE homologue, which doubles in the bio-
synthesis of ubiquinone, hence its name (Lee et al., 1997)—as query detected
the product of gene At1g23360 as a likely orthologue (Lohmann et al., 2006).
Expression of the cognate cDNA fully rescued phylloquinone biosynthesis in
the Synechocystis demethylphylloquinone methyltransferase knockout
(Lohmann et al., 2006). In parallel, a T-DNA line corresponding to an
insertion in the seventh exon of At1g23360 was found to be devoid of
246 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
phylloquinone and to accumulate 2-phytyl-1,4-naphthoquinone, thus estab-
lishing definite evidence that this gene encodes the only demethylphylloqui-
none methyltransferase in Arabidopsis (Lohmann et al., 2006).
G. MUTANT PHENOTYPE
Arabidopsis lines corresponding to phyllo (the fused SEPHCHC synthase–
SHCHC synthase–OSB synthase), aae14 (OSB-CoA ligase) and abc4
(DHNA phytyl transferase) knockouts and to the double knockout ics1/
ics2 (isochorismate synthase 1 and 2) display loss of photoautotrophy and
are seedling lethal (Gross et al., 2006; Kim et al., 2008; Shimada et al., 2005).
A few pale-green leaves can be obtained from these mutants providing that
they are grown on a medium containing sucrose and under low illumination,
but even so the plants eventually stop developing. Analyses of the phyllo and
abc4 mutants showed that the lack of phylloquinone results in the disruption
of photosystem I assembly (Gross et al., 2006; Shimada et al., 2005). Plasto-
quinone level and photosystem II activity were also shown to be dramatically
reduced in the abc4 knockout, while photosystem II was found to be only
moderately affected in the phyllo mutant (Gross et al., 2006; Shimada et al.,
2005). In contrast, green algal and cyanobacterial mutants, which are
blocked in the formation of the naphthoquinone ring or its prenylation, are
able to recruit plastoquinone into the A
1
site of photosystem I in place of
phylloquinone and—though being sensitive to high light intensity—can grow
photoautotrophically (Johnson et al., 2000; Lefebvre-Legendre et al., 2007).
The demethylphylloquinone methyltransferase knockout is the sole viable
phylloquinone-deficient mutant in plants. The reduction in photosynthetic
efficiency and number of photosystem I subunits observed in the cognate
Arabidopsis insertion line indicate nonetheless that the replacement of
phylloquinone by demethylphylloquinone is not fully functional (Lohmann
et al., 2006).
H. SUBCELLULAR LOCALIZATION OF PHYLLOQUINONE
BIOSYNTHETIC ENZYMES
Early radiolabelling experiments showed that the prenylation and methylation
steps of plant phylloquinone biosynthesis were associated with the chloroplast
membranes; the two activities appeared to be localized in separate subfractions:
the prenylation occurring in the chloroplast envelope and the methylation in
thylakoids (Gaudillie
`re et al., 1984; Kaiping et al., 1984; Schultz et al.,1981).
Cloning of Arabidopsis DHNA phytyl transferase and demethylphylloquinone
methyl transferase later confirmed that both enzymes possess N-terminal
VITAMIN K
1
(PHYLLOQUINONE) 247
Author's personal copy
signalling peptides and are indeed targeted to plastids (Lohmann et al., 2006;
Shimada et al.,2005). Similar findings were obtained for isochorismate synthase
1and2(Garcion et al., 2008; Strawnet al., 2007), PHYLLO (Gross et al.,2006)
and OSB-CoA ligase (Kim et al.,2008). Contrasting with this apparent all-
plastidial localization of the pathway, predicted DHNA-CoA synthases from
dicots and monocots have N-terminal extensions that contain a canonical
peroxisomal targeting signal type 2 (RLx
5
HL) and proteomic approaches
have identified the putative Arabidopsis enzyme and its spinach orthologue in
purified peroxisomes (Babujee et al., 2010; Reumann et al.,2007). Expression in
onion epidermal cells of the Arabidopsis proteinfused at its C-terminal end to a
fluorescent reporter protein further verified that the resulting construct was
imported into peroxisomes (Babujee et al.,2010). The green alga C. reinhardtii
and moss P. patens orthologues, however, lack a peroxisomal targeting signal,
so as do their cyanidiale C. merolae and C. caldarium and cercozoan P. chro-
matophora counterparts, which are chloroplast or chromatophore encoded,
thus indicating that the targeting of DHNA-CoA synthase to peroxisome is
not ubiquitous in phylloquinone-synthesizing eukaryotes.
Interestingly, the preceding enzyme, OSB-CoA ligase, displays in most
monocotyledonous and dicotyledonous species a predicted peroxisomal target-
ing signal. In this case, it corresponds to a C-terminal tripeptide (SSL, SNL,
SRL or SKL depending on the species) that typifies a peroxisomal targeting
signal type 1 (Babujee et al.,2010). N-terminally fused fluorescent versions of
Arabidopsis OSB-CoA ligase (AAE14) or of its last 10 residues containing the
SSL signal were expressed in onion epidermal cells and confirmed here again
that the hybrid proteins were targeted to peroxisomes (Babujee et al.,2010).
These exciting observations imply that the activation of OSB and its cyclization
into DHNA-CoA occur in peroxisomes, thus requiring the shuttling of phyllo-
quinone biosynthetic precursors in and out of plastids and peroxisomes. One
should note, however, that transient expressions of C-terminally tagged fluo-
rescent versions of AAE14 or its first 120 residues in Arabidopsis leaf proto-
plasts and tobacco leaf mesophyll cells, respectively, have demonstrated that the
enzyme also bears a functional plastid targeting presequence and is targeted to
chloroplasts (Kim et al., 2008). The obvious bias of each of the aforementioned
fusion strategies is that the reporter protein conceals either the peroxisomal
targeting signal type 1 (C-terminal fusion) or the plastid targeting presequence
(N-terminal fusion). Although the current view is that AAE14 could be dual
targeted, direct identification using proteomics approaches and/or assays of
OSB-CoA ligase in purified peroxisomes and chloroplasts is needed to precisely
determine the subcellular localization of this enzyme.
As for the hydrolysis of DHNA-CoA, preliminary data from our labora-
tory indicated that although DHNA-CoA thioesterase was detectable in
248 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
whole extracts of pea and Arabidopsis leaves, it was absent in the
corresponding preparations of purified chloroplasts (unpublished results).
As previously suggested (Babujee et al., 2010), this makes the localization of
DHNA-CoA hydrolysis in peroxisomes all the more likely. Figure 7 sum-
marizes the current knowledge concerning the subcellular compartmentation
of the phylloquinone biosynthesis pathway in Arabidopsis.
Fig. 7. Subcellular localization of the phylloquinone biosynthetic enzymes in
Arabidopsis. Letters in brackets specify the type of experimental evidence: fusion to
a fluorescent reporter protein and transient expression in Arabidopsis leaf or leaf
protoplasts (a), onion epidermal cell (b), tobacco mesophyll cells or leaf protoplasts
(c); C-terminal fusion to V5-6xHis epitope and stable expression under the control of
native promoter in Arabidopsis transgenics, and immunolocalization (d); in vitro
import assay in chloroplasts purified from pea seedling (e); subcellular fractionation
and identification using mass sprectrometry (f). Dashed arrows indicate putative
transport steps between plastid and peroxisome, or the possible occurrence of
DHNA-CoA hydrolysis in peroxisome (reaction 7). Evidence from our laboratory
indicates that DHNA-CoA thioesterase activity is lacking in chloroplasts
(Widhalm J.R., unpublished data).
VITAMIN K
1
(PHYLLOQUINONE) 249
Author's personal copy
VI. EVOLUTION OF NAPHTHOQUINONE
BIOSYNTHESIS IN PHOTOSYNTHETIC EUKARYOTES
Isoprenoid naphthoquinones are evolutionarily the most ancient of all con-
jugated quinones, their biosynthesis in prokaryotes predating the rise of
atmospheric dioxygen level ca. 2.5 billion years ago (Schoepp-Cothenet
et al., 2009). Menaquinones have thus been detected in most prokaryotic
lineages (Collins and Jones, 1981), rhodophytes (Yoshida et al., 2003) and
diatoms (Ikeda et al., 2008). Phylloquinone in contrast appears to be restrict-
ed to some cyanobacterial species, green algae and plants (Collins and Jones,
1981; Lefebvre-Legendre et al., 2007; Oostende et al., 2008).
All the phylloquinone biosynthetic enzymes identified so far in photosyn-
thetic eukaryotes are nuclear encoded. However, the cyanidiale orthologues—
with the exception of the MenG orthologue (see below)— are plastid encoded,
indicating that the cognate genes have likely been retained from the former
cyanobacterial endosymbiont. While it might therefore seem that such genes
are merely of direct cyanobacterial descent, some phylogenetic studies suggest
that this is not so for the menF, menD, menC, menE and menB orthologues,
which are in fact more closely related to the chlorobi/-proteobacteria lineage
(Gross et al., 2008). To reconcile this surprising genealogy with the
fact that the corresponding enzymes are plastid encoded in cyanidiales, it
has been proposed that men genes originating from an organism of the
chlorobi/-proteobacteria descent have been captured through horizontal
gene transfer by the free living cyanobacterial progenitor of plastids, that is,
prior to its endosymbiosis, and have replaced their pre-existing cyanobacterial
counterparts (Gross et al., 2008). The remnant of this horizontal gene transfer
would now ‘survive’ as a men gene cluster in the plastid genome of modern-day
cyanidiales and in the nuclear genome of diatoms, green algae and plants—the
aforementioned tetramodular PHYLLO locus (Gross et al., 2008).
In contrast, the cyanidiales menA homologue—also part of such a men
gene cluster—would have been acquired through an independent horizontal
gene transfer with a prokaryotic donor, whose identity remains unclear
(Gross et al.,2008). So is the case for the menA homologue of diatoms,
which would define another event of horizontal gene transfer that occurred in
the nucleus (Gross et al., 2008). As for the nuclear-encoded menG homo-
logue, phylogenies suggest that it would originate from -proteobacteria in
both cyanidiales and diatoms (Gross et al., 2008). Plants and green algae
further complicate the picture, having menE homologues that would branch
from the -proteobacteria lineage and menA and menG homologues that
would be of cyanobacterial ancestry (Gross et al., 2008). In essence, accord-
ing to such phylogenetic reconstructions, the eukaryotic genes involved in the
250 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
formation of isoprenoid naphthoquinones display a high degree of evolu-
tionary chimerism that varies with the lineage considered, owing to multiple
and unrelated events of horizontal gene transfer and/or gene losses (summar-
ized in Fig. 8).
Fig. 8. Tentative scenario for the evolution of the men genes in photosynthetic
eukaryotes as proposed by Gross et al. (2008). Arrows symbolize gene transfers. Note
that menH homologues are not detected in the plastid genomes of cyanidiales. As for
menA, the cyanobacterial progenitor of plastids would have harboured two cognate
homologues: menA1, of cyanobacterial descent and menA2, acquired by HGT from
an unknown prokaryotic donor. Cyanidiales would have lost menA1 and retained
menA2; the opposite would have happened in plants and green algae. An alternative
explanation would be that menA1 was acquired from cyanobacteria by direct HGT in
the nucleus of the green algal ancestor. HGT, horizontal gene transfer.
VITAMIN K
1
(PHYLLOQUINONE) 251
Author's personal copy
One should, however, point that comparative genomics of modern cyano-
bacteria, cyanidiales and chlorobi/-proteobacteria question some parts of this
evolutionary model. For instance, the genomic organization of the men and
DHNA-CoA thioesterase homologues of certain present-day cyanobacteria
and of cyanidiales displays striking similarities that are not conserved in the
chlorobi/-proteobacteria lineage (Fig. 9). Although such a conservation might
be purely coincidental or driven by identical selective constraints (e.g. transcrip-
tional regulation), it could also point to an overlooked phylogenetic closeness
between the menaquinone biosynthetic genes of cyanidiales and some of their
homologues involved in phylloquinone biosynthesis in cyanobacteria.
Fig. 9. Organization of the phylloquinone/menaquinone biosynthetic gene clus-
ters in representative species of cyanobacteria/cercozoan, cyanidiales, -proteobac-
teria and chlorobi. The dashed frame highlights the conserved arrangement of the men
and DHNA-CoA thioesterase (THIO) homologues in cyanobacteria (N. punctiforme,
Nostoc punctiforme;P. marinus,Prochlorococcus marinus; S. sp. CC9605, Synechoc-
coccus sp. CC9605), cercozoan (Cerc.) (P. chromatophora,Paulinella chromatophora)
and cyanidiales (C. caldarium,Cyanidium caldarium;C. merolae,Cyanidioschyzon
merolae). The gene cluster of the cercozoan species P. chromatophora is located in a
plastid-like organelle called the chromatophore; the later is thought to originate from
a recent endosymbiosis of a cyanobacterium of the Prochlorococcus/Synechococcus
lineage. A. hydrophila,Aeromonas hydrophila;C. limicola,Chlorobium limicola;
C. ferrooxidans,Chlorobium ferrooxidans.
252 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
VII. PHYLLOQUINONE TURNOVER
Our knowledge of vitamin K metabolism is almost exclusively restricted to
mammals and is largely extrapolated from the data obtained with tocopher-
ols. Pharmacological studies have shown that phylloquinone is rapidly cat-
abolized into shortened side-chain carboxylic acids, which are excreted in
urine as water-soluble glucuronic acid conjugates (Harrington et al., 2007;
Landes et al., 2003). The enzymatic reactions that lead to the shortening of
the side chain are not known sensu stricto. Nevertheless, as tocopherols and
phylloquinone share the same phytyl side chain and in mammals the pro-
ducts of tocopherol catabolism come from !-hydroxylation and subsequent
-oxidation of the side chain, it is believed that phylloquinone follows a
similar catabolic route (Harrington et al., 2007; Landes et al., 2003).
Close to nothing is known about the catabolism of vitamin K in plants. One
study showed that non-physiological doses of phylloquinone can be fed to pea
stem sections, and that more than 90% of the incorporated vitamin could be
recovered after 18-h incubation (Gaunt and Stowe, 1967). The occurrence of
degradation products of phylloquinone in plants is not documented.
VIII. ENGINEERING OF PHYLLOQUINONE
IN PLANTS
There has not been so far any dedicated engineering of phylloquinone in
plants; the only data available are for tobacco transgenics engineered for
salicylic acid biosynthesis, and functional complementation experiments in
Arabidopsis. Thus, in tobacco, the overexpression of an E. coli isochorismate
synthase targeted to plastids led to a fourfold increase of phylloquinone
above wild-type levels (Verberne et al., 2007). In Arabidopsis, overexpression
of demethylphylloquinone methyltransferase or OSB-CoA ligase did not
change phylloquinone content compared to that of wild-type plants (Kim
et al., 2008; Lohmann et al., 2006).
IX. CONCLUDING REMARKS
Findings from the most recent studies of phylloquinone metabolism in plants
illustrate once more that the architecture of plant secondary metabolism is
hardly inferable from previous work in microorganisms. Witness the unprec-
edented and extraordinary multifunctional PHYLLO, the occurrence of
functional redundancies (isochorismate synthases), the apparent lack of
VITAMIN K
1
(PHYLLOQUINONE) 253
Author's personal copy
orthology (DHNA-CoA thioesterase) and the split of phylloquinone biosyn-
thesis between plastids and peroxisomes.
Besides the identification of the ‘missing’ DHNA-CoA thioesterase and
the characterization of DHNA-CoA synthase—especially with regards to its
possible regulation by carbonate—one of the priorities of the research on
phylloquinone biosynthesis in plants is now to determine the arrangement of
the plastid and peroxisomal branches that lead to the formation of the
naphthoquinone ring. One cannot indeed overemphasize that the cognate
transport steps between these two organelles are as determinant for the flux
of phylloquinone production as the biosynthetic enzymes themselves. It will
therefore need to be established which biosynthetic intermediates are trans-
ported, and if specific transporters are involved. Such future investigations
are predictably challenging owing to the very low abundance and high
instability of most of the naphthoquinone ring’s biosynthetic intermediates,
and to the difficulty inherently attached to the isolation of reasonably pure
plant organelles and to the functional study of integral proteins.
Another area to further explore is the integration of phylloquinone in the
metabolic network of plastids. As mentioned earlier, there is experimental
evidence in Arabidopsis and tobacco that the biosynthetic pathways of
phylloquinone and salicylic acid intersect through isochorismate. It is prob-
able that plants tightly regulate this metabolic node because salicylate is
massively produced in response to certain stresses, while phylloquinone is
absolutely needed as a redox cofactor. One fascinating hypothesis could be
that the flux of isochorismate usage towards phylloquinone biosynthesis
depends on a—yet-to-be demonstrated— physical association between iso-
chorismate synthase and the multifunctional PHYLLO, thus creating a
metabolon from chorismate up to OSB. Such a scenario could also explain
why flowering plants have maintained a truncated and catalytically inactive
MenF domain in PHYLLO; it could serve for instance as a recognition/
binding domain to assemble the metabolon.
Although species specific, the flux split at the level of DHNA between the
naphthoquinone and anthraquinone biosynthetic pathways is even more
enigmatic. Here, again one can expect that anthraquinone-producing species
must have evolved strategies to commit a steady flux of DHNA towards
phylloquinone biosynthesis.
Through phytyl-diphosphate as a common precursor of the isoprenyl
side chain, we also know that phylloquinone is connected to the metabolism
of tocopherols and chlorophyll. However, we cannot currently tell to
what extent a change in the biosynthesis flux of one of this compound will
impact the others. Answering this question is important for the basic under-
standing of the metabolic network of plastid isoprenoids as well as for
254 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
engineering purposes. For instance, an increase in phylloquinone level that
would occur at the expense of chlorophyll and/or tocopherols could nega-
tively impact photosynthesis or paradoxically decrease the nutritional value
of the derived plant products. The remark stands for the engineering of
tocopherol levels.
As for the fused VKORC1-PDI enzyme, the research problem is now to
identify the proteins whose folding is connected to the reduction of phyllo-
quinone in plastids. The fate of the formed quinol is also intriguing because
the phylloquinone/phylloquinol ratio appears to be remarkably stable in
plants and Synechocystis (Oostende et al., 2008; Widhalm et al., 2009).
A tacit conclusion is that phylloquinol is re-oxidized, and that these organ-
isms can sense the redox status of their phylloquinone pool.
ACKNOWLEDGEMENTS
G. J. B. and C. v. O. dedicate this chapter to the memory of Dr. Philippe
Raymond, whose mentoring and support have been seminal to their works.
Research in our laboratory is made possible in part by National Science
Foundation Grant MCB-0918258 to G. J. B. and by startup funds provided
by the Center for Plant Science Innovation and the Nebraska Tobacco
Settlement Biomedical Research Development Funds.
REFERENCES
Academy and of Pediatrics, Committee on Fetus and Newborn (2003). Policy state-
ment. Controversies concerning vitamin K and the newborn. Pediatrics 112,
191–192.
Almquist, H. J. and Stokstad, E. L. R. (1935). Hemorrhagic chick disease of dietary
origin. The Journal of Biological Chemistry 111, 105–113.
Babujee, L., Wurtz, V., Ma, C., Lueder, F., Soni, P., van Dorsselaer, A. and
Reumann, S. (2010). The proteome map of spinach leaf peroxisomes indi-
cates partial compartmentalization of phylloquinone (vitamin K1) biosyn-
thesis in plant peroxisomes. Journal of Experimental Botany 61, 1441–1453.
Bach, A. U., Anderson, S. A., Foley, A. L., Williams, E. C. and Suttie, J. W. (1996).
Assessment of vitamin K status in human subjects administered ‘‘minidose’’
warfarin. The American Journal of Clinical Nutrition 64, 894–902.
Ben-Shem, A., Frolow, F. and Nelson, N. (2003). Crystal structure of plant photo-
system I. Nature 426, 630–635.
Binkley, S. B., MacCorquodale, D. W., Thayer, S. A. and Doisy, E. A. (1939). The
isolation of vitamin K
1
.The Journal of Biological Chemistry 130, 219–234.
Booth, S. L. and Sadowski, J. A. (1997). Determination of phylloquinone in foods by
high-performance liquid chromatography. Methods in Enzymology 282,
446–456.
Booth, S. L. and Suttie, J. W. (1998). Dietary intake and adequacy of vitamin K. The
Journal of Nutrition 128, 785–788.
VITAMIN K
1
(PHYLLOQUINONE) 255
Author's personal copy
Booth, S. L., Tucker, K. L., Chen, H., Hannan, M. T., Gagnon, D. R.,
Cupples, L. A., Wilson, P. W., Ordovas, J., Schaefer, E. J., Dawson-
Hughes, B. and Kiel, D. P. (2000). Dietary vitamin K intakes are associated
with hip fracture but not with bone mineral density in elderly men and
women. The American Journal of Clinical Nutrition 71, 1201–1208.
Boudreaux, B., MacMillan, F., Teutloff, C., Agalarov, R., Gu, F., Grimaldi, S.,
Bittl, R., Brettel, K. and Redding, K. (2001). Mutations in both sides of
the photosystem I reaction center identify the phylloquinone observed by
electron paramagnetic resonance spectroscopy. The Journal of Biological
Chemistry 276, 37299–37306.
Brettel, K., Se
´tif, P. and Mathis, P. (1987). Flash-induced absorption changes in
photosystem I at low temperatures: Evidence that the electron acceptor A
1
is vitamin K
1
.FEBS Letters 203, 220–224.
Bridge, A., Barr, R. and Morre
´, D. J. (2000). The plasma membrane NADH oxidase
of soybean has vitamin K1 hydroquinone oxidase activity. Biochimica et
Biophysica Acta 1463, 448–458.
Chu, P. H., Huang, T. Y., Williams, J. and Stafford, D. W. (2006). Purified vitamin K
epoxide reductase alone is sufficient for the conversion of vitamin K
epoxide to vitamin K and vitamin K to vitamin KH
2
.Proceedings of the
National Academy of Sciences of the United States of America 103,
19308–19313.
Collins, M. D. and Jones, D. (1981). Distribution of isoprenoid quinone structural
types in bacteria and their taxonomic implications. Microbiological Reviews
45, 316–354.
Dam, H. (1929). Cholesterinositoffweschel in huhnereiern und huhnchen. Biochem-
ische Zeitschrift 215, 475–492.
Dam, H. (1935). The antihaemorrhagic vitamin of the chick. The Biochemical Journal
29, 1273–1285.
Dam, H. and Schønheyder, F. (1934). A deficiency disease in chicks resembling
scurvy. The Biochemical Journal 28, 1355–1359.
Dam, H. and Schønheyder, F. (1936). The occurrence and chemical nature of vitamin
K. The Biochemical Journal 30, 897–901.
Dansette, P. and Azerad, R. (1970). A new intermediate in naphthoquinone and
menaquinone biosynthesis. Biochemical and Biophysical Research Commu-
nications 40, 1090–1095.
Davidson, K. W. and Sadowski, J. A. (1997). Determination of vitamin K compounds
in plasma or serum by high-performance liquid chromatography using
postcolum chemical reduction and fluorimetric detection. Methods in Enzy-
mology 282, 408–421.
Doisy, E. A., Binkley, S. B. and Thayer, S. A. (1941). Vitamin K. Chemical Reviews
28, 477–517.
Feskanich, D., Weber, P., Willett, W. C., Rockett, H., Booth, S. L. and Colditz, G. A.
(1999). Vitamin K intake and hip fractures in women: A prospective study.
The American Journal of Clinical Nutrition 69, 74–79.
Food and Nutrition Board, Institute of Medicine (2001). In Dietary Reference Intakes
for Vitamin A, Vitamin K, Arsenic, Boron, Chromium, Copper, Iodine,
Iron, Manganese, Molybdenum, Nickel, Silicon, Vanadium, and Zinc,
pp. 162–196. National Academy Press, Washington, DC.
Fraser, P. D., Pinto, M. E. S., Holloway, D. E. and Bramley, P. M. (2000). Applica-
tion of high-performance liquid chromatography with photodiode array
detection to the metabolic profiling of plant isoprenoids. The Plant Journal
24, 551–558.
256 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
Fromme, P. and Grotjohann, I. (2006). Structural analysis of cyanobacterial photo-
system I. In Photosystem I: The Light-Driven plastocyanin:Ferredoxin
Oxidoreductase, (J. H. Golbeck, ed.), pp. 205–222. Springer, Dordrecht.
Furt, F., Oostende, C. V., Widhalm, J. R., Dale, M. A., Wertz, J. and Basset, G. J. C.
(2010). A bimodular oxidoreductase mediates the specific reduction of
phylloquinone (vitamin K
1
) in chloroplasts. The Plant Journal 64, 38–46.
Garcion, C., Lohmann, A., Lamodie
`re, E., Catinot, J., Buchala, A., Doermann, P.
and Me
´traux, J. P. (2008). Characterization and biological function of the
ISOCHORISMATE SYNTHASE2 gene of Arabidopsis. Plant Physiology
147, 1279–1287.
Gaudillie
`re, J.-P., d’Harlingue, A., Camara, B. and Mone
´ger, R. (1984). Prenylation
and methylation reactions in phylloquinone (vitamin K
1
) synthesis in Cap-
sicum annuum plastids. Plant Cell Reports 3, 240–242.
Gaunt, J. K. and Stowe, B. B. (1967). Uptake and metabolism of vitamins E and K by
pea stem sections. Plant Physiology 42, 859–862.
Goodstadt, L. and Ponting, C. P. (2004). Vitamin K epoxide reductase: Homology,
active site and catalytic mechanism. Trends in Biochemical Sciences 29,
289–292.
Gross, J., Cho, W. K., Lezhneva, L., Falk, J., Krupinska, K., Shinozaki, M., Seki, M.,
Hermann, R. G. and Meurer, J. (2006). A plant locus essential for phyllo-
quinone (vitamin K
1
) biosynthesis originated from a fusion of four eubac-
terial genes. The Journal of Biological Chemistry 281, 17189–17196.
Gross, J., Meurer, J. and Bhattacharya, D. (2008). Evidence of a chimeric genome in
the cyanobacterial ancestor of plastids. BMC Evolutionary Biology 8, 117.
Guergova-Kuras, M., Boudreaux, B., Joliot, A., Joliot, P. and Redding, K. (2001).
Evidence for two active branches for electron transfer in photosystem I.
Proceedings of the National Academy of Sciences of the United States of
America 98, 4437–4442.
Harrington, D. J., Booth, S. L., Card, D. J. and Shearer, M. J. (2007). Excretion of the
urinary 5C- and 7C-aglycone metabolites of vitamin K by young adults
responds to changes in dietary phylloquinone and dihydrophylloquinone
intakes. The Journal of Nutrition 137, 1763–1768.
Heide, L., Kolkmann, R., Arendt, S. and Leistner, E. (1982). Enzymic synthesis of
o-succinylbenzoyl-CoA in cell-free extracts of anthraquinone producing
Galium mollugo L. cell suspension cultures. Plant Cell Reports 1, 180–182.
Hiratsuka, T., Furihata, K., Ishikawa, J., Yamashita, H., Itoh, N., Seto, H. and
Dairi, T. (2008). An alternative menaquinone biosynthetic pathway
operating in microorganisms. Science 321, 1670–1673.
Holst, W. F. and Halbrook, E. R. (1933). A ‘scurvy-like’ disease in chicks. Science 77,
354.
Hutson, K. G. and Threlfall, D. R. (1980). Asymetric incorporation of 4-(20-carbox-
yphenyl)-4-oxobutyrate into phylloquinone by Zea mays.Phytochemistry
19, 535–537.
Ichikawa, T., Horie-Inoue, K., Ikeda, K., Blumberg, B. and Inoue, S. (2006).
Steroid and xenobiotic receptor SXR mediates vitamin K2-activated
transcription of extracellular matrix-related genes and collagen accumula-
tion in osteoblastic cells. The Journal of Biological Chemistry 281,
16927–16934.
Ikeda, Y., Komura, M., Watanabe, M., Minami, C., Koike, H., Itoh, S., Kashino, Y.
and Satoh, K. (2008). Photosystem I complexes associated with fucoxan-
thin-chlorophyll-binding proteins from a marine centric diatom, Chaeto-
ceros gracilis.Biochimica et Biophysica Acta 1777, 351–361.
VITAMIN K
1
(PHYLLOQUINONE) 257
Author's personal copy
Inaba, K., Takahashi, Y. H. and Ito, K. (2004). DsbB elicits a red-shift of bound
ubiquinone during the catalysis of DsbA oxidation. The Journal of
Biological Chemistry 279, 6761–6768.
Iwamoto, J., Takeda, T. and Ichimura, S. (2001). Effect of menatetrenone on bone
mineral density and incidence of vertebral fractures in postmenopausal
women with osteoporosis: A comparison with the effect of etidronate.
Journal of Orthopaedic Science 6, 487–492.
Jiang, M., Chen, M., Guo, Z. F. and Guo, Z. (2010). A bicarbonate cofactor
modulates 1,4-dihydroxy-2-naphthoyl-coenzyme a synthase in menaqui-
none biosynthesis of Escherichia coli.The Journal of Biological Chemistry
285, 30159–30169.
Johnson, T. W., Shen, G., Zybailov, B., Kolling, D., Reategui, R., Beauparlant, S.,
Vassiliev, I. R., Bryant, D. A., Jones, A. D., Golbeck, J. H. and
Chitnis, P. R. (2000). Recruitment of a foreign quinone into the A(1) site
of photosystem I. I. Genetic and physiological characterization of phyllo-
quinone biosynthetic pathway mutants in Synechocystis sp. PCC 6803. The
Journal of Biological Chemistry 275, 8523–8530.
Jordan, P., Fromme, P., Witt, H. T., Klukas, O., Saenger, W. and Krauss, N. (2001).
Three-dimensional structure of cyanobacterial photosystem I at 2.5 A reso-
lution. Nature 411, 909–917.
Kaiping, S., Soll, J. and Schultz, G. (1984). Site of methylation of 2-phytyl-1,
4-naphthoquinol in phylloquinone (vitamin K
1
) synthesis in spinach chlor-
oplasts. Phytochemistry 23, 89–91.
Kim, H. U., van Oostende, C., Basset, G. J. and Browse, J. (2008). The AAE14 gene
encodes the Arabidopsis o-succinylbenzoyl-CoA ligase that is essential for
phylloquinone synthesis and photosystem I function. The Plant Journal 54,
272–283.
Knapen, M. H., Schurgers, L. J. and Vermeer, C. (2007). Vitamin K2 supplementa-
tion improves hip bone geometry and bone strength indices in postmeno-
pausal women. Osteoporosis International 18, 963–972.
Kolkmann, R. and Leistner, E. (1987). 4-(20-Carboxyphenyl)-4-oxobutyryl coenzyme
A ester, an intermediate in vitamin K2 (menaquinone) biosynthesis. Zeits-
chrift fu
¨r Naturforschung. C 42, 1207–1214.
Landes, N., Birringer, M. and Brigelius-Flohe
´, R. (2003). Homologous metabolic and
gene activating routes for vitamins E and K. Molecular Aspects of Medicine
24, 337–344.
Lee, P. T., Hsu, A. Y., Ha, H. T. and Clarke, C. F. (1997). A C-methyltransferase
involved in both ubiquinone and menaquinone biosynthesis: Isolation and
identification of the Escherichia coli ubiE gene. Journal of Bacteriology 179,
1748–1754.
Lefebvre-Legendre, L., Rappaport, F., Finazzi, G., Ceol, M., Grivet, C.,
Hopfgartner, G. and Rochaix, J. D. (2007). Loss of phylloquinone in
Chlamydomonas affects plastoquinone pool size and photosystem II synthe-
sis. The Journal of Biological Chemistry 282, 13250–13263.
Li, J., Lin, J. C., Wang, H., Peterson, J. W., Furie, B. C., Furie, B., Booth, S. L.,
Volpe, J. J. and Rosenberg, P. A. (2003). Novel role of vitamin K in
preventing oxidative injury to developing oligodendrocytes and neurons.
The Journal of Neuroscience 23, 5816–5826.
Li, W., Schulman, S., Dutton, R. J., Boyd, D., Beckwith, J. and Rapoport, T. A.
(2010). Structure of the bacterial homologue of vitamin K epoxide reduc-
tase. Nature 463, 507–512.
258 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
Lichtenthaler, H. K. (1999). The 1-deoxy-D-xylulose-5-phosphate pathway of iso-
prenoid biosynthesis in plants. Annual Review of Plant Physiology and Plant
Molecular Biology 50, 47–65.
Lohmann, A., Schottler, M. A., Brehelin, C., Kessler, K., Bock, R., Cahoon, E. D.
and Dormann, P. (2006). Deficiency in phylloquinone (vitamin K
1
) methyl-
ation affects prenyl quinone distribution, photosystem I abundance, and
anthocyanin accumulation in the Arabidopsis AtmenG mutant. The Journal
of Biological Chemistry 281, 40461–40472.
Lu
¨thje, S., Van Gestelen, P., Co
´rdoba-Pedregosa, M. C., Gonza
´lez-Reyes, J. A.,
Asard, H., Villalba, J. M. and Bo
¨ttger, M. (1998). Quinones in plant plasma
membranes—A missing link? Protoplasma 205, 43–51.
MacCorquodale, D. W., Binkley, S. B., Thayer, S. A. and Doisy, E. A. (1939a). On
the constitution of vitamin K
1
.Journal of the American Chemical Society 61,
1928–1929.
MacCorquodale, D. W., Cheney, L. W., Binkley, S. B., Holcomb, W. F.,
McKee, R. W., Thayer, S. A. and Doisy, E. A. (1939b). The constitution
and synthesis of vitamin K
1
.The Journal of Biological Chemistry 131,
357–370.
McCarthy, P. T., Harrington, D. J. and Shearer, M. J. (1997). Assay of phylloquinone
in plasma by high-performance liquid chromatography with electrochemical
detection. Methods in Enzymology 282, 421–438.
McFarlane, W. D., Graham, W. R., Jr. and Hall, G. E. (1931). Studies in protein
nutrition of the chick. I. The influence of different protein concentrates on
the growth of baby chicks, when fed as the source of protein in various
simplified diets. The Journal of Nutrition 4, 331–349.
McKee, R. W., Binkley, S. B., MacCorquodale, D. W., Thayer, S. A. and Doisy, E. A.
(1939). The isolation of vitamin K
1
and K
2
.Journal of the American Chemi-
cal Society 61, 1295.
McKenna, M., Henninger, M. D. and Crane, F. L. (1964). A second naphthoquinone
in spinach chloroplasts. Nature 203, 524–525.
McLean, R. R., Booth, S. L., Kiel, D. P., Broe, K. E., Gagnon, D. R., Tucker, K. L.,
Cupples, L. A. and Hannan, M. T. (2006). Association of dietary and
biochemical measures of vitamin K with quantitative ultrasound of the
heel in men and women. Osteoporosis International 17, 600–607.
Meganathan, R. and Bentley, R. (1979). Menaquinone (vitamin K
2
) biosynthesis:
Conversion of o-succinylbenzoic acid to 1,4-dihydroxy-2-naphthoic acid by
Mycobacterium phlei enzymes. Journal of Bacteriology 140, 92–98.
Mimuro, M., Tsuchiya, T., Inoue, H., Sakuragi, Y., Itoh, Y., Gotoh, T.,
Miyashita, H., Bryant, D. A. and Kobayashi, M. (2005). The secondary
electron acceptor of photosystem I in Gloeobacter violaceus PCC 7421 is
menaquinone-4 that is synthesized by a unique but unknown pathway.
FEBS Letters 579, 3493–3496.
Oostende, C., Widhalm, J. R. and Basset, G. J. (2008). Detection and quantification
of vitamin K(1) quinol in leaf tissues. Phytochemistry 69, 2457–2462.
Petersen, J., Stehlik, D., Gast, P. and Thurnauer, M. (1987). Comparison of the
electron spin polarized spectrum found in plant photosystem I and in iron-
depleted bacterial reaction centers with time-resolved K-band EPR; evi-
dence that the photosystem I acceptor A
1
is a quinone. Photosynthesis
Research 14, 15–29.
Rejnmark, L., Vestergaard, P., Charles, P., Hermann, A. P., Brot, C., Eiken, P. and
Mosekilde, L. (2006). No effect of vitamin K1 intake on bone mineral
density and fracture risk in perimenopausal women. Osteoporosis Interna-
tional 17, 1122–1132.
VITAMIN K
1
(PHYLLOQUINONE) 259
Author's personal copy
Reumann, S., Babujee, L., Ma, C., Wienkoop, S., Siemsen, T., Antonicelli, G. E.,
Rasche, N., Lu
¨der, F., Weckwerth, W. and Jahn, O. (2007). Proteome
analysis of Arabidopsis leaf peroxisomes reveals novel targeting peptides,
metabolic pathways, and defense mechanisms. The Plant Cell 19,
3170–3193.
Rohmer, M. (2003). Mevalonate-independent methylerythritol phosphate pathway
for isoprenoid biosynthesis. Elucidation and distribution. Pure and Applied
Chemistry 75, 375–387.
Sakuragi, Y. and Bryant, D. A. (2006). Genetic manipulation of quinone biosynthesis
in cyanobacteria. In Photosystem I: The Light-Driven Plastocyanin:Ferre-
doxin Oxidoreductase, (J. H. Golbeck, ed.), pp. 205–222. Springer,
Dordrecht.
Savage, D. and Lindenbaum, J. (1983). Clinical and experimental human vitamin K
deficiency. In Nutrition in Hematology, (J. Lindenbaum, ed.), pp. 271–320.
Churchill-Livingstone, New York.
Schoepp-Cothenet, B., Lieutaud, C., Baymann, F., Verme
´glio, A., Friedrich, T.,
Kramer, D. M. and Nitschke, W. (2009). Menaquinone as pool quinone in
a purple bacterium. Proceedings of the National Academy of Sciences of the
United States of America 106, 8549–8554.
Schønheyder, F. (1935). The anti-hemorrhagic vitamin of the chick. Measurement
and biological action. Nature 135, 653.
Schopfer, P., Heyno, E. and Krieger-Liszkay, A. (2008). Naphthoquinone-dependent
generation of superoxide radicals by quinone reductase isolated from the
plasma membrane of soybean. Plant Physiology 147, 864–878.
Schultz, G., Ellerbrock, B. and Soll, J. (1981). Site of prenylation reaction in synthesis
of phylloquinone (vitamin K
1
) by spinach chloroplasts. European Journal of
Biochemistry 117, 329–332.
Shevchuk, Y. M. and Conly, J. M. (1990). Antibiotic-associated hypoprothrombine-
mia: A review of prospective studies, 1966–1988. Reviews of Infectious
Diseases 12, 1109–1126.
Shimada, H., Ohno, R., Shibata, M., Ikegami, I., Onai, K., Ohto, M. A. and
Takamiya, K. (2005). Inactivation and deficiency of core proteins of photo-
systems I and II caused by genetical phylloquinone and plastoquinone
deficiency but retained lamellar structure in a T-DNA mutant of Arabidop-
sis. The Plant Journal 41, 627–637.
Sigfridsson, K., Hansson, O. and Brzezinski, P. (1995). Electrogenic light reactions in
photosystem I: Resolution of electron-transfer rates between the iron-sulfur
centers. Proceedings of the National Academy of Sciences of the United States
of America 92, 3458–3462.
Singh, A. K., Bhattacharyya-Pakrasi, M. and Pakrasi, H. B. (2008). Identification of
an atypical membrane protein in the formation of protein disulfide bonds in
oxygenic photosynthetic organisms. The Journal of Biological Chemistry
283, 15762–15770.
Strawn, M. A., Marr, S. K., Inoue, K., Inada, N., Zubieta, C. and Wildermuth, M. C.
(2007). Arabidopsis isochorismate synthase functional in pathogen-induced
salicylate biosynthesis exhibits properties consistent with a role in diverse
stress responses. The Journal of Biological Chemistry 282, 5919–5933.
Suttie, J. W. (1995). The importance of menaquinones in human nutrition. Annual
Review of Nutrition 15, 399–417.
Szulc, P., Chapuy, M. C., Meunier, P. J. and Delmas, P. D. (1996). Serum under-
carboxylated osteocalcin is a marker of the risk of hip fracture: A three year
follow-up study. Bone 18, 487–488.
260 CHLOE
¨VAN OOSTENDE ET AL.
Author's personal copy
Takahashi, Y. H., Inaba, K. and Ito, K. (2004). Characterization of the menaqui-
none-dependent disulfide bond formation pathway of Escherichia coli.The
Journal of Biological Chemistry 279, 47057–47065.
Thomas, G. and Threlfall, D. R. (1974). Incorporation of shikimate and 4-(20-carbox-
yphenyl)-4-oxobutyrate into phylloquinone. Phytochemistry 13, 807–813.
Truglio, J. J., Theis, K., Feng, Y., Gajda, R., Machutta, C., Tonge, P. J. and
Kisker, C. (2003). Crystal structure of Mycobacterium tuberculosis MenB,
a key enzyme in vitamin K2 biosynthesis. The Journal of Biological Chemis-
try 278, 42352–42360.
Unden, G. (1988). Differential roles for menaquinone and demethylmenaquinone in
anaerobic electron transport of E. coli and their fnr-independent expression.
Archives of Microbiology 150, 499–503.
van der Est, A. (2006). Electron transfer involving phylloquinone in photosystem I.
In Photosystem I: The Light-Driven Plastocyanin:Ferredoxin Oxidoreduc-
tase, (J. H. Golbeck, ed.), pp. 387–411. Springer, Dordrecht.
Verberne, M. C., Sansuk, K., Bol, J. F., Linthorst, H. J. M. and Verpoorte, R. (2007).
Vitamin K1 accumulation in tobacco plants overexpressing bacterial genes
involved in the biosynthesis of salicylic acid. Journal of Biotechnology 128,
72–79.
Vermeer, C., Shearer, M. J., Zittermann, A., Bolton-Smith, C., Szulc, P., Hodges, S.,
Walter, P., Rambeck, W., Stocklin, E. and Weber, P. (2004). Beyond
deficiency: Potential benefits of increased intakes of vitamin K for bone
and vascular health. European Journal of Nutrition 43, 325–335.
Widhalm, J. R., van Oostende, C., Furt, F. and Basset, G. J. (2009). A dedicated
thioesterase of the Hotdog-fold family is required for the biosynthesis of the
naphthoquinone ring of vitamin K1. Proceedings of the National Academy of
Sciences of the United States of America 106, 5599–5603.
Wildermuth, M. C., Dewdney, J., Wu, G. and Ausubel, F. M. (2001). Isochorismate
synthase is required to synthesize salicylic acid for plant defence. Nature 414,
562–565.
Yoshida, E., Nakamura, A. and Watanabe, T. (2003). Reversed-phase HPLC deter-
mination of chlorophyll a0and naphthoquinones in photosystem I of red
algae: Existence of two menaquinone-4 molecules in photosystem I of
Cyanidium caldarium.Analytical Sciences 19, 1001–1005.
Yu, G., Nguyen, T. T., Guo, Y., Schauvinhold, I., Auldridge, M. E., Bhuiyan, N.,
Ben-Israel, I., Iijima, Y., Fridman, E., Noel, J. P. and Pichersky, E. (2010).
Enzymatic functions of wild tomato methylketone synthases 1 and 2. Plant
Physiology 154, 67–77.
VITAMIN K
1
(PHYLLOQUINONE) 261
Author's personal copy
... It has been reported that foreign gene acquisitions by HGT in plastid genomes are rare compared to that of mitochondrial and nuclear genomes (Khan et al., 2007;Rice & Palmer, 2006). Menaquinone (men) genes encode functions involved in the conversion of chorismate into menaquinone (Van Oostende et al., 2011). The ...
... TAXONOMIC REVISION OF CYANIDIOPHYCEAE men gene cluster is widely distributed in bacteria, plastids of the Cyanidiophyceae, and the nuclear genomes of Viridiplantae and diatoms (Gross et al., 2008). As an example, based on the presence of menD and menF genes with comprehensive phylogenetic analysis, the Chlaymidial origin through HGT was suggested for the Cyanidiophyceae (Cenci et al., 2018;Gross et al., 2008;Van Oostende et al., 2011). Meanwhile, we confirmed that these HGT-derived genes (i.e., men, hupA) were positioned among neighboring plastid genes, which were continuously mapped by long-read Pacbio and short-read Illumina data in the plastid genomes of Cyanidium sp. ...
Article
Full-text available
The Cyanidiophyceae, a extremophilic red algal class, is distributed worldwide in extreme environments. Species grow either in acidic hot environments or in dim light conditions (e.g., “cave Cyanidium”). The taxonomy and classification systems are currently based on morphological, eco‐physiological, and molecular phylogenetic characters; however, previous phylogenetic results showed hidden diversity of the Cyanidiophyceae and suggested a revision of the classification system. To clarify phylogenetic relationships within this red algal class, we employ a phylogenomic approach based on 15 plastomes (ten new) and 15 mitogenomes (seven new). Our phylogenies show a consistent relationship among four lineages (Galdieria, “cave Cyanidium”, Cyanidium, and Cyanidioschyzon lineages). Each lineage is distinguished by organelle genome characteristics. The “cave Cyanidium” lineage is a distinct clade that diverged after the Galdieria clade but within a larger monophyletic clade that included the Cyanidium and Cyanidioschyzon lineages. Because the “cave Cyanidium” lineage is a mesophilic lineage that differs substantially from the other three thermoacidophilic lineages, we describe it as a new order (Cavernulicolales). Based on this evidence, we reclassified the Cyanidiophyceae into four orders: Cyanidiales, Cyanidioschyzonales, Cavernulicolales ord. nov., and Galdieriales ord. nov. The genetic distance between these four orders is comparable to, or greater than, the distances found between other red algal orders and subclasses. Three new genera (Cavernulicola, Gronococcus, Sciadococcus), five new species (Galdieria javensis, Galdieria phlegrea, Galdieria yellowstonensis, Gronococcus sybilensis, Sciadococcus taiwanensis) and a new nomenclatural combination (Cavernulicola chilensis) are proposed.
... Garcion et al. 7 previously found the Arabidopsis double mutant ics1 ics2 to be completely deficient of phylloquinone, which is essential for electron transfer in photosystem I. Isochorismate is an early precursor for phylloquinone, synthesized via 1,4-dihydroxy-2-naphthoic acid (NA) in a multistep pathway. 18,19 Thus, we were able to rescue the barley ics plants to almost wild-type size by spraying with NA ( Figure 1). ICS is located on barley chromosome 5, whereas the genes, encoding the enzymes mediating the biosynthetic steps between isochorismate and NA, 19 are located on barley chromosomes 3, 4, 6 and 7. Thus, the rescuing by NA, together with the co-segregation of ics and wilting, confirms that the ics mutation prevents isochorismate synthesis. ...
... 18,19 Thus, we were able to rescue the barley ics plants to almost wild-type size by spraying with NA ( Figure 1). ICS is located on barley chromosome 5, whereas the genes, encoding the enzymes mediating the biosynthetic steps between isochorismate and NA, 19 are located on barley chromosomes 3, 4, 6 and 7. Thus, the rescuing by NA, together with the co-segregation of ics and wilting, confirms that the ics mutation prevents isochorismate synthesis. ...
Article
Full-text available
Salicylic acid (SA) is an important signaling hormone in plant immunity. It can be synthesized by either the phenylpropanoid pathway or the isochorismate pathway, but mutant studies of this have been scarce in other species than Arabidopsis. Here we identified a mutation that introduced a stop-codon early in the barley gene for isochorismate synthase (ICS). We found that homozygous ics plants wilted if not sprayed with 1,4-dihydroxy-2-naphthoic acid, a precursor of phylloquinone, also synthesized via the isochorismate pathway. Interestingly, ics had unchanged SA, suggesting that the basal level of SA is synthesized via the phenylpropanoid pathway. Previous studies have failed seeing increased SA levels in barley after attack by the powdery mildew fungus, Blumeria graminis f.sp. hordei (Bgh), and indeed, we saw no changes in the interaction of ics with this fungus. Overall, we hope this mutant will be useful for other studies of SA in barley.
... ClCG08G017810 is an ortholog of the MenG gene in Arabidopsis thaliana, which encodes a 2-phytyl-1,4-beta-naphthoquinone methyltransferase protein involved in phylloquinone (vitamin K1) biosynthesis [35]. Phylloquinone is synthesized in plants, green algae, and some cyanobacteria and acts as a major electron transporter in photosystem I [36]. ...
Article
Full-text available
Watermelon fruit rind color (RC) and bloom formation (BF) affect product value and consumer preference. However, information on the candidate gene(s) for additional loci involved in dark green (DG) RC and the genetic control of BF and its major chemical components is lacking. Therefore, this study aimed to identify loci controlling RC and BF using QTL-seq of the F2 population derived by crossing ‘FD061129’ with light-green rind and bloom and ‘SIT55616RN’ with DG rind and bloomless. Phenotypic evaluation of the F1 and 219 F2 plants indicated the genetic control of two complementary dominant loci, G1 and G2, for DG and a dominant locus, Bf, for BF. QTL-seq identified a genomic region on Chr.6 for G1, Chr.8 for G2, and Chr.1 for Bf. G1 and G2 helped determine RC with possible environmental effects. Chlorophyll a-b binding protein gene-based CAPS (RC-m5) at G1 matched the highest with the RC phenotype. In the 1.4 cM Bf map interval, two additional gene-based CAPS markers were designed, and the CAPS for a nonsynonymous SNP in Cla97C01G020050, encoding a CSC1-like protein, cosegregated with the BF trait in 219 F2 plants. Bloom powder showed a high Ca2+ concentration (16,358 mg·kg−1), indicating that the CSC1-like protein gene is possibly responsible for BF. Our findings provide valuable information for marker-assisted selection for RC and BF and insights into the functional characterization of genes governing these watermelon-fruit-related traits.
... Vitamin K has the 2-methyl-1,4-naphthoquinone ring as a core structure, and the side-chain at the C3-position of the naphthoquinone ring structure ( Figure 1A) [5]. Phylloquinone has a partially unsaturated side chain consisting of one isopentenyl followed by three isopentyl units, while menaquinones have a fully unsaturated side chain composed of 2 to 13 isopentenyl units [7]. Menaquinone has various subtypes that have a different number (n) of isoprenoid units called menaquinone-n (MKn) [8]. ...
Article
Full-text available
Vitamin K is a fat-soluble vitamin that mainly exists as phylloquinone or menaquinone in nature. Vitamin K plays an important role in blood clotting and bone health in humans. For use as a nutraceutical, vitamin K is produced by natural extraction, chemical synthesis, and microbial fermentation. Natural extraction and chemical synthesis methods for vitamin K production have limitations, such as low yield of products and environmental concerns. Microbial fermentation is a more sustainable process for industrial production of natural vitamin K than two other methods. Recent advanced genetic technology facilitates industrial production of vitamin K by increasing the yield and productivity of microbial host strains. This review covers (i) general information about vitamin K and microbial host, (ii) current titers of vitamin K produced by wild-type microorganisms, and (iii) vitamin K production by engineered microorganisms, including the details of strain engineering strategies. Finally, current limitations and future directions for microbial production of vitamin K are also discussed.
... In I. balsamina, only phylloquinone and lawsone are directly derived from DHNA ( Fig. 1), with lawsone being the precursor of MNQ 32 . Three enzymatic reactions are required to convert DHNA into phylloquinone and this pathway has been fully characterized due to the latter's importance as an electron carrier in photosystem I (PSI) during photosynthesis 44 . However, the enzymes for specialized 1,4-naphthoquinones biosynthesis downstream of DHNA have not been identified, including that for MNQ biosynthesis. ...
Article
Full-text available
Impatiens balsamina L. is a tropical ornamental and traditional medicinal herb rich in natural compounds, especially 2-methoxy-1,4-naphthoquinone (MNQ) which is a bioactive compound with tested anticancer activities. Characterization of key genes involved in the shikimate and 1,4-dihydroxy-2-naphthoate (DHNA) pathways responsible for MNQ biosynthesis and their expression profiles in I. balsamina will facilitate adoption of genetic/metabolic engineering or synthetic biology approaches to further increase production for pre-commercialization. In this study, HPLC analysis showed that MNQ was present in significantly higher quantities in the capsule pericarps throughout three developmental stages (early-, mature- and postbreaker stages) whilst its immediate precursor, 2-hydroxy-1,4-naphthoquinone (lawsone) was mainly detected in mature leaves. Transcriptomes of I. balsamina derived from leaf, flower, and three capsule developmental stages were generated, totalling 59.643 Gb of raw reads that were assembled into 94,659 unigenes (595,828 transcripts). A total of 73.96% of unigenes were functionally annotated against seven public databases and 50,786 differentially expressed genes (DEGs) were identified. Expression profiles of 20 selected genes from four major secondary metabolism pathways were studied and validated using qRT-PCR method. Majority of the DHNA pathway genes were found to be significantly upregulated in early stage capsule compared to flower and leaf, suggesting tissue-specific synthesis of MNQ. Correlation analysis identified 11 candidate unigenes related to three enzymes (NADH-quinone oxidoreductase, UDP-glycosyltransferases and S-adenosylmethionine-dependent O-methyltransferase) important in the final steps of MNQ biosynthesis based on genes expression profiles consistent with MNQ content. This study provides the first molecular insight into the dynamics of MNQ biosynthesis and accumulation across different tissues of I. balsamina and serves as a valuable resource to facilitate further manipulation to increase production of MNQ.
... Both PhQs are active in PSI electron transport, but one in each pair appears to exhibit greater and PhQ (2-methyl-3-phytyl-1,4-naphthoquinone) is a conjugated isoprenoid made up of a redox-active naphthoquinone ring attached to a partially saturated C-20 phytyl side chain. PhQ biosynthesis comprises two separate metabolic branches: one for the naphthoquinone ring (PhQ head group) and the other for the phytyl side chain (Van Oostende et al., 2011). In Arabidopsis, chorismate serves as the precursor for the formation of the naphthoquinone ring and is converted into 1,4-dihydroxy-2naphthoate (DHNA) by a series of enzymatic reactions (Gross et al., 2006;Kim et al., 2008;Widhalm et al., 2012;Reumann, 2013) (Supplemental Figure 1). ...
Article
Full-text available
Phytyl-diphosphate, providing phytyl moieties as a common substrate in both tocopherol and phylloquinone biosynthesis, derives from de novo isoprenoid biosynthesis or a salvage pathway via phytol phosphorylation. However, very little is known about the role and origin of the phytyl moiety for phylloquinone biosynthesis. Since VTE6, a phytyl-phosphate kinase, is a key enzyme for phytol phosphorylation, we characterized Arabidopsis vte6 mutants to gain insight into the roles of phytyl moieties in phylloquinone biosynthesis and of phylloquinone in photosystem I (PSI) biogenesis. The VTE6 knock-out mutants vte6-1 and vte6-2 lacked detectable phylloquinone, whereas the phylloquinone content in the VTE6 knock-down mutant vte6-3 was 90% lower than that in wild-type. In vte6 mutants, PSI function was impaired and accumulation of the PSI complex was defected. The PSI core subunits PsaA/B were efficiently synthesized and assembled into the PSI complex in vte6-3. However, the degradation rate of PSI subunits in the assembled PSI complex was more rapid in vte6-3 than in wild-type. In vte6-3, PSI was more susceptible to high-light damage than in wild-type. Our results provide the first genetic evidence that the phytol phosphorylation pathway is essential for phylloquinone biosynthesis, and that phylloquinone is required for PSI complex stability.
Article
Full-text available
Vitamin K (VK) has long been known for its essential role in blood coagulation. However, over the past decade, evidence has mounted for its intrinsic and essential roles in other functions within the body, including bone metabolism, calcification, brain development and glucose metabolism. Thus, VK should no longer be considered a single-function ‘haemostasis vitamin’, but rather as a ‘multi-function vitamin’. While current research has focused on its emerging role in human nutrition, the role that VK plays in other species such as the horse has not been well described, with most of our current understanding having been extrapolated from other species, especially rodents. This review assesses the current state of knowledge of VK as it pertains to human and animal nutrition, and, where data exist, its metabolism and nutrition in the horse is explored. Future research on the roles of VK as they pertain to horses, particularly extra-hepatic functions, is necessary. Such insight will allow a greater understanding of how VK is metabolised, facilitating the development of recommendations to assist in the health, growth, and longevity of horses.
Article
Full-text available
Plant 1,4-naphthoquinones encompass a class of specialized metabolites known to mediate numerous plant–biotic interactions. This class of compounds also presents a remarkable case of convergent evolution. The 1,4-naphthoquinones are synthesized by species belonging to nearly 20 disparate orders spread throughout vascular plants, and their production occurs via one of four known biochemically distinct pathways. Recent developments from large-scale biology and genetic studies corroborate the existence of multiple pathways to synthesize plant 1,4-naphthoquinones and indicate that extraordinary events of metabolic innovation and links to respiratory and photosynthetic quinone metabolism probably contributed to their independent evolution. Moreover, because many 1,4-naphthoquinones are excreted into the rhizosphere and they are highly reactive in biological systems, plants that synthesize these compounds also needed to independently evolve strategies to deploy them and to resist their effects. In this review, we highlight new progress made in understanding specialized 1,4-naphthoquinone biosynthesis and trafficking with a focus on how these discoveries have shed light on the convergent evolution and diversification of this class of compounds in plants. We also discuss how emerging themes in metabolism-based herbicide resistance may provide clues to mechanisms plants employ to tolerate allelopathic 1,4-naphthoquinones.
Article
Full-text available
Pale green lethal (PGL) is a recessive genetic disorder of apple (Malus) characterized by severe chlorophyll deficiency and seedling lethality. Following germination, seedlings cannot photosynthesize and die at the cotyledon stage. We previously reported that the genetic and biochemical basis of PGL is due to a loss-of-function mutation in a gene required for the biosynthesis of phylloquinone (vitamin K1), a molecule essential for photosynthesis. For the present study, we used Illumina high-throughput RNA sequencing to identify genes differentially regulated between wild-type and PGL cotyledons. Changes in the expression of chlorophyll-related genes alone cannot explain the reduced chlorophyll content of PGL seedlings. However, genes putatively responding to numerous stress-related conditions including carbohydrate starvation, water deficit, and senescence were differentially regulated. This pattern of transcript accumulation suggests PGL seedlings alter many physiological and metabolic processes such as sorbitol metabolism, osmoprotectant production, and abscisic acid activity. The functions of individual genes relating to specific stresses are discussed. These findings provide insight into possible mechanisms PGL seedlings employ during stress response. Pale green lethal disorder may be a useful model for studying abiotic stress and senescence in rosaceous fruit tree species.
Article
o-Succinylbenzoic acid (OSB) is an intermediate in the biosynthesis of shikimatederived anthraquinones. The cell free activation of o-succinylbenzoic acid in extracts of anthraquinone producing cells of Galium mollugo L. is demonstrated for the first time. This activation depends on the presence of ATP, coenzyme A and Mg(2+). The o-succinylbenzoic acid coenzyme A ester was identified by converting it to 1,4-dihydroxy-2-naphthoic acid by a bacterial enzyme, viz. naphthoatesynthase. It is thus demonstrated that the o-succinylbenzoic acid coenzyme A ester derived from bacteria and from Galium mollugo cells are identical.
Article
The biosynthesis of phylloquinone (vitamin K1) was examined using Capsicum fruit chloroplasts and chromoplasts (apparently phylloquinone free). In both cases, the synthesis of phylloquinone from α-naphthoquinone, dihydro-α-naphthoquinone, 1,4-dihydroxy-2-naphthoic acid (as precursors of the ring moiety) and (S)-adenosyl-L-methionine was achieved. In the presence of phytylpyrophosphate, the biosynthesis of phylloquinone in both organelles is particularly enhanced when 1,4-dihydroxy-2-naphthoic acid is used.
Article
The patterns of incorporation of d-[G-14C]shikimate and variously labelled 14C-4-(2′-carboxy-phenyl)-4-oxobutyrate into the naphthoquinone nucleus of phylloquinone by maize shoots have been investigated. The results show that (a) the alicyclic ring and C-7 of shikimate give rise to Ring A and either C-1 or C-4, and (b) the phenyl ring, 2′-carboxy and C-4, and C-2 and -3 of 4-(2′-carboxyphenyl)-4-oxobutyrate give rise to Ring A, C-1 and -4 and C-2 and -3. Radioactivity from α-[1-14C]naphthol, 1,4-[1,4-14C]naphthoquinone and [Me-14C]menadione is not incorporated into phylloquinone to any significant extent.
Article
Radioactivity from 4-(2′-carboxyphenyl)-4-oxobutyrate-[2-14C] and 4-(2′-carboxyphenyl)- 4-oxobutyrate-[3-14C] was incorporated into C-3 and C-2 respectively of phylloquinone in maize shoots. These results show that this substrate is incorporated in the same asymmetric manner into phylloquinone as it is into the bacterial menaquinones.
Article
The suggestion that the electron acceptor A1 in plant photosystem I (PSI) is a quinone molecule is tested by comparisons with the bacterial photosystem. The electron spin polarized (ESP) EPR signal due to the oxidized donor and reduced quinone acceptor (P 870+Q-) in iron-depleted bacterial reaction centers has similar spectral characteristics as the ESP EPR signal in PSI which is believed to be due to P 700+A 1-, the oxidized PSI donor and reduced A1. This is also true for better resolved spectra obtained at K-band (∼24 GHz). These same spectral characteristics can be simulated using a powder spectrum based on the known g-anisotropy of reduced quinones and with the same parameter set for Q- and A1-. The best resolution of the ESP EPR signal has been obtained for deuterated PSI particles at K-band. Simulation of the A1- contribution based on g-anisotropy yields the same parameters as for bacterial Q- (except for an overall shift in the anisotropic g-factors, which have previously been determined for Q-). These results provide evidence that A1 is a quinone molecule. The electron spin polarized signal of P700+ is part of the better resolved spectrum from the deuterated PSI particles. The nature of the P700+ ESP is not clear; however, it appears that it does not exhibit the polarization pattern required by mechanisms which have been used so far to explain the ESP in PSI.