Encyclopedia_of_Chemical_Biology_(Wiley,_2008).pdf

(86002 KB) Pobierz
``Alkaloid Biosynthesis''. In Wiley Encyclopedia of Chemical Biology
Alkaloid Biosynthesis
Sarah E. O’Connor, Department of Chemistry, Massachusetts Institute of
Technology, Cambridge, Massachusetts
Advanced Article
Article Contents
Biologic Background
Isoquinoline Alkaloids
doi: 10.1002/9780470048672.wecb004
Terpenoid Indole Alkaloids
Tropane Alkaloids
How nature synthesizes complex secondary metabolites, or natural
products, can be studied only by working within the disciplines of both
chemistry and biology. Alkaloids are a complex group of natural products
with diverse mechanisms of biosynthesis. This article highlights the
biosynthesis of four major classes of plant-derived alkaloids. Only plant
alkaloids for which significant genetic information has been obtained were
chosen for review. Isoquinoline alkaloid, terpenoid indole alkaloid, tropane
alkaloid, and purine alkaloid biosynthesis are described here. The article is
intended to provide an overview of the basic mechanism of biosynthesis for
selected members of each pathway. Manipulation of these pathways by
metabolic engineering is highlighted also.
Purine Alkaloids
Alkaloids are a highly diverse group of natural products re-
lated only by the presence of a basic nitrogen atom located
at some position in the molecule. Even among biosynthetically
related classes of alkaloids, the chemical structures are often
highly divergent. Although some classes of natural products
have a recognizable biochemical paradigm that is centrally ap-
plied throughout the pathway, for example, the “assembly line”
logic of polyketide biosynthesis (1), the biosynthetic pathways
of alkaloids are as diverse as the structures. It is difficult to
predict the biochemistry of a given alkaloid based solely on
precedent, which makes alkaloid biosynthesis a challenging, but
rewarding, area of study.
have been cloned successfully, and many more enzymes have
been purified from alkaloid-producing plants or cell lines (4–6).
Identification and study of the biosynthetic enzymes has a sig-
nificant impact on the understanding of the biochemistry of the
pathway. Furthermore, genetic information also can be used to
understand the complicated localization patterns and regulation
of plant pathways. This article focuses on the biochemistry re-
sponsible for the construction of plant alkaloids and summarizes
the biosynthetic genes that have been identified to date. Some of
these pathways have been the subject of metabolic engineering
studies; the results of these studies are mentioned here also. An
excellent, more detailed review that covers the biochemistry and
genetics of plant alkaloid biosynthesis up until the late 1990s is
available also (7).
Biologic Background
Hundreds of alkaloid biosynthetic pathways have been studied
by chemical strategies, such as isotopic labeling experiments
(2, 3). However, modern molecular biology and genetic method-
ologies have facilitated the identification of alkaloid biosyn-
thetic enzymes. This article focuses on pathways for which a
significant amount of genetic and enzymatic information has
been obtained. Although alkaloid natural products are produced
by insects, plants, fungi, and bacteria, this article focuses on
four major classes of plant alkaloids: the isoquinoline alkaloids,
the terpenoid indole alkaloids, the tropane alkaloids, and the
purine alkaloids.
In general, plant biosynthetic pathways are understood poorly
when compared with prokaryotic and fungal metabolic path-
ways. A major reason for this poor understanding is that genes
that express complete plant pathways typically are not clus-
tered together on the genome. Therefore, each plant enzyme
often is isolated individually and cloned independently. How-
ever, several enzymes involved in plant alkaloid biosynthesis
Isoquinoline Alkaloids
The isoquinoline alkaloids include the analgesics morphine
and codeine as well as the antibiotic berberine ( Fig. 1a ).
Morphine and codeine are two of the most important analgesics
used in medicine, and plants remain the main commercial
source of the alkaloids (8). Development of plant cell cultures
of Eschscholzia californica , Papaver somniferum ,and Coptis
japonica has aided in the isolation and cloning of many enzymes
involved in the biosynthesis of isoquinoline alkaloids (9).
Early steps of isoquinoline biosynthesis
Isoquinoline biosynthesis begins with the substrates dopamine
and p -hydroxyphenylacetaldehyde ( Fig. 1b ). Dopamine is made
from tyrosine by hydroxylation and decarboxylation. Enzymes
that catalyze the hydroxylation and decarboxylation steps in
either order exist in the plant, and the predominant pathway
WILEY ENCYCLOPEDIA OF CHEMICAL BIOLOGY © 2008, John Wiley & Sons, Inc.
1
308553743.020.png 308553743.021.png 308553743.022.png 308553743.023.png 308553743.001.png 308553743.002.png 308553743.003.png
Alkaloid Biosynthesis
(a)
(b)
Figure 1 (a) Representative isoquinoline alkaloids. (b) Early biosynthetic steps of the isoquinoline pathway yield the biosynthetic intermediate
(S)-reticuline, the central biosynthetic intermediate for all isoquinoline alkaloids. (c) Berberine and sanguinarine biosynthesis pathways. (d)Morphine
biosynthesis. NCS, norcoclaurine synthase; 6-OMT, norcoclaurine 6-O-methyltransferase; CNMT, coclaurine N-methyltransferase (Cyp80B); NMTC,
N-methylcoclaurine 3 -hydroxylase; 4 -OMT, 3 -hydroxy-N-methylcoclaurine 4 -O-methyltransferase; BBE, berberine bridge enzyme; SOMT, scoulerine
9-O-methyltransferase; CS, canadine synthase; TBO, tetrahydroprotoberberine oxidase; CHS, cheilanthifoline synthase; SYS, stylopine synthase; NMT,
N-methyltransferase; NMSH, N-methylstylopine hydroxylase; P6H protopine 6-hydroxylase; DHPO, dihydrobenzophenanthridine oxidase; RO, reticuline
oxidase; DHR, dihydroreticulinium ion reductase; STS, salutaridine synthase; SalR, salutaridine reductase; SalAT, salutaridinol acetyltransferase; COR,
codeinone reductase.
for formation of dopamine from tyrosine is not clear. The
second substrate, p -hydroxyphenylacetaldehyde, is generated by
transamination and decarboxylation of tyrosine (10, 11).
Condensation of dopamine and p -hydroxyphenylacetaldehyde
is catalyzed by norcoclaurine synthase to form (S)-norcoclaurine
( Fig. 1b ). Two norcoclaurine synthases with completely unre-
lated sequences were cloned ( Thalictrum flavum and C. japon-
ica ) and heterologously expressed in E. coli (12–14). One is ho-
mologous to iron-dependent diooxygenases, whereas the other
is homologous to a pathogenesis-related protein. Undoubtedly,
future experiments will shed light on the mechanism of these
enzymes and on how two such widely divergent sequences can
catalyze the same reaction.
One of the hydroxyl groups of (S)-norcoclaurine is methy-
lated by a S-adenosyl methionine-(SAM)-dependent O-methyl
transferase to yield (S)-coclaurine. This enzyme has been
cloned, and the heterologously expressed enzyme exhibited
the expected activity (15–17). The resulting intermediate is
2
WILEY ENCYCLOPEDIA OF CHEMICAL BIOLOGY © 2008, John Wiley & Sons, Inc.
308553743.004.png 308553743.005.png 308553743.006.png 308553743.007.png
Alkaloid Biosynthesis
(c)
(d)
Figure 1 ( continued )
WILEY ENCYCLOPEDIA OF CHEMICAL BIOLOGY © 2008, John Wiley & Sons, Inc.
3
308553743.008.png 308553743.009.png 308553743.010.png 308553743.011.png
Alkaloid Biosynthesis
then N-methylated to yield N-methylcoclaurine, an enzyme
that has been cloned recently (18, 19). N-methylcoclaurine, in
turn, is hydroxylated by a P450-dependent enzyme (CYP80B),
N-methylcoclaurine 3 -hydroxylase, that has been cloned (20,
21). The 4 hydroxyl group then is methylated by the
enzyme 3 -hydroxy-N-methylcoclaurine 4 -O-methyltransferase
(4 -OMT) to yield (S)-reticuline, the common biosynthetic inter-
mediate for the berberine, benzo(c)phenanthridine, and morphi-
nan alkaloids ( Fig. 1b ). The gene for this methyl transferase also
has been identified (15, 22). These gene sequences also were
used to identify the corresponding T. flavum genes that encode
the biosynthetic enzymes for reticuline from a cDNA library
(23). At this point, the biosynthetic pathway then branches to
yield the different structural classes of isoquinoline alkaloids.
P450 enzyme, protopine 6-hydroxylase, to yield an intermediate
that rearranges to dihydrosanguinarine (40). This intermediate
also serves as the precursor to the benzo(c)phenanthridine alka-
loid macarpine ( Fig. 1a ). The copper-dependent oxidase dihy-
drobenzophenanthridine oxidase, which has been purified (41,
42), then catalyzes the formation of sanguinarine from dihy-
drosanguinarine.
Additional enzymes from other benzo(c)phenanthidine alka-
loids have been cloned. For example, an O-methyl transferase
implicated in palmitine biosynthesis has been cloned recently
(43).
Morphine biosynthesis
Berberine biosynthesis
The later steps of morphine biosynthesis have been inves-
tigated in P. somniferum cells and tissue. Notably, in mor-
phine biosynthesis, (S)-reticuline is converted to (R)-reticuline,
thereby epimerizing the stereocenter generated by norcoclaurine
synthase at the start of the pathway ( Fig. 1d ). (S)-reticuline is
converted to (R)-reticuline through a 1,2-dehydroreticuline in-
termediate. Dehydroreticuline synthase catalyzes the oxidation
of (S)-reticuline to 1,2-dehydroreticulinium ion (44). This en-
zyme has not been cloned but has been purified partially and
shown to be membrane-associated. This intermediate then is
reduced by dehydroreticuline reductase, an NADPH-dependent
enzyme that stereoselectively transfers a hydride to dehydroreti-
culinium ion to yield (R)-reticuline. This enzyme has not been
cloned yet but has been purified to homogeneity (45).
Next, the key carbon–carbon bond of the morphinan alka-
loids is formed by the cytochrome P450 enzyme salutaridine
synthase. Activity for this enzyme has been detected in microso-
mal preparations, but the sequence has not been identified (46).
The keto moiety of the resulting product, salutaridine, then is
stereoselectively reduced by the NADPH-dependent salutaridine
reductase to form salutardinol. The enzyme has been purified
(47), and a recent transcript analysis profile of P. sominiferum
has resulted in the identification of the clone (48). Salutaridinol
acetyltransferase, also cloned, then transfers an acyl group from
acetyl-CoA to the newly formed hydroxyl group, which results
in the formation of salutaridinol-7-O-acetate (49). This modifi-
cation sets up the molecule to undergo a spontaneous reaction
in which the acetate can act as a leaving group. The resulting
product, thebaine, then is demethylated by an as yet uncharac-
terized enzyme to yield neopinione, which exists in equilibrium
with its tautomer codeinone. The NADPH-dependent codeinone
reductase catalyzes the reduction of codeinone to codeine and
has been cloned (50, 51). Finally, codeine is demethylated by
an uncharacterized enzyme to yield morphine.
The localization of isoquinoline biosynthesis has been in-
vestigated at the cellular level in intact poppy plants by using
in situ RNA hybridization and immunoflouresence microscopy.
The localization of 4 -OMT (reticuline biosynthesis), berberine
bridge enzyme (saguinarine biosynthesis), salutaridinol acetyl-
transferase (morphine biosynthesis), and codeinone reductase
(morphine biosynthesis) has been probed. 4 -OMT and salu-
taridinol acetyltransferase are localized to parenchyma cells,
whereas codeinone reductase is localized to laticifer cells in
(S)-reticuline is converted to (S)-scoulerine by the action
of a well-characterized flavin-dependent enzyme, berberine
bridge enzyme ( Fig. 1c ). This enzyme has been cloned from
several plant species, and the mechanism of this enzyme
has been studied extensively (24–28). (S)-scolerine is then
O-methylated by scoulerine 9-O-methyltransferase to yield
(S)-tetrahydrocolumbamine. Heterologous expression of this
gene in E. coli yielded an enzyme that had the expected
substrate specificity (29). A variety of O-methyl transferases
also have been cloned from Thalictrum tuberosum (30). The
substrate-specific cytochrome P450 oxidase canadine synthase
(31) that generates the methylene dioxy bridge of (S)-canadine
has been cloned recently (32). The final step of berberine
biosynthesis is catalyzed by a substrate-specific oxidase, tetrahy-
droprotoberberine oxidase, the sequence of which has not been
identified yet (33).
Overproduction of berberine in C. japonica cell suspension
cultures was achieved by selection of a high-producing cell
line (34) with reported productivity of berberine reaching 7 g/L
(35). This overproduction is one of the first demonstrations of
production of a benzylisoquinoline alkaloid in cell culture at
levels necessary for economic production. This cell line has
facilitated greatly the identification of the biosynthetic enzymes.
Sanginarine biosynthesis
The biosynthesis of the highly oxidized benzo(c)phenanthidine
alkaloid sanguinarine is produced in a variety of plants and
competes with morphine production in opium poppy. The path-
way to sanguinarine has been elucidated at the enzymatic
level ( Fig. 1c ) (36). Sanguinarine biosynthesis starts from
(S)-scoulerine, as in berberine biosynthesis. Methylenedioxy
bridge formation then is catalyzed by the P450 cheilanthifo-
line synthase to yield cheilanthifoline (37). A second P450
enzyme, stylopine synthase, catalyzes the formation of the sec-
ond methyenedioxy bridge of stylopine (37). Stylopine syn-
thase from E. californica has been cloned recently (38). Sty-
lopine then is N-methylated by (S)-tetrahydroprotoberberine
cis -N-methyltransferase to yield (S)- cis -N-methylstylopine, an
enzyme that has been cloned recently from opium poppy (39).
A third P450 enzyme, (S)- cis -N-methylstylopine hydroxylase,
then forms protopine. Protopine is hydroxylated by a fourth
4
WILEY ENCYCLOPEDIA OF CHEMICAL BIOLOGY © 2008, John Wiley & Sons, Inc.
308553743.012.png 308553743.013.png 308553743.014.png 308553743.015.png
Alkaloid Biosynthesis
sections of capsule (fruit) and stem from poppy plants. Berber-
ine bridge enzyme is found in parenchyma cells in roots. There-
fore, this study suggests that two cell types are involved in iso-
quinoline biosynthesis in poppy and that intercellular transport
is required for isoquinoline alkaloid biosynthesis (52). Another
study, however, implicates a single cell type (sieve elements and
their companion cells) in isoquinoline alkaloid biosynthesis (53,
54). Therefore, it is not clear whether transport of pathway in-
termediates is required for alkaloid biosynthesis or whether the
entire pathway can be performed in one cell type. Localization
of enzymes in alkaloid biosynthesis is difficult, and, undoubt-
edly, future studies will provide more insight into the trafficking
involved in plant secondary metabolism.
Early steps of terpenoid indole alkaloid
biosynthesis
All terpenoid indole alkaloids are derived from tryptophan and
the iridoid terpene secologanin ( Fig. 2b ). Tryptophan decar-
boxylase, a pyridoxal-dependent enzyme, converts tryptophan
to tryptamine (62, 63). The enzyme strictosidine synthase cat-
alyzes a stereoselective Pictet–Spengler condensation between
tryptamine and secologanin to yield strictosidine. Strictosidine
synthase (64) has been cloned from the plants C. roseus (65),
Rauwolfia serpentine (66), and, recently, Ophiorrhiza pumila
(67). A crystal structure of strictosidine synthase from R. ser-
pentina has been reported (68, 69), and the substrate specificity
of the enzyme can be modulated (70).
Strictosidine then is deglycosylated by a dedicated
β -glucosidase, which converts it to a reactive hemiacetal inter-
mediate (71–73). This hemiacetal opens to form a dialdehyde
intermediate, which then forms dehydrogeissoschizine. The enol
form of dehydrogeissoschizine undergoes 1,4 conjugate addition
to produce the heteroyohimbine cathenamine (74–76). A variety
of rearrangements subsequently act on deglycosylated strictosi-
dine to yield a diversity of indole alkaloid products (77).
Metabolic engineering of morphine
biosynthesis
In attempts to accumulate thebaine and decrease produc-
tion of morphine (a precursor to the recreational drug hero-
ine), codeinone reductase in opium poppy plant was down-
regulated by using RNAi (8). Silencing of codeinone re-
ductase results in the accumulation of (S)-reticuline but not
the substrate codeinone or other compounds on the pathway
from (S)-reticuline to codeine. However, the overexpression
of codeinone reductase in opium poppy plants did result, in
fact, in an increase in morphine and other morphinan alkaloids,
such as morphine, codeine, and thebaine, compared with control
plants (55). Gene expression levels in low morphine-producing
poppy plants have been analyzed also (56). Silencing of berber-
ine bridge enzyme in opium poppy plants also resulted in a
change in alkaloid profile in the plant latex (57).
The cytochrome P450 responsible for the oxidation of
(S)-N-methylcoclaurine to (S)-3 -hydroxy-N-methylcocluarine
has been overexpressed in opium poppy plants, and morphi-
nan alkaloid production in the latex is increased subsequently
to 4.5 times the level in wild-type plants (58). Additionally,
suppression of this enzyme resulted in a decrease in morphi-
nan alkaloids to 16% of the wild-type level. Notably, analysis
of a variety of biosynthetic gene transcript levels in these ex-
periments supports the hypothesis that this P450 enzyme plays
a regulatory role in the biosynthesis of benzylisoquinoline al-
kaloids. Collectively, these studies highlight that the complex
metabolic networks found in plants are not redirected easily or
predictably in all cases.
Ajmaline biosynthesis
The biosynthetic pathway for ajmaline in R. serpentina is one
of the best-characterized terpenoid indole alkaloid pathways.
Much of this progress has been detailed in a recent extensive
review (78). Like all other terpenoid indole alkaloids, ajmaline,
an antiarrhythmic drug with potent sodium channel-blocking
properties (79), is derived from deglycosylated strictosidine
( Fig. 2c ).
A membrane–protein fraction of an R. serpentina extract
transforms labeled strictosidine (80, 81) into sarpagan-type alka-
loids. The enzyme activity is dependent on NADPH and molec-
ular oxygen, which suggests that sarpagan bridge enzyme may
be a cytochrome P450 enzyme. Polyneuridine aldehyde esterase
hydrolyzes the polyneuridine aldehyde methyl ester, which gen-
erates an acid that decarboxylates to yield epi-vellosamine.
This enzyme has been cloned from a Rauwolfia cDNA library,
heterologously expressed in E. coli , and subjected to detailed
mechanistic studies (82, 83).
In the next step of the ajmaline pathway, vinorine synthase
transforms the sarpagan alkaloid epi-vellosamine to the ajmalan
alkaloid vinorine (84). Vinorine synthase also has been purified
from Rauwolfia cell culture, subjected to protein sequencing,
and cloned from a cDNA library (85, 86). The enzyme, which
seems to be an acetyl transferase homolog, has been expressed
heterologously in E. coli . Crystallization and site-directed muta-
genesis studies of this protein have led to a proposed mechanism
(87).
Vinorine hydroxylase hydroxylates vinorine to form vom-
ilene (88). Vinorine hydroxylase seems to be a P450 en-
zyme that requires an NADPH-dependent reductase. This
enzyme is labile and has not been cloned yet. Next, the
indolenine bond is reduced by an NADPH-dependent re-
ductase to yield 1,2-dihydrovomilenene. A second enzyme,
1,2-dihydrovomilenene reductase, then reduces this product to
Terpenoid Indole Alkaloids
The terpenoid indole alkaloids have a variety of chemical struc-
tures and a wealth of biologic activities ( Fig. 2a ) (59, 60).
Terpenoid indole alkaloids are used as anticancer, antimalarial,
and antiarrhythmic agents. Although many biosynthetic genes
from this pathway remain unidentified, recent studies have cor-
related terpenoid indole alkaloid production with the transcript
profiles of Catharanthus roseus cell cultures (61).
WILEY ENCYCLOPEDIA OF CHEMICAL BIOLOGY © 2008, John Wiley & Sons, Inc.
5
308553743.016.png 308553743.017.png 308553743.018.png 308553743.019.png
Zgłoś jeśli naruszono regulamin