You are on page 1of 11

REVIEWS

FUNGAL SECONDARY
METABOLISM  FROM
BIOCHEMISTRY TO GENOMICS
Nancy P. Keller*, Geoffrey Turner‡ and Joan W. Bennett§
Abstract | Much of natural product chemistry concerns a group of compounds known as
secondary metabolites. These low-molecular-weight metabolites often have potent
physiological activities. Digitalis, morphine and quinine are plant secondary metabolites,
whereas penicillin, cephalosporin, ergotrate and the statins are equally well known fungal
secondary metabolites. Although chemically diverse, all secondary metabolites are produced
by a few common biosynthetic pathways, often in conjunction with morphological
development. Recent advances in molecular biology, bioinformatics and comparative
genomics have revealed that the genes encoding specific fungal secondary metabolites are
clustered and often located near telomeres. In this review, we address some important
questions, including which evolutionary pressures led to gene clustering, why closely related
species produce different profiles of secondary metabolites, and whether fungal genomics will
accelerate the discovery of new pharmacologically active natural products.

The fungal kingdom includes many species with chemists. Secondary metabolites are often bioactive,
unique and unusual biochemical pathways. The usually of low molecular weight, and are produced as
products of these pathways include important pharma- families of related compounds at restricted parts of
ceuticals such as penicillin, cyclosporin and statins; the life cycle, with production often correlated with
*University of Wisconsin– potent poisons, including aflatoxins and tricho- a specific stage of morphological differentiation.
Madison, Department of thecenes; and some Janus-faced metabolites that Secondary metabolites share the enigmatic properties
Plant Pathology, are both toxic and pharmaceutically useful, such of cellular dispensability — producer organisms can
882 Russell Labs, 1630 as the ERGOT ALKALOIDS. All of these natural products, grow without synthesizing these metabolites — and
Linden Drive, Madison,
Wisconsin 53706, USA. along with many other low-molecular-weight fun- restricted taxonomic distribution (only a small group

Department of Molecular gal metabolites, are classified together as secondary of organisms produces each metabolite)5,6.
Biology and Biotechnology, metabolites and have been reviewed elsewhere1–3. The systematic study of fungal secondary meta-
Firth Court, University of What are secondary metabolites? Julian Davies once bolites began in 1922 under the leadership of Harold
Sheffield, S10 2TN, UK.
§ joked that “the simplest definition is that they are not Raistrick, who eventually characterized more than
Department of Cell and
Molecular Biology, Tulane generally included in the standard metabolic charts”4. 200 mould metabolites7. However, it was not until
University, New Orleans, Secondary metabolites are customarily distinguished the discovery and development of penicillin BOX 1
Louisiana 70118, USA. from primary metabolites, which are the almost that widespread attention was focused on fungal
e-mails: universally distributed compounds of INTERMEDIARY metabolites. Pharmaceutical companies instigated
npk@plantpath.wisc.edu;
METABOLISM. Primary metabolism has been the domain extensive screening programmes, and by 1950 a
G.Turner@sheffield.ac.uk;
profmycogirl@yahoo.com of biochemists, whereas until recently, secondary treasure trove of microbial products with pharma-
doi:10.1038/nrmicro1286 metabolism has largely been the domain of organic ceutical applications had been discovered. This

NATURE REVIEWS | MICROBIOLOGY VOLUME 3 | DECEMBER 2005 | 937


© 2005 Nature Publishing Group
REVIEWS

Box 1 | Penicillin
Penicillin, the first broad-spectrum antibiotic, is the most famous fungal secondary metabolite. Penicillin
transformed the practice of medicine, changed the trajectory of pharmaceutical research, influenced the course
of World War II and saved countless lives. The penicillin story has been told many times95–98. In 1929, Alexander
Fleming discovered the antibacterial action of a ‘mould juice’ from Penicillium notatum and named the
biological activity ‘penicillin’. A decade later, Howard Florey, working at Oxford University with Ernst Chain,
Norman Heatley and others, purified enough penicillin to demonstrate its clinical efficacy, completing their
first experiments during the evacuation of Dunkirk during World War II. Florey and Heatley, fearing a German
invasion, travelled clandestinely to the USA with cultures of the mould. Penicillin yields were low and the
surface-culture method used for growing the mould was cumbersome, so government scientists in Peoria,
Illinois screened fungi from all over the world in search of a higher-yielding strain that could grow in
submerged culture. An ardent technician named Mary Hunt, soon dubbed ‘Mouldy Mary’, scoured Peoria
markets for mouldy produce, and found the rotting cantaloupe that eventually yielded a strain of Penicillium
chrysogenum that was selected for large-scale production99. As scale-up methods were perfected in preparation
for D-Day, chemists worked hard to elucidate the chemical structure of the antibiotic. It became apparent that
there was more than one penicillin. The main metabolite obtained by surface culture of the Fleming strain of
P. notatum was different from the main metabolite obtained by submerged fermentation of the Peoria strain of
P. chrysogenum. These metabolites are now known, respectively, as 2-pentenylpenicillin (‘penicillin I’) and
benzylpenicillin (‘penicillin II’)100.
It had been expected that a complete chemical synthesis for penicillin would replace fungal biosynthesis in
commercial production. Instead, efficient industrial-scale fermentation methods were developed. These
fermentation technologies were not only good preparation for industrial production of the streptomycin family of
antibiotics by actinomycetes, but have also provided a platform for the biotechnological production of mammalian
hormones and other gene-encoded products using genetically engineered microorganisms.

search for bioactive secondary metabolites has more than half of these molecules had antibacterial,
continued unabated, and thousands of compounds antifungal or antitumour activity8. Given the empha-
that inhibit the growth of bacteria, fungi, protozoa, sis on physiological activity, one common classifica-
parasites, insects, viruses and even human tumour tion method for secondary metabolites is through
cells have been discovered. Many other molecules defining their impact on human interests, for exam-
with cytotoxic, mutagenic, carcinogenic, teratogenic, ple, plant and animal toxins, growth hormones and
immunosuppressive, enzyme inhibitory, ALLELOPATHIC pharmaceuticals. Another, more chemically rational
and other biological effects also have been found. A system of classification reflects the fact that, despite
recent literature survey of fungal metabolites, which their enormous chemical complexity and diversity, all
examined 1,500 compounds that were isolated and secondary metabolites arise from a limited number
characterized between 1993 and 2001, showed that of precursors from primary metabolism. This review

Box 2 | Aflatoxins
Mycotoxins, or mould poisons, are less well known than mushroom poisons but cause a higher incidence of disease.
Eating toxic mushrooms is easier to avoid than the inadvertent consumption of mould-contaminated foods.
Synthetic contaminants, food additives and pesticide residues get more press attention, but mycotoxins are almost
certainly the main non-infectious dietary risk factor in the human food supply101. The most famous mycotoxins are
the aflatoxins. These molecules were discovered in the early 1960s when thousands of turkey poults mysteriously
ERGOT ALKALOID died in hatcheries in and around London. All of the dead turkeys had been fed the same Brazilian peanut meal. The
Any of a group of about 30 meal was heavily contaminated with a common species of mould, so suspicion focused on the fungus, and soon a
indole alkaloids obtained from family of toxic metabolites was isolated. These toxins were named aflatoxins after the producer species, Aspergillus
the sclerotial phase of the
flavus. The four major aflatoxins — aflatoxin B1, B2, G1 and G2 — were identified based on their blue or green
fungus Claviceps purpurea.
fluorescence under ultraviolet light and their relative chromatographic mobility during thin-layer chromatography
INTERMEDIARY METABOLISM with silica gel101. Further studies showed that aflatoxin B1 was one of the most toxic and carcinogenic compounds
Enzyme-catalysed processes ever discovered102,103. The ease with which A. flavus grew on most major crop plants and the prevalence of aflatoxin
within cells that metabolize contamination of foods and feeds led to major international research efforts that included elucidation of most of the
macronutrients, carbohydrate, genes in the biosynthetic pathway13,104,105.
fat and protein.
There is considerable evidence that the Iraqi government stockpiled aflatoxins as part of their chemical-warfare
ALLELOPATHIC
programme during the 1980s106. Perhaps Saddam Hussein, or one of his henchmen, was influenced by Graham
Describes secondary Greene’s 1979 spy novel, The Human Factor107. The plot revolves around the toxicologically improbable murder of a
metabolites released by plants, man whose whisky had been laced with aflatoxin. When the victim, a heavy drinker, succumbs to liver cancer, no
bacteria, fungi or viruses that one suspects foul play. Nonetheless, despite their rightly deserved notoriety, aflatoxins are a poor choice of poison
have a direct or indirect, for both novelists and bioterrorists. They do most of their damage in developing countries, where their prevalence
harmful or even beneficial
in the food supply is an all too common phenomenon108.
effect on another organism.

938 | DECEMBER 2005 | VOLUME 3 www.nature.com/reviews/micro


© 2005 Nature Publishing Group
REVIEWS

a Peptides Classes of fungal secondary metabolites


CH3 O
O O There are several classes of secondary metabolites,
N CH3
NH discussed below.
HN N
S HO
O N
H3C O
O Polyketides. Polyketides are the most abundant
N
H3C N fungal secondary metabolites. The genetically
O
Penicillin G O OH best-characterized polyketides include the yellow
O A. nidulans spore-pigment intermediate naphtho-
N CH3 N CH3 pyrone (WA)9, the carcinogen aflatoxin BOX 2 and
O
O
O
the commercially important cholesterol-lowering
CH3 compound lovastatin10 (FIG. 1). Fungal polyketides are
S N HN NH synthesized by type I polyketide synthases (PKSs),
N S OH CH3 O
O
which are multidomain proteins that are related to
N N
O O H eukaryotic fatty-acid synthases and contain similar
OH
DOMAIN structures (FIG. 2). For both enzymes, short-
Gliotoxin Cyclosporin chain carboxylic acids — usually acetyl coenzyme A
b Alkaloids (acetyl CoA) and malonyl CoA — are condensed to
2
H R O OH form carbon chains of varying lengths. The main dif-
O
O N
O
ference between polyketides and fatty acids is the full
N
N H3C N reduction of the β-carbon in fatty acids, which is an
O
O N
N optional event in polyketide synthesis. In the fungal
N R3 H PKSs, the ketoacyl CoA synthase (KS), acyltransferase
CH3 O
(AT) and acyl carrier (ACP) domains are essential for
polyketide synthesis, whereas the ketoreductase (KR),
HN
Ergopeptides Fumitremorgen C dehydratase (DH) and enoyl reductase (ER) domains
R1
that are required for ketone reduction in fatty acids are
c Terpenes
OH
not present in all fungal PKS enzymes (FIG. 2). Bacteria
O
O also have multidomain, type I PKS enzymes11, for
O O OH
example, the multimodular enzyme DEBS (6-deoxy-
O
CO erythronolide B synthase) is required for biosynthesis
O O
HO of the antibiotic erythromycin A. It consists of three
O
O OH different subunits, each of which is composed of two
O
MODULES. The six modules function successively in a
Aristolochene Trichothecene T2 toxin Gibberellin GA3
six-step reaction to incorporate methylmalonyl CoA
d Polyketides O into a growing polyketide chain, which is then released
HO O
O
and cyclized through addition of a thioesterase domain
O
O at the end of module six12.
HO CH3
Unlike bacterial PKS type I enzymes, which have
HN O
O O O separate modules for each methylmalonyl CoA addi-
Fusarin C tion, fungal PKSs are limited to one module, with
O O which they can carry out repeated biosynthetic reac-
O
tions, and are therefore called ‘iterative PKSs’. For
OH OH O Lovastatin example, the WA PKS, which has the minimal domain
O set of KS–AT–ACP, uses repeated cycles of condensa-
OH
O
CH3 tion without reduction to generate a heptaketide. The
O HO O
C-terminal region of the enzyme, which has a thio-
Aflatoxin B1 WA esterase domain motif, is responsible for a CLAISENTYPE
Figure 1 | The main groups of fungal secondary metabolites. Ergopeptides such as CYCLIZATION that results in the formation of the aromatic
ergotamine are tryptophan-derived alkaloids to which peptides are added by a non-ribosomal polyketide naphthopyrone9. How the number of cycles
peptide synthetase (NRPS). T2 toxin is a trichothecene of Fusarium sporotrichioides. WA of condensation is controlled to stop at the hexaketide
(a naphthopyrone), is a yellow pigment produced by the polyketide synthase WA, encoded by is not understood for this or for any other fungal type I
the wA gene of Aspergillus nidulans. It is converted to the green spore pigment by a laccase.
PKS enzyme.
Fusarin C synthesis is dependent on a hybrid PKS (polyketide synthase)–NRPS.
The diversity of fungal polyketide structures results
from the number of iteration reactions, the number of
will start by classifying secondary metabolites reduction reactions, which extender unit is used and,
DOMAIN according to the enzyme classes involved in their bio- in the case of aromatic polyketides, cyclizations of the
In a polyketide synthase or synthesis, and ends with a description of the wealth nascent polyketide chain. Further variety is achieved
non-ribosomal peptide of putative metabolites that have been revealed by by the introduction of many different post-polyketide-
synthetase, a stretch of
conserved amino acids that
sequencing the genomes of three Aspergillus spe- synthesis steps. For example, in addition to the PKS
defines a specific biochemical cies: Aspergillus nidulans, Aspergillus fumigatus and and a fatty-acid synthase that is required for a hexanoyl
function or active site region. Aspergillus oryzae. CoA starter unit, in A. nidulans the genes in the cluster

NATURE REVIEWS | MICROBIOLOGY VOLUME 3 | DECEMBER 2005 | 939


© 2005 Nature Publishing Group
REVIEWS

tethered amino acids. The resulting peptide is then


KS AT (DH) (MT) (ER) (KR) ACP (CYC) (TE) released by a thioesterase-like domain that is usually
located at the C-terminal end of the final module.
Figure 2 | Fungal polyketide synthase (PKS) domain structure. The minimal structure is The domains that carry out these activities are named
KS–AT–ACP. Optional domains are in brackets. Polyketide synthesis is initiated when acetyl A (adenylation), P (pantothenylation/peptidyl carrier),
and malonyl coenzyme A (CoA) are loaded as thioesters on to the 4′-phosphopantotheine of
an acyl carrier (ACP) domain by means of the acyltransferase (AT) domain. Condensation then
C (condensation/peptide-bond formation) and TE
occurs with another thioester intermediate bound to the ketoacyl CoA synthase (KS) domain, (thioesterase) (FIG. 3). The first fungal NRPS identified,
and decarboxylation of the ACP-bound intermediate occurs. The resulting β-ketothioester δ-(l-α-aminoadipyl)-l-cysteinyl-d-valine synthetase
can then be reduced by the action of the ketoreductase (KR) domain, followed by dehydration (ACVS), catalyses the first committed step in the bio-
by the dehydratase (DH) domain. If an enoyl reductase (ER) domain is present, an unsaturated synthesis of the β-lactam antibiotics penicillin and
intermediate is formed. Some PKSs contain a methyltransferase (MT) domain that methylates cephalosporin17 BOX 1. ACVS condenses l-α-amino-
the α-carbon of the thioester. CYC, cyclase; TE, thioesterase.
adipic acid, l-cysteine and l-valine, and epimerizes
l-valine to d-valine18 to form a linear peptide (ACV)
that is subsequently cyclized to isopenicillin N by the
MODULE that is required for sterigmatocystin synthesis encode action of isopenicillin N synthase.
In a polyketide synthase or five monoxygenases, four dehydrogenases, an esterase, Diversity among non-ribosomal peptides arises
non-ribosomal peptide an O-methyltransferase, a reductase, an oxidase and a in the length of peptides produced, whether the
synthetase, the complete set of DNA-binding protein13. Similar clusters of genes that peptide is cyclized, and variations in the functions
domains that is required for one
round of chain elongation and
code for PKS enzymes and 15 or more genes encoding of the domains. Another example of an NRPS that
modification. post-PKS-step enzymes are present in the clusters that produces a linear peptide is the peptaibol synthetase
are required for aflatoxin synthesis in Aspergillus flavus of Trichoderma virens19. This enzyme produces an
CLAISENTYPE CYCLIZATION and Aspergillus parasiticus14,15. 18-residue linear peptide that includes the rare amino
Claisen condensations are a
acid aminoisobutyric acid. The N terminus is acylated
common mechanism in
biological systems for synthesis Non-ribosomal peptides. Non-ribosomal peptides by an enzyme domain that resembles a fatty-acid
of carbon–carbon bonds. The are derived from both PROTEINOGENIC AMINO ACIDS and synthase, and the C-terminal amino acid is hydroxy-
product is a β-ketoester. A non-proteinogenic amino acids by multidomain, lated. In Tolypocladium niveum, an NRPS synthesizes
similar reaction is also used to multimodular enzymes named non-ribosomal pep- the cyclic undecapeptide cyclosporin20, which is an
cyclize the heptaketide product
of the wA gene to form an
tide synthetases16 (NRPSs) (FIGS 1,3). Each module immunosuppressive drug that is used to treat patients
aromatic ring. in an NRPS contains several domains that allow after organ-transplant surgery. This NRPS lacks a TE
recognition, activation and then covalent bind- domain. The peptide is released by cyclization, and
ing of a module-specific amino acid as a thioester some of the modules contain an additional methyl-
to the 4′-phosphopantetheine cofactor, which is ation domain between the A and P domains, which
attached to each module through a conserved serine. catalyses the methylation of amino acids. The enzyme
Subsequently, peptide bonds are formed between the modules and amino-acid sequences are collinear with
a repeating pattern (APC)n of domain structure. Other
NRPSs in bacteria and fungi have a more diverse
A P C A P C A P C TE domain arrangement. For example, in A. nidulans
the NRPS gene that is involved in the biosynthesis of
O S O S O S the siderophore ferricrocin, a cyclic hexapeptide with
the structure Gly–Ser–Gly–(N5-acetyl–N5-hydroxy-
H2N H2N
ornithine)3, has the domain structure (APC)3 (PC)2
REF. 21. It is not clear how the modules and domains
NH2 SH
L-cysteine L-valine
are used, although some iterative use of modules is
HO O
hypothesized.
L-α-aminoadipate
SH Terpenes. The best-known terpenes are odoriferous
H
H2N N plant metabolites such as camphor and turpentine,
O
but fungi also synthesize several important terpenes,
HO O O N
H
O including the aristolochenes, carotenoids, gibberellins,
HO
indole-diterpenes and trichothecenes. All terpenes are
δ-(L-α-aminoadipyl)-L-cysteinyl-D-valine composed of several isoprene units, can be linear or
cyclic, saturated or unsaturated, and can be modified
in various ways (FIGS 1,4).
β-lactam antibiotics Common classes of terpenes include the mono-
Figure 3 | ACV synthetase, a trimodular non-ribosomal peptide synthetase. terpenes, which are generated from geranyl pyro-
δ-(L-α-aminoadipyl) L-cysteinyl-D-valine (ACV) synthetase catalyses the first committed step in phosphate (also known as geranyl diphosphate);
penicillin and cephalosporin biosynthesis. Each amino acid is recognized and activated by the sesquiterpenes, which are generated from farnesyl
cognate adenylation domain (A), and attached as a thioester to 4′-phosphopantetheine at the
pyrophosphate (also known as farnesyl diphosphate);
peptidyl carrier domain (P). Peptide bonds are formed with the involvement of the condensation
domain (C). The final tripeptide, attached to the peptidyl carrier domain of the C-terminal and diterpenes and carotenoids, which are generated
module, is released by the integrated thioesterase domain (TE), with the L-valine isomerized to from geranylgeranyl pyrophosphate (also known as
D-valine. The tripeptide is subsequently cyclized to isopenicillin N. geranylgeranyl diphosphate). The defining enzyme in

940 | DECEMBER 2005 | VOLUME 3 www.nature.com/reviews/micro


© 2005 Nature Publishing Group
REVIEWS

Mevalonate pathway Carotenoid biosynthesis has been studied in


Neurospora crassa25. Only one geranylgeranyl diphos-
phate synthase has been found in the genome of
P P O N. crassa, in contrast to the two distinct genes encoding
P P
O geranylgeranyl diphosphate synthase that are found in
Dimethylallyl diphosphate Isopentenyl diphosphate gibberellin- and indole-diterpene-producing fungi.
Secondary metabolites Primary metabolites
Of the genes encoding these cyclases, one (named
ggs2 in G. fujikuroi and paxG in Penicillium paxilli) is
Indole alkaloids P P dedicated to secondary metabolic pathways for gib-
O
berellin and paxilline, respectively; this might reflect
Dimethylallyl diphosphate (DMAPP) compartmentalization of primary and secondary
O
P P
terpene metabolism26.

Indole alkaloids. Indole alkaloids are usually derived


Monoterpenes P P
O from tryptophan and dimethylallyl pyrophosphate,
Geranyl diphosphate (GPP) although sometimes amino acids other than tryptophan
O
are used as precursors BOX 3. The best-understood
P P
pathway is ergotamine synthesis in Claviceps purpurea
and related species27. The first committed step is the
Sesquiterpenes P P Steroids PRENYLATION of tryptophan by dimethylallyl tryptophan
O synthetase (DMATS). Following methylation of
Farnesyl diphosphate (FPP) dimethylallyl tryptophan, a series of oxidation steps
O proceed through agroclavine to lysergic acid. Lysergic
P P
acid is then activated by a single-module NRPS, con-
densed with a tripeptide that is produced by a second
Diterpenes P P Carotenoids NRPS, and released as ergotamine.
O
Geranylgeranyl diphosphate (GGPP) Other tryptophan-derived alkaloids such as the
fumigaclavines and fumitremorgens of A. fumigatus
undergo one or more prenylation steps. The details of
C50 Coenzyme Q these pathways are yet to be elucidated, but it is likely
Figure 4 | Terpene biosynthetic pathway. Isopentenyl diphosphate and its isomer dimethylallyl that the fumigaclavine biosynthetic pathway proceeds
diphosphate (DMAPP), products of the mevalonate pathway, are the building blocks (5C isoprene through agroclavine and might therefore have some
units) for the linear polyprenyl diphosphates, which are precursors of steroids, carotenoids and early steps in common with the ergotamine pathway.
coenzyme Q in many species. A family of isoprenyl diphosphate synthases is responsible for Fumitremorgens are derived from proline, tryptophan
chain elongation. DMAPP and the isoprenoid intermediates are also the starting points for a wide and dimethylallyl pyrophosphate28. Tryptophan and
range of secondary metabolites, including indole alkaloids, monoterpenes, sesquiterpenes and
diterpenes. The terpenes are produced by cyclization of the isoprenoids. For example, farnesyl
proline are condensed to form a diketopiperazine called
diphosphate is cyclized to produce a large variety of sesquiterpenes, depending on the cyclase brevianamide F, possibly by means of an unidentified
structure and function. DMAPP can be added as a side chain to various aromatic compounds. NRPS. It is hypothesized that brevianamides E and A
For example, ergotamine is derived from dimethylallyl tryptophan. from Penicillium brevicompactum and fumitremorgens
A, B and C, and tryprostatins from A. fumigatus are
derived from brevianamide F by diverse pathways
terpene synthesis is terpene cyclase, which is essential that require various oxidases, methylases and prenyl
for the production of different terpenes from different transferases.
diphosphates. Although terpene cyclases have struc-
tural homology, they have little primary sequence Genetics
similarity and seem to have diverged relatively rapidly Gene clusters. Until recently, research into fungal
from a common ancestor. secondary metabolism was dominated by drug com-
Several fungal terpene cyclases have been char- panies, which have screening programmes for new
acterized, including a bifunctional terpene cyclase and commercially viable bioactive products that are
from Gibberella fujikuroi22, a trichodiene synthase coupled with the structural elucidation of bioactive
from Fusarium sporotrichioides23 and an aristolo- secondary metabolites. The biochemical pathways
chene cyclase from Aspergillus terreus and Penicillium that lead to the production of a few specific second-
PROTEINOGENIC AMINO ACIDS roquefortii24. The crystal structures of the trichodiene ary metabolites, including aflatoxin, penicillins and
Those amino acids that are
synthase and aristolochene cyclase have been solved. ergot alkaloids, have however been analysed using
found in proteins and that are
coded for in the standard A conserved DDXXD/E motif (where X is any amino blocked mutants and isotopic tracers29–31. In general,
genetic code. Proteinogenic acid) is involved in binding Mg2+ ions, which are biochemical and genetic studies were rare owing to
means ‘protein-building’. required for the enzyme to bind to prenyl phosphates. the formidable technical challenges: the enzymes that
Trichothecene and aristolochene are the substrates synthesize secondary metabolites are only present in
PRENYLATION
The enzymatic addition of
for further modifications, including acetylations, minute quantities, and many of the producer fungal
prenyl moieties to secondary esterifications and oxygenations that generate various species lacked the sexual systems that were required
metabolic intermediates. sesquiterpenoid mycotoxins. for genetic analysis. The advent of recombinant DNA

NATURE REVIEWS | MICROBIOLOGY VOLUME 3 | DECEMBER 2005 | 941


© 2005 Nature Publishing Group
REVIEWS

Box 3 | Ergot alkaloids


Ergot alkaloids are found in the sclerotia of Claviceps, a genus of plant pathogens that parasitizes rye, wheat and
other grasses. When sclerotia are ground and baked into bread, inadvertent human consumption can lead to
convulsions, vasoconstriction and possible hallucinations, a disease syndrome variously called St Anthony’s Fire or
ergotism. Claviceps sclerotia contain a cocktail of ergot alkaloids from which modern scientists have identified three
main groups: the clavine type, the lysergic-acid type and the peptide alkaloids. Members of each group have
different physiological properties, some of which have been exploited for human use. For example,
ethnomycologists have hypothesized that the Eleusian mysteries, holy rituals carried out in ancient Greece, used an
elixir containing water-soluble, hallucinogenic ergot alkaloids. Medieval midwives adopted Claviceps sclerotia to
hasten labour and to induce abortion. In the twentieth and twenty-first centuries, ergot alkaloids have been used in
obstetrics as well as in the treatment of migraine headaches109.
The famous hallucinogen lysergic acid diethyl amide (LSD) was discovered by Hofmann at the Sandoz
Laboratories in 1938, in a research project that involved making novel derivatives of methergine R, a compound
then widely used to assuage haemorrhage during childbirth. Hofmann accidentally swallowed a minute quantity of
his semi-synthetic derivative and discovered its hallucinogenic properties. For a while, LSD was marketed to
psychiatrists, used unsuccessfully to treat schizophrenia, and adopted by the American CIA as a truth serum for
investigating suspected communists110. Most famously, during the 1960s, LSD was the favourite recreational drug of
the counterculture movement. Popular songs such as the Beatles’ ‘Lucy in the Sky with Diamonds’ and Jefferson
Airplane’s ‘White Rabbit’ glamourized its psychedelic effects. By the mid-1970s, however, government efforts to
criminalize its use, coupled with increased public awareness about the possibility of ‘bad trips’, succeeded in making
LSD less socially acceptable.
One provocative hypothesis developed by historians of the era was that ergotism might have contributed to the
notorious seventeenth-century witchcraft trials at Salem, Massachusetts111,112. More recently, Robin Cook, a popular
novelist, used the Salem–ergot hypothesis as the plot line for Acceptable Risk113. In this thriller, biotechnologists
isolate Claviceps spores from a damp New England cellar, culture the fungus, and extract a new alkaloid that
enhances intellect, stamina, libido and ego. Too late, they discover that their ‘billion-dollar drug’ has side effects,
turning those who consume it into flesh-eating zombies. In a crude way, this story highlights the paradox of most
powerful drugs: often the distinction between a toxin and a remedy is a fraction of a decimal point, the addition of a
methyl group or the removal of a double bond.

methodologies in the 1980s enabled dramatic progress ‘broad’-domain transcription factors. The narrow
in the genetics and biochemistry of fungal secondary pathway-specific regulators are usually found in the
metabolism. This rapid progress was facilitated by what cluster and positively regulate gene expression. These
is now considered a hallmark characteristic of second- proteins are often Zn(II)2Cys6 zinc binuclear cluster
ary metabolic biosynthetic pathways — the grouping proteins48–50, which are a class of proteins so far only
of pathway genes in a contiguous cluster. found in fungi. The archetypal protein in this group
Metabolic gene clustering was not predicted by the is AflR (aflatoxin regulator), the Zn(II)2Cys6 protein
fungal research community. In fact, in the era between that is required for aflatoxin and sterigmatocystin bio-
the discovery of the bacterial operon and the advent synthetic gene activation49–52. Typical for this group of
of large-scale eukaryotic gene cloning, it had become DNA-binding proteins, AflR recognizes and binds to
dogma that eukaryotic genes that are involved in a palindromic sequence found in the promoters of
functionally related pathways are not linked. By 1990 the biosynthetic genes50. Other transcription factors
however, this tenet had been abandoned owing to the that are encoded in biosynthetic gene clusters include
almost routine discovery of gene clusters in fungi for Cys2His2 zinc-finger proteins (Tri6 and MRTRI6 for
phenotypes as varied as nutrient use, mating type, trichothecene production48) and an ankyrin repeat
pathogenicity and secondary metabolism32,33. In less protein (ToxE for HC-toxin production53). Cluster
than a decade, it was shown that the genes for the pro- regulators not found in the cluster itself include
duction of a broad range of secondary metabolites were a two-peptide forkhead complex (AcFKH1 and
located adjacent to one another; in addition, a pathway- CPCR1) for cephalosporin production 54 and an
specific regulatory gene was often embedded in these HAP-like transcriptional complex (PENR1) for peni-
gene clusters. Fungal secondary metabolite clusters cillin55. Additionally, PENR1 has also been shown to
characterized by gene cloning include aflatoxins14,15, be important in taka-amylase, xylanase and cellobio-
cephalosporin 34 , compactin 35,36 , ergot alkaloids27 , hydrolase production56.
fumonisin37, gibberellins38,39, HC toxin40, lovastatin10, Secondary metabolite biosynthesis has long been
melanin41,42, paxillin26, penicillin43,44, sterigmatocystin13, known to be responsive to environmental cues,
sirodesmin45 and trichothecenes46,47. including the carbon and nitrogen source, ambient
temperature, light and pH57,58. Broad-domain factors
Transcription factors. The co-regulation of the clusters are transcription factors that are important in inte-
of genes that code for the synthesis of natural products grating cellular responses to these parameters. Several
can, in part, be explained by coordinated transcrip- studies59,60 indicate that responses to environmental
tional control of biosynthetic genes by ‘narrow’- or signals are transmitted through Cys2His2 zinc-finger

942 | DECEMBER 2005 | VOLUME 3 www.nature.com/reviews/micro


© 2005 Nature Publishing Group
REVIEWS

A. parasiticus73. AflR is inactivated by PkaA-mediated


phosphorylation, and the transcription of aflR is repressed
FlbA FadA PkaA Colony
growth
by PkaA activity74; the repression of aflR transcription is
mediated by PkaA through a novel methyltransferase,
LaeA75 (discussed later).
Increasing research into the roles of G-protein
Post- Transcriptional signalling in fungal development indicates that signal-
transcriptional
transduction pathways often positively or negatively
BrlA AflR LaeA
regulate secondary metabolism. For example, FadA
negatively regulates aflatoxin and sterigmatocystin
synthesis, but positively regulates penicillin produc-
Conidiation Sterigmatocystin Penicillin
tion in A. nidulans, and a FadA homologue also
Figure 5 | Integrating signal-transduction controls in positively regulates trichothecene production in
spore production (conidiation) and secondary
metabolism in Aspergillus nidulans. The model shown
F. sporotrichioides76. G proteins also regulate secondary
uses solid lines to indicate known pathways and dashed lines metabolism in Botrytis cinerea and Trichoderma atro-
to indicate hypothesized pathways. FadA, α subunit viride, among others77,78. In addition to its role in the
heterotrimeric G protein; FlbA, a regulatory protein that has regulation of secondary metabolism and sporulation,
an RGS (regulator of G protein signalling) motif; PkaA, G-protein signalling is crucial for pathogenicity. For
catalytic subunit of protein kinase A; BrlA, a conidiation- instance, deletion of cpg1, which encodes the α-subunit
specific transcription factor; AflR (aflatoxin regulator), a
of a heterotrimeric G protein in the chestnut blight
sterigmatocystin/aflatoxin-specific transcription factor.
fungus Cryphonectria parasitica, results in a marked
reduction in the fungal growth rate, reduced levels of
spore production, decrease in pigmentation and a loss
global transcription factors that mediate carbon61 of virulence79.
(CreA), nitrogen62 and pH63,64 (PacC) signalling. Broad
transcription factors can positively (PacC regulation of Epigenetic controls? The role of narrow pathway-
penicillin) or negatively (CreA regulation of penicillin) specific regulators and broad global transcription
regulate metabolite production and are conserved in all factors such as PacC in penicillin biosynthesis63 and
fungi and other eukaryotes. Regulation by both narrow- in regulation of secondary metabolite gene clusters
and broad-domain transcription factors ensures — coupled with the G-protein/cAMP/protein-kinase-
that secondary metabolite pathways can respond to mediated regulation of sterigmatocystin and aflatoxin
the demands of general cellular metabolism and the biosynthesis — provided the research community with
presence of specific pathway inducers. insights into the mechanisms of secondary-metabolite
regulation in fungi. None of these findings, however,
Secondary metabolism and fungal development. An could explain why metabolite-specific synthesis and
association between natural product formation and regulatory genes were clustered.
morphological development has been observed for The evolution and maintenance of gene clusters has
decades57,65. For example, one unusual class of mutants received a great deal of attention in bacteria80–82 and
in A. parasiticus named ‘sec’ are deficient in both fungi32,33,83,84. Evidence for horizontal transfer of the
sporulation and aflatoxin production66. Although the penicillin cluster from bacteria to fungi was met with
genes that encode the enzymes for aflatoxin synthesis great excitement, as it suggested a reason for both the
are present67, expression of the pathway-specific regu- origin and maintenance of metabolic clustering83,84,85.
lator, aflR, is 5–10-fold reduced in the toxigenic strains However, other functionally conserved metabolites,
compared with parental strains68. A similar example such as gibberellin production in the fungus G. fujikuroi
involves a mutation in the A. parasiticus fluP gene and in plants, are unlikely to have arisen from hori-
that results in a fluffy hyphal morphology phenotype, zontal transfer between kingdoms22, and it is likely
a reduction in the number of asexual spores produced that several evolutionary mechanisms account for the
and a reduction in the level of aflatoxin production69. presence of secondary-metabolite clusters. Although a
The correlation of secondary metabolism with unifying model to explain the clustering of secondary-
fungal development has recently been reviewed70, and metabolite genes in fungi is lacking, the preservation
the authors made special note of the joint regulation of of clusters, now known to be extensive through scru-
these processes through a G-protein–protein kinase A tiny of fungal genomes, might indicate an underlying
signal-transduction pathway. FadA, a G protein in advantage to maintaining clustering.
A. nidulans, and PkaA, a protein kinase that is down- An exciting advance in support of a global require-
stream in the FadA regulatory pathway, are particularly ment for fungal secondary-metabolite gene clustering
well characterized. Both proteins negatively regulate has arisen from identification of an Aspergillus methyl-
aflatoxin and sterigmatocystin synthesis71,72 (FIG. 5). transferase, LaeA75. LaeA was originally identified by
PkaA regulates production of sterigmatocystin tran- complementation of a sterigmatocystin biosynthetic
scriptionally and post-transcriptionally through aflR mutant. Deletion and overexpression of laeA in
and its gene product72, and it is postulated that a simi- A. nidulans, A. fumigatus and A. terreus strains either
lar regulatory cascade controls aflatoxin regulation in silences or increases, respectively, the production

NATURE REVIEWS | MICROBIOLOGY VOLUME 3 | DECEMBER 2005 | 943


© 2005 Nature Publishing Group
REVIEWS

Heterochromatin nucleosome biosynthetic genes are detected by sequence similar-


ity to fungal terpene cyclases and with the DMATSs
of Claviceps species. As terpene cyclase primary
sequences have diverged, it is more difficult to detect
Euchromatin Heterochromatin Euchromatin the genes that encode these proteins.
LaeA
Oxidoreductases, methylases, acetylases, esterases
and transporters are not exclusive to secondary metab-
olism, so homology searches are more problematic for
these genes. Nevertheless, if good candidate genes
encoding these enzymes are found adjacent to signa-
ture secondary metabolic genes such as NRPS, PKS
and DMATS homologues, there is an improved chance
Euchromatin Euchromatin Euchromatin that they are involved in production of a secondary
Figure 6 | Model of LaeA function. Secondary metabolite gene clusters such as the
metabolite.
sterigmatocystin gene cluster are maintained in silenced heterochromatin but are surrounded Using these criteria, several important genes have
by expressed euchromatin. The LaeA protein functions to initiate a process that converts been identified from fungal genomic-DNA-sequence
heterochromatin to euchromatin, perhaps by interfering with methylases or deacetylases that data; a summary is given in TABLE 1. Most of these are
are associated with heterochromatin. Sterigmatocystin genes are designated with thick present in putative gene clusters. In some cases, we
arrows, thin arrows indicate genes that are located on either side of the sterigmatocystin know the function of the cluster from pre-genomic
cluster and that are not regulated by LaeA.
studies, but the function of most of these putative
clusters will remain speculative until functional stud-
ies have been carried out. More putative secondary
of several secondary metabolites and the expression metabolite clusters are present than are needed to
of their respective genes 75. Sequence analysis of laeA account for the known products, and some of these
and analysis of the encoded protein indicated that it clusters might not be expressed under laboratory con-
is a protein methyltransferase75. The highest sequence ditions. For example, the aflatoxin cluster of A. oryzae
similarity is to histone and arginine methyltransferases, is not expressed under conditions that are favourable
which have important roles in the regulation of gene to aflatoxin expression in A. flavus and A. parasiticus
expression, such as defining the boundaries of EUCHRO (reviewed in REF. 33).
MATIC and HETEROCHROMATIC chromosomal domains Comparative studies allow a confident predic-
in the mating locus of yeast and the β-globin locus in tion of the likely gliotoxin cluster of A. fumigatus,
mice86–88. Although LaeA functions are not yet fully based on its resemblance to the sirodesmin cluster
characterized, we speculate that this protein is involved of Leptosphaeria maculans. Both compounds are
in chromatin modification, perhaps reminiscent of piperazines. Sirodesmin is derived from tyrosine and
mating-type locus expression in yeast. Methylation serine, and gliotoxin is derived from phenylalanine
is involved in both gene repression and activation of and serine 45 . Each cluster includes a bimodular
heterochromatic and euchromatic regions of eukaryo- NRPS, which is presumably required for condensa-
tic chromosomes89, so one model of LaeA-mediated tion of the precursor amino acids. The A. fumigatus
regulation of clusters invokes a role in heterochromatin genome also has seven significant hits in the public
repression (FIG. 6). databases (ranging from 22%, expect 5e-15, to 52%,
expect e-115 in a BLAST search) with the DMATS of
Bioinformatics and gene predictions C. purpurea, and the best hit (TIGR assigned number
Automated and manual annotations of the N. crassa, Afu2g18040 accessible at the Central Aspergillus Data
A. fumigatus, A. nidulans and A. oryzae genomes Repository (see Online links box)) is present in a clus-
predict gene functions based on homology to char- ter that contains other genes resembling those found
acterized genes and their products. It is easy to find in C. purpurea. No NRPS is present in the putative
genes encoding putative NRPSs and PKSs based on A. fumigatus cluster, as expected from the absence
protein domain structures. Terpene and alkaloid of reports of ergopeptine synthesis from this species,

Table 1 | Summary of secondary-metabolite gene classes in Aspergillus


EUCHROMATIC
Describes chromosome regions Protein Aspergillus oryzae Aspergillus fumigatus Aspergillus nidulans
with actively transcribed genes. PKS 30 14 27
Generally these regions stain
poorly or not at all. NRPS 18 14 14
FAS 5 1 6
HETEROCHROMATIC
Describes chromosomal regions Sesquiterpene cyclase 1 Not detected 1
that are generally genetically
inert. The chromatin is tightly DMATS 2 7 2
coiled throughout the cell cycle PKS, polyketide synthase; NRPS, non-ribosomal peptide synthetase; FAS, fatty-acid synthase; DMATS, dimethylallyl tryptophan
and stains well. synthetase.

944 | DECEMBER 2005 | VOLUME 3 www.nature.com/reviews/micro


© 2005 Nature Publishing Group
REVIEWS

PARALOGUES which produces its own group of clavines90,91. Further has a homologue of the aristolochene cyclase gene found
Genes that are derived from a homologues of DMATS genes are found in this and in P. roquefortii. A. terreus and A. oryzae have a homo-
common ancestor by another cluster, and these might be prenyl transferases logue of trichothecene cyclase, but no obvious terpene
duplication. They can have
that are required for prenylation steps required in cyclase is detectable in A. fumigatus.
related functions.
fumigaclavine and fumitremorgen biosynthesis.
ORTHOLOGUES There are still many PKS and NRPS genes (and Conclusions
Genes that are derived from a associated clusters) for which the putative products Secondary metabolites are low-molecular-weight natural
common ancestor by a cannot be predicted. In bacteria, prediction of NRPS products with restricted taxonomic distribution, often
speciation event. They usually
have equivalent function in
products has been partially successful through the use synthesized after active growth has ceased, which do
their respective species. of a derivation of a module-recognition code that is not have an obvious function in producer species. The
based on correlation of NRPS adenylation domain β-lactam antibiotic penicillin, which is synthesized by
structure to a particular amino acid92. However, this an NRPS, and the polyketides aflatoxin and sterigmato-
algorithm is not sufficiently robust to predict the cystin, which are synthesized through a polyketide
amino-acid specificity of fungal NRPS modules, partly pathway, are among the best studied fungal secondary
because of the limited experimental information metabolites, and their pathways have become paradigms.
available. Similarly, it is not yet possible to predict the The discovery that genes involved in their production
structure of any polyketide of an iterative, type I fungal are clustered, as are the genes that code for the produc-
PKS through analysis of domain and motif structures tion of the vast majority of other secondary metabolites
alone. For example, automated annotation might label that have been studied, has important implications for
some fungal PKSs as ‘lovastatin biosynthesis’ genes. gene regulation and evolution. Pragmatically, it means
This classification simply reflects similarity to one that putative biosynthetic pathway genes for secondary
of the lovastatin PKS genes, and probably indicates a metabolism can easily be detected by in silico analysis
related PARALOGUE rather than a true ORTHOLOGUE. of genomic data. Although bioinformatics alone can-
Several sequenced fungal genomes contain a hybrid not predict pathway products, or even determine if the
PKS–NRPS gene. This arrangement has been observed genes are expressed, these analyses can reveal molecular
previously in bacteria, and illustrates how genome data evidence of many hitherto undiscovered pathways.
add to our knowledge of the diversity of PKS and NRPS The recent characterization of LaeA, which is a
structures found in nature. A hybrid gene in Fusarium global regulator of secondary metabolism, provides
verticilliodes (formerly Fusarium moniliforme) produces a tool for detecting gene clusters, and might reveal
a precursor of the toxin fusarin C 93, and a PKS/NRPS is novel chromatin-based mechanisms of transcriptional
a virulence factor in the rice blast fungus Magnaporthe control of these clusters. The laeA gene has been found
grisea94. Genes of this hybrid class can be detected in in all aspergilli examined to date; however, it remains
several fungal genomes; however, sequence compari- to be seen if LaeA functions in other fungi.
sons indicate that most are paralogues to the fusarin C
gene, and in most cases the products are unknown. Note added in proof
Bioinformatic analysis of the Aspergillus genomes Three recent papers describe the genomes of A. fumi-
has revealed few terpene biosynthetic genes. A. nidulans gatus114, A. nidulans115 and A. oryzae116.

1. Turner, W. B. Fungal Metabolites (Academic Press, London, 12. Donadio S., Staver, M. J., McAlpine, J. B., Swanson, S. J. & encoded by a giant 45.8-kilobase open reading frame.
1971). Katz, L. Modular organization of gene required for complex Curr. Genet. 26, 120–125 (1994).
2. Turner, W. B. & Aldridge, D. C. Fungal Metabolites II polyketide biosynthesis. Science 252, 675–679 (1991). 21. Eisendle, M., Oberegger, H., Zadra, I. & Haas, H. The
(Academic Press, London, 1983). 13. Brown, D. et al. Twenty-five co-regulated transcripts define siderophore system is essential for viability of Aspergillus
3. Cole, R. & Schweikert, M. Handbook of Secondary Fungal a sterigmatocystin gene cluster in Aspergillus nidulans. nidulans: functional analysis of two genes encoding L-ornithine
Metabolites Volumes 1–3 (Elsevier, Amsterdam, 2003). Proc. Natl Acad. Sci. USA 93, 1418–1422 (1996). N5-monooxygenase (sidA) and a nonribosomal peptide
4. Davies, J. Recombinant DNA and the Production of Small 14. Yu, J. et al. Clustered pathway genes in aflatoxin bio- synthetase (sidC). Mol. Microbiol. 49, 359–375 (2003).
Molecules (ASM Press, Washington DC, 1985). synthesis. Appl. Environ. Microbiol. 70, 1253–1262 (2004). 22. Tudzynski, B., Hedden, P. Carrera, E. & Gaskin, P.
5. Bennett, J. W. & Bentley, R. What’s in a name? Microbial 15. Yu, J., D. Bhatnagar, D. & Cleveland, T. D. Completed The 450–4 gene of Gibberella fujikuroi encodes ent-
secondary metabolism. Adv. Appl. Microbiol. 34, 1–28 (1989). sequence of aflatoxin pathway gene cluster in Aspergillus kaurine oxidase in the gibberellin biosynthesis pathway.
6. Ciba Foundation Symposium 171. Secondary Metabolites: parasiticus. FEBS Lett. 564, 126–130 (2004). Appl. Environ. Microbiol. 67, 3514–3522 (2001).
Their Function and Evolution (John Wiley & Sons, 16. Finking, R. & Marahiel, M. Biosynthesis of nonribosomal 23. Rynkiewicz, M. J., Cane, D. E. & Christianson, D. W.
Chicester, 1992). peptides. Annu. Rev. Microbiol. 58, 453–488 (2004). Structure of trichodiene synthase from Fusarium
7. Raistrick, H. A region of biosynthesis. Proc. R. Soc. Lond. 17. Smith, D. J., Earl, A. J., Turner, G. The multifunctional sporotrichioides provides mechanistic inferences on the
B Biol. Sci. 136, 481–508 (1950). peptide synthetase performing the first step of penicillin terpene cyclization cascade. Proc. Natl Acad. Sci. USA
8. Pelaez, F. Biological activities of fungal metabolites. biosynthesis in Penicillium chrysogenum is a 421073 98, 13543–13548 (2001).
in Handbook of Industrial Mycology (ed. An, Z.) 49–92 dalton protein similar to Bacillus brevis peptide antibiotic 24. Carruthers, J., Kang, I., Rynkiewicz, M., Cane, D. &
(Marcel Dekker, New York, 2005). synthetases. EMBO J. 9, 2743–2750 (1990). Christianson, D. Crystal structure determination of
9. Fujii, I., Watanabe, A., Sankawa, U. & Ebizuka, Y. First indication of the multimodular structure of aristolochene synthase from the cheese mold,
Identification of Claisen cyclase domain in fungal peptide synthetases: three modules were detected Penicillium roquefortii. J. Biol. Chem. 275, 25533–25539
polyketide synthase WA, a naphthopyrone synthase of in this tripeptide synthetase. (2000).
Aspergillus nidulans. Chem. Biol. 8, 189–197 (2001). 18. Kallow, W., Kennedy, J., Arezi, B., Turner, G. & von Doehren, Structural determination of a fungal terpenecyclase,
10. Kennedy, J. et al. Modulation of polyketide synthase H. Thioesterase domain of δ-(L-α-aminoadipyl)-L-cysteinyl- showing that whereas the primary sequences of
activity by accessory proteins during lovastatin D-valine synthetase: alteration of stereospecificity by site- terpene cyclases are not well conserved between
biosynthesis. Science 284, 1368–1372 (1999). directed mutagenesis. J. Mol. Biol. 297, 395–408 (2000). plants and fungi, the tertiary structure is conserved.
The first biochemical dissection of fungal polyketide 19. Wiest, A. et al. Identification of peptaibols from Also, terpene cyclases might all be derived from a
synthase and use of the model system A. nidulans to Trichoderma virens and cloning of a peptaibol synthetase. common ancestor.
help decipher lovastatin assembly in A. terreus. J. Biol. Chem. 277, 20862–20868 (2002). 25. Schmidhauser, T., Lauter, F., Russo, V. & Yanofsky, C.
11. Bentley, R. & Bennett, J. W. Constructing polyketides: 20. Weber, G., Schorgendorfer, K., Schneider-Scherzer, E. & Cloning, sequence, and photoregulation of al-1,
from Collie to combinatorial biosynthesis. Annu. Rev. Leitner, E. The peptide synthetase catalyzing a carotenoid biosynthetic gene of Neurospora crassa.
Microbiol. 53, 411–446 (1999). cyclosporine production in Tolypocladium niveum is Mol. Cell. Biol. 10, 5064–5070 (1990).

NATURE REVIEWS | MICROBIOLOGY VOLUME 3 | DECEMBER 2005 | 945


© 2005 Nature Publishing Group
REVIEWS

26. Young, C., McMillan, L., Telfer, E. & Scott, B. Molecular 49. Woloshuk, C. et al. Molecular characterization of aflR, a 71. Hicks, J., Yu, J., Keller, N. & Adams, T. Aspergillus
cloning and genetic analysis of an indole-diterpene gene regulatory locus for aflatoxin biosynthesis. Appl. Environ. sporulation and mycotoxin production both require
cluster from Penicillium paxilli. Mol. Microbiol. 39, Microbiol. 60, 2408–2414 (1994). inactivation of the FadA G α protein-dependent signaling
754–764 (2001) Identification of the first Zn(II)2Cys6 that regulates a pathway. EMBO J. 16, 4916–4923 (1997).
27. Tudzynski, P. et al. Evidence for an ergot alkaloid gene secondary metabolite gene cluster. This paper reported the genetic connection of
cluster in Claviceps purpurea. Mol. Gen. Genet. 261, 50. Fernandes, M., Keller, N. & Adams, T. Sequence-specific sporulation and secondary metabolism through a
133–141 (1999). binding by Aspergillus nidulans AflR, a C6 zinc cluster G-protein signalling pathway.
Identification of the ergot alkaloid gene cluster, which protein regulating mycotoxin biosynthesis. Mol. 72. Shimizu, K. & Keller, N. Genetic involvement of a
includes an NRPS required for ergotamine biosynthesis. Microbiol. 28, 1355–1365 (1998). cAMP-dependent protein kinase in a G protein signaling
28. von Nussbaum, F. Stephacidin B — a new stage of 51. Chang, P., Ehrlich, K., Yu, J., Bhatnagar, D. & Cleveland, pathway regulating morphological and chemical
complexity within prenylated indole alkaloids from fungi. T. Increased expression of Aspergillus parasiticus aflR, transitions in Aspergillus nidulans. Genetics 157,
Angew. Chem. Int. Ed. Engl. 42, 3068–3071 (2003). encoding a sequence-specific DNA-binding protein, 591–600 (2001).
29. Bennett, J. W., Chang, P.-K. & Bhatnagar, D. One gene relieves nitrate inhibition of aflatoxin biosynthesis. 73. Roze, L., Beaudry, R., Keller, N. & Linz, J. Regulation of
to whole pathway: the role of norsolorinic acid in Appl. Environ. Microbiol. 61, 2372–2377 (1995). aflatoxin synthesis by FadA/cAMP/protein kinase A
aflatoxin research. Adv. Appl. Microbiol. 45, 1–15 (1997). 52. Yu, J. et al. Conservation of structure and function of the signaling in Aspergillus parasiticus. Mycopathologia 158,
30. Luengo, J. M. & Penalva, M. A. Penicillin Biosynthesis aflatoxin regulatory gene aflR from Aspergillus nidulans 219–232 (2004).
in Aspergillus: 50 Years On (eds Martinelli, S. D & and A. flavus. Curr. Genet. 29, 549–555 (1996). 74. Shimizu, K., Hicks, J., Huang T.-P. & Keller, N. P. Pka,
Kinghorn, J. R.) 603–638 (Elsevier, Amsterdam,1994). 53. Pedley, K. & Walton, J. Regulation of cyclic peptide Ras and RGS protein interactions regulate activity of
31. Rehacek, Z. & Sajdl, P. Ergot Alkaloids: Chemistry, biosynthesis in a plant pathogenic fungus by a novel AflR, a Zn(II)2Cys6 transcription factor in Aspergillus
Biological Effects, Biotechnology (Academia, Prague, 1990) transcription factor. Proc. Natl Acad. Sci. USA 98, nidulans. Genetics 165, 1095–1104 (2003).
32. Keller, N. & Hohn, T. Metabolic pathway gene clusters in 14174–14179 (2001). 75. Bok, J. & Keller, N. LaeA, a regulator of secondary
filamentous fungi. Fungal Genet. Biol. 21, 17–29 (1997). 54. Schmitt, E. K., Hoff, B. & Kuck, U. AcFKH1, a novel metabolism in Aspergillus. Euk. Cell 3, 527–535 (2004).
33. Zhang, Y.-Q., Wilkinson, H., Keller, N. P. & Tsitsigiannis, D. member of the forkhead family, associates with the RFX Discovery of novel global regulator of several
Secondary metabolite gene clusters. in Handbook of transcription factor CPCR1 in the cephalosporin C- Aspergillus secondary metabolites.
Industrial Microbiology (ed. An, Z.) 355–386 (Marcel producing fungus Acremonium chrysogenum. Gene 76. Tag, A. et al. G-protein signalling mediates differential
Dekker, New York, 2005). 342, 269–281 (2004). production of toxic secondary metabolites. Mol.
34. Gutierrez, S., Velasco, J., Fernandez, F. J. & Martin, J. F. In contrast to penicillin regulation in A. nidulans, Microbiol. 38, 658–665 (2000).
The cefG gene of Cephalosporium acremonium is linked cephalosporin regulation is by a forkhead 77. Schulze Gronover, C., Schorn, C. & Tydzynski, B.
to the cefEF gene and encodes a deacetylcephalosporin transcription factor in Acremonium. Identification of Botrytis cinerea genes up-regulated
C acetyltransferase closely related to homoserine 55. Litzka, O., Papagiannopolus, P., Davis, M., Hynes, M. & during infection and controlled by the G α subunit
O-acetyltransferase. J. Bacteriol. 174, 3056–3064 (1992). Brakhage, A. The penicillin regulator PENR1 of BCG1 using suppression subtractive hybridization
35. Abe, Y. et al. Effect of increased dosage of the ML-236B Aspergillus nidulans is a HAP-like transcriptional (SSH). Mol. Plant Microbe Interact. 17, 537–546
(compactin) biosynthetic gene cluster on ML-236B complex. Eur. J. Biochem. 251, 758–767 (1998). (2004).
production in Penicillium citrinum. Mol. Gen. Genet. 268, A HAP-like CCAAT-binding complex regulates 78. Reithner, B. et al. The G protein α subunit Tga1 of
130–137 (2002). penicillin biosynthesis. Trichoderma atroviride is involved in chitinase formation
36. Abe, Y., Ono, C., Hosobuchi, M. & Yoshikawa, H. 56. Brakhage, A. et al. HAP-like CCAAT-binding complexes and differential production of antifungal metabolites.
Functional analysis of mlcR, a regulatory gene for in filamentous fungi: implications for biotechnology. Fungal Genet. Biol. 42, 749–760 (2005).
ML-236B (compactin) biosynthesis in Penicillium Fungal Genet. Biol. 27, 243–252 (1999). 79. Gao, S. & Nuss, D. Distinct roles for two G protein α
citrinum. Mol. Genet. Genomics 268, 352–361 (2002). 57. Bennett, J. & Ciegler, A. (eds) Secondary Metabolism and subunits in fungal virulence, morphology, and
37. Proctor, R. H., Brown, D. W., Plattner, R. D. & Differentiation in Fungi. (Marcel Dekker, New York, 1983). reproduction revealed by targeted gene disruption.
Desjardins, A. E. Co-expression of 15 contiguous genes 58. Berry, D. R. (ed.) Physiology of Industrial Fungi (Blackwell Proc. Natl Acad. Sci. USA 93, 14122–14127 (1996).
delineates a fumonisin biosynthetic gene cluster in Scientific Publishing, Oxford, 1988). 80. Lawrence, J. G. & Roth, J. R. Selfish operons: horizontal
Gibberella moniliformis. Fungal Genet. Biol. 38, 237–249 59. Ehrlich, K., Montalbano, B. & Cotty, P. Sequence transfer may drive the evolution of gene clusters.
(2003). comparison of aflR from different Aspergillus species Genetics 143, 1843–1860 (1996).
38. Hedden, P., Phillips, A., Rojas, M., Carrera, C. & provides evidence for variability in regulation of aflatoxin 81. Lawrence, J. G. Selfish operons and speciation by gene
Tudzynski, B. Gibberellin biosynthesis in plants and production. Fungal Genet. Biol. 38, 63–74 (2003). transfer. Trends Microbiol. 5, 355–359 (1997).
fungi: a case of convergent evolution? J. Plant Growth 60. Tudzynski, B., Homann, V., Feng, B. & Marzluf, G. 82. Lawrence, J. G. Gene transfer, speciation, and the
Regul. 20, 319–331 (2002). Isolation, characterization and disruption of the areA evolution of bacterial genomes. Curr. Opin. Microbiol. 2,
39. Tudzynski, B. Biosynthesis of gibberellins in Gibberella nitrogen regulatory gene of Gibberella fujikuroi. Mol. Gen. 519–523 (1999).
fujikuroi: biomolecular aspects. Appl. Environ. Microbiol. Genet. 261, 106–114 (1999). 83. Rosewich, U. & Kistler, H. Role of horizontal gene
52, 298–310 (1999). 61. Dowzer, C. & Kelly, J. Cloning of the creA gene from transfer in the evolution of fungi. Annu. Rev. Phytopathol.
40. Ahn, J., Cheng, Y. & Walton, J. An extended physical Aspergillus nidulans: a gene involved in carbon 38, 325–363 (2000).
map of the TOX2 locus of Cochliobolus carbonum catabolite repression. Curr. Genet. 15, 457–459 (1989). 84. Walton, J. J. Horizontal gene transfer and the evolution
required for biosynthesis of HC-toxin. Fungal Genet. 62. Kudla, B. et al. The regulatory gene areA mediation of secondary metabolite gene clusters in fungi: an
Biol. 35, 31–38 (2002). nitrogen metabolite repression in Aspergillus nidulans. hypothesis. Fungal Genet. Biol. 30, 167–171 (2000).
41. Kimura, N. & Tsuge, T. Gene cluster involved in melanin Mutations affecting specificity of gene activation alter a 85. Smith, M. W., Feng, D.-F. & Doolittle, R. F. Evolution by
biosynthesis of the filamentous fungus Alternaria loop residue of a putative zinc finger. EMBO J. 9, acquisition: the case for horizontal gene transfers.
alternata. J. Bacteriol. 175, 4427–4435 (1993). 1355–1364 (1990). Trends Biochem. Sci. 17, 489–493 (1992).
42. Tsai, H., Wheeler, M., Chang, Y. & Kwon-Chung, K. 63. Tilburn, J. et al. The Aspergillus PacC zinc finger 86. Litt, M., Simpson, M., Gaszner, M., Allis, D. & Felsenfeld,
A developmentally regulated gene cluster involved in transcription factor mediates regulation of both acid- G. Correlation between histone lysine methylation and
conidial pigment biosynthesis in Aspergillus fumigatus. and alkaline-expressed genes by ambient pH. developmental changes at the chicken β-globin locus.
J. Bacteriol. 181, 6469–6477 (1999). EMBO J. 14, 779–790 (1995). Science 293, 2453–2455 (2001).
43. Smith, D. J. et al. β-lactam antibiotic biosynthetic genes Important contribution that showed that pH 87. Recillas-Targa, F. et al. Position-effect protection and
have been conserved in clusters in prokaryotes and regulates the penicillin gene cluster through a enhancer blocking by the chicken β-globin insulator are
eukaryotes. EMBO J. 9, 741–747 (1990). global transcription factor, PacC. separable activities. Proc. Natl Acad. Sci. USA 99,
Showed that secondary metabolic genes were 64. Martin, J. Molecular control of expression of penicillin 6883–6888 (2002).
clustered in filamentous fungi, and revealed the biosynthesis genes in fungi: regulatory proteins interact 88. Noma, K., Allis, C. & Grewal, S. Transitions in distinct
close relationship between the β-lactam with a bidirectional promoter region. J. Bacteriol. 182, histone H3 methylation patterns at the heterochromatin
biosynthetic genes of bacteria and fungi. 2355–2362 (2000). domain boundaries. Science 293, 1150–1155 (2001).
44. Brakhaage, A. A. Molecular regulation of β-lactam 65. Luckner, M. Secondary Metabolism in Microorganisms, 89. Lee, D. Y., Teyssier, C., Strahl, B. D. & Stallcup, M. R.
biosynthesis in filamentous fungi. Microbiol. Mol. Biol. Plants and Animals (Springer–Verlag, Berlin, 1990). Role of protein methylation I regulation of transcription.
Rev. 62, 547–585 (1998). 66. Kale, S., Bhatnagar D. & Bennett, J. Isolation and Endocr. Rev. 26, 147–170 (2005).
45. Gardiner, D., Cozijnsen, A., Wilson, L., Pedras, M. & characterization of morphological variants of Aspergillus 90. Spilsbury, J. F. & Wilkinson, S. The isolation of festuclavine
Howlett, B. The sirodesmin biosynthetic gene cluster of parasiticus deficient in secondary metabolite production. and two new clavine alkaloids from Aspergillus fumigatus
the plant pathogenic fungus Leptosphaeria maculans. Mycol. Res. 98, 645–652 (1994). Fres. J. Chem. Soc. 5, 2085–2091 (1961).
Mol. Microbiol. 53, 1307–1318 (2004). 67. Kale, S., Cary, J., Bhatnagar, D. & Bennett, J. 91. Cole, R. J. et al. Mycotoxins produced by Aspergillus
46. Trapp, S., Hohn T., McCormick, S. & Jarvis, B. Characterization of an experimentally induced, fumigatus species isolated from molded silage.
Characterization of the gene cluster for biosynthesis of nonaflatoxigenic variant strains of Aspergillus parasiticus. J. Agric. Food Chem. 25, 826–830 (1977).
macrocyclic trichothecenes in Myrothecium roridum. Appl. Environ. Microbiol. 62, 3999–3404 (1996). 92. Challis, G. L, Ravel, J. & Townsend, C. A. Predictive,
Mol. Gen. Genet. 257, 421–432 (1998). 68. Kale, S. et al. Genetic analysis of morphological variants structure-based model of amino acid recognition by
47. Brown, D., McCormick, S., Alexander, N., Proctor, R. & of Aspergillus parasiticus deficient in secondary nonribosomal peptide synthetase adenylation domains.
Desjardins, A. A genetic and biochemical approach to metabolite production. Mycol. Res. 107, 831–840 Chem. Biol. 7, 211–224 (2000).
study trichothecene diversity in Fusarium (2003). 93. Song, Z., Cox, R. J., Lazarus, C. M. & Simpson, T. J.
sporotrichioides and Fusarium graminearum. Fungal 69. Zhou, R., Rasooly, R. & Linz, J. Isolation and analysis of Fusarin C biosynthesis in Fusarium moniliforme and
Genet. Biol. 32, 121–133 (2001). fluP, a gene associated with hyphal grown and Fusarium venenatum. ChembioChem. 5, 1196–1203
48. Proctor, R, Hohn, T., McCormick, S. & Desjardins, A. sporulation in Aspergillus parasiticus. Mol. Gen. Genet. (2004).
Tri6 encodes an unusual zinc finger protein involved in 264, 514–520 (2000). 94. Bohnert, H. U. et al. A putative polyketide synthase/
regulation of trichothecene biosynthesis in Fusarium 70. Calvo, A., Wilson, R., Bok, J. & Keller, N. Relationship peptide synthetase from Magnaporthe grisea signals
sporotrichioides. Appl. Environ. Microbiol. 61, 1923– between secondary metabolism and fungal pathogen attack to resistant rice. Plant Cell 16,
1930 (1995). development. Mol. Microbiol. Rev. 66, 447–459 (2002). 2499–2513 (2004).

946 | DECEMBER 2005 | VOLUME 3 www.nature.com/reviews/micro


© 2005 Nature Publishing Group
REVIEWS

95. Hobby, G. Penicillin: Meeting the Challenge (Yale 107. Green, G. The Human Factor (Everyman’s Library, London, Competing interests statement
University Press, New Haven, 1985). 1979). The authors declare no competing financial interests.
96. Wainwright, M. Miracle Cure: the Story of Penicillin and the 108. Centers for Disease Control and Prevention (CDC).
Golden Age of Antibiotics (Blackwell Publishing, Oxford, Outbreak of aflatoxin poisoning — eastern and central
1990). provinces, Kenya. January–July 2004. Morb. Mortal. Online links
97. Bennett, J. & Chung, K. Alexander Fleming and the discovery Wkly Rep. 53, 790–793 (2004).
of penicillin. Adv. Appl. Microbiol. 49, 163–184 (2001). 109. Bennett, J. W. & Bentley, R. Pride and prejudice: the DATABASES
98. Lax, A. The Mold in Dr Florey’s Coat. The Story of the story of ergot. Persp. Biol. Med. 42, 333–355 (1999). The following terms in this article are linked online to:
Penicillin Miracle (Henry Holt & Company, New York, 110. Ulrich, R. F. & Paten, B. M. The rise, decline and fall of LSD. Entrez: http://www.ncbi.nlm.nih.gov/Entrez
2004) Persp. Biol. Med. 34, 561–578 (1991). Aspergillus flavus | Aspergillus parasiticus | Aspergillus terreus |
99. Scoutaris, M. “Moldy Mary” and the Illinois Fruit and 111. Caporeal, L. Ergotism: the Satan loosed in Salem? Fusarium sporotrichioides | Magnaporthe grisea |
Vegetable Company. Pharm. Hist. 38, 175–177 (1996). Science 192, 21–26 (1976). Neurospora crassa
100. Bentley, R. The molecular structure of penicillin. 112. Matossian, M. Ergot and the Salem witchcraft affair. Am.
J. Chem. Ed. 81, 1462–1470 (2004). Scientist 70, 355–357 (1982). FURTHER INFORMATION
101. Kuiper-Goodman, T. Food safety: mycotoxins and 113. Cook, R. Acceptable Risk (Barkley, New York, 1996). Nancy P. Keller’s homepage:
phycotoxins in perspective. In Mycotoxins and Phycotoxins 114. Nierman, W. C. et al. Genomic sequence of the http://www.plantpath.wisc.edu/NewFacPage/faculty.htm
— Developments in Chemistry, Toxicology and Food pathogenic and allergenic filamentous fungus Aspergillus Geoffrey Turner’s homepage:
Safety. (eds Miraglia, M., van Edmond, H., Brera, C. & fumigatus. Nature (in the press). http://www.shef.ac.uk/mbb/staff/turner
Gilbert, J.) 25–48 (Alaken Inc., Fort Collins, 1998) 115. Galagan, J. et al. Sequencing and comparative analysis of The Aspergillus website:
102. Squire, R. A. Ranking animal carcinogens. A proposed Aspergillus nidulans. Nature (in the press). http://www.aspergillus.man.ac.uk
regulatory approach. Science 214, 877–880 (1981). 116. Machida, M. et al. Genome sequencing and analysis of The Aspergillus nidulans Database:
103. Eaton, D. & Groopman, J. (eds) The Toxicology of Aspergillus oryzae. Nature (in the press). http://www.broad.mit.edu/annotation/fungi/aspergillus
Aflatoxins: Human Health, Veterinary, and Agricultural Central Aspergillus Data Repository:
Significance (Academic Press, San Diego, 1998). Acknowledgements http://www.cadre.man.ac.uk
104. Payne, G. & Brown, M. Genetics and physiology of Genomic data for Aspergillus fumigatus were provided by The Database of the Genomes Analysed at the National
aflatoxin biosynthesis. Annu. Rev. Plant Path. 36, 329–362 Institute for Genomic Research and The Wellcome Trust Sanger Institute of Advanced Industrial Science and Technology:
(1998). Institute; genomic data for Aspergillus nidulans were provided http://www.bio.nite.go.jp/dogan/Top
105. Hicks, J., Shimizu, K. & Keller, N. Genetics and by The Broad Institute; and genomic data for Aspergillus oryzae The TIGR Aspergillus fumigatus Genome Project:
biosynthesis of aflatoxins and sterigmatocystin. in The were provided by The National Institute of Advanced Industrial http://www.tigr.org/tdb/e2k1/afu1
Mycota. Volume XI. Agricultural Applications, (ed. Science and Technology. Coordination of the analyses of these The Wellcome Trust Sanger Institute Aspergillus fumigatus
Kempken, F.) 55–69 (Springer–Verlag, Berlin, 2002). data was enabled by an international collaboration involving Genome Project:
106. Zilinskas, R. A. Iraq’s biological weapons. The past as more than 50 institutions from 10 countries and coordinated http://www.sanger.ac.uk/Projects/A_fumigatus
future? J. Amer. Med. Assoc. 278, 418–424 (1997). from Manchester, UK. Access to this interactive links box is free online.

NATURE REVIEWS | MICROBIOLOGY VOLUME 3 | DECEMBER 2005 | 947


© 2005 Nature Publishing Group

You might also like