You are on page 1of 12

ARTICLE IN PRESS

Metabolic Engineering 10 (2008) 281–292

Contents lists available at ScienceDirect

Metabolic Engineering
journal homepage: www.elsevier.com/locate/ymben

Improving production of bioactive secondary metabolites in actinomycetes


by metabolic engineering
Carlos Olano, Felipe Lombó, Carmen Méndez, José A. Salas 
Departamento de Biologı́a Funcional e Instituto Universitario de Oncologı́a del Principado de Asturias (IUOPA), Universidad de Oviedo, 33006 Oviedo, Spain

a r t i c l e in fo abstract

Article history: Production of secondary metabolites is a process influenced by several physico-chemical factors
Received 10 June 2008 including nutrient supply, oxygenation, temperature and pH. These factors have been traditionally
Received in revised form controlled and optimized in industrial fermentations in order to enhance metabolite production. In
8 July 2008
addition, traditional mutagenesis programs have been used by the pharmaceutical industry for strain
Accepted 9 July 2008
Available online 15 July 2008
and production yield improvement. In the last years, the development of recombinant DNA technology
has provided new tools for approaching yields improvement by means of genetic manipulation of
Keywords: biosynthetic pathways. These efforts are usually focused in redirecting precursor metabolic fluxes,
Streptomyces deregulation of biosynthetic pathways and overexpression of specific enzymes involved in metabolic
Metabolism
bottlenecks. In addition, efforts have been made for the heterologous expression of biosynthetic gene
Antibiotic biosynthesis
clusters in other organisms, looking not only for an increase of production levels but also to speed the
Precursor engineering
Heterologous expression process by using rapidly growing and easy to manipulate organisms compared to the producing
Regulation organism. In this review, we will focus on these genetic approaches as applied to bioactive secondary
Ribosome engineering metabolites produced by actinomycetes.
Genome shuffling & 2008 Elsevier Inc. All rights reserved.

1. Introduction yields increase then costs are reduced. The large-scale production
of drugs from microbial fermentation has been the basis of the
Microorganisms are the source of many drugs including industry since the development of penicillin in the 1940s. The
antibiotics, antitumor compounds, immunosuppressants, antivir- titers of products made by industrial cultures nowadays are very
al, antiparasitic agents and enzyme-inactivating compounds. high after years of intense improvement programs using the
About 23,000 bioactive secondary metabolites produced by traditional ‘‘mutate-and-screen’’ method of strain improvement
microorganisms have been reported, and only 150 of them are that was early developed for the penicillin strain. Today, penicillin
being used in pharmacology, agriculture or other fields. Over producer Penicillium chrysogenum makes over 70 g/l of drug
10,000 of these compounds are produced by actinomycetes, starting from a strain capable of producing only 60 mg/l, which
representing 45% of all bioactive microbial metabolites discov- represents a 1000-fold increase. Other impressive examples of
ered, 80% if we only consider those compounds in practical use. strain improvement are the production of riboflavin by Ashbya
Among actinomycetes, around 7600 compounds are produced by gossypii that has been improved 40,000 times and the production
Streptomyces species (Bérdy, 2005). of a 100,000-fold excess of vitamin B12 by Pseudomonas
Drugs in commercial use are obtained at industrial scale either denitrificans (Demain, 2006).
by fermentative production, chemical synthesis or semisynthetic Current methods to increase the productivity of industrial
processes. When fermentation is used, it requires microbial microorganisms go from the classical random mutagenesis
strains producing high titers of compound. However, wild-type performed in close association with optimization of large-scale
strains isolated from nature usually produce only discrete industrial fermentations, to the use of more rational methods. One
amounts of a particular secondary metabolite, which in terms of of these is metabolic engineering where, in order to maximize
its isolation implies the need for production improvement to meet product yields, primary metabolic fluxes are redirected by
commercial requirements. This improvement will, in addition, the introduction of genetic modifications through recombinant
determine in most cases whether a new natural product goes to DNA technology, in a manner that supports high secondary
market or is abandoned, since it is clear that when production metabolite productivities (Adrio and Demain, 2006; Nielsen,
1998). In addition, the development of modern technologies such
as DNA sequencing, transcription profiling, genomics, proteomics,
 Corresponding author. Fax: +34 985103652. metabolomics, transcriptomics and metabolite profiling has
E-mail address: jasalas@uniovi.es (J.A. Salas). created new opportunities to engineer microorganisms for the

1096-7176/$ - see front matter & 2008 Elsevier Inc. All rights reserved.
doi:10.1016/j.ymben.2008.07.001
ARTICLE IN PRESS

282 C. Olano et al. / Metabolic Engineering 10 (2008) 281–292

production of natural products in high yields (Bro and Nielsen, increase of acetyl-CoA, precursor of undecylprodigiosin and
2004). actinorhodin, increasing production of these antibiotics (Fig. 2)
In this review, we focus on the different genetic approaches (Butler et al., 2002).
used for the production improvement of secondary metabolites
produced by actinomycetes. In general, a particular metabolite
2.2. Fatty acid precursors
can be overproduced following different approaches: (i) altering
the metabolic flux distribution of its different precursors,
Engineering the availability of coenzyme A (CoA) activated
(ii) deregulating its specific biosynthetic pathway, (iii) increasing
fatty acid precursors has been reported for the enhanced
self-resistance or inducing resistance to several antibiotics,
production of several polyketides such as erythromycin, oligomy-
(iv) overexpressing structural genes coding for enzymes involved
cin, monensin B and actinorhodin in the producer organisms. In
in the biosynthesis of the metabolite, (v) using global genetic
the case of erythromycin, two of its precursors are malonyl-CoA
approaches such as genome shuffling and (vi) by expressing the
derived from the carboxylation of acetyl-CoA and methylmalonyl-
biosynthetic gene cluster in a heterologous host or an industrially
CoA, which can be synthesized through different pathways such
optimized strain. Examples of each of these approaches are
as carboxylation of propionyl-CoA and rearrangement of succinyl-
described below, summarized in Table 1 and schematized in Fig. 1.
CoA (Fig. 3). Engineering the methylmalonyl-CoA metabolic node
in erythromycin producers Saccharopolyspora erythraea and Aero-
2. Precursor engineering microbium erythreum leads to the enhanced production of the
antibiotic depending on the fermentation medium used. Over-
The availability of biosynthetic precursors is a key factor production of erythromycin was achieved by inactivating the
determining the productivity of secondary metabolites. Primary methylmalonyl-CoA mutase (MCM) gene mutB and cultivation of
metabolism is the supplier for those precursors that are generally Sac. erythraea or A. erythreum in a carbohydrate-based medium
formed through the catabolism of various carbon substrates such while in an oil-based medium erythromycin production dimin-
as fatty acids, monosaccharides or proteins. The identification and ished (Reeves et al., 2004, 2006). In a carbohydrate-based
genetic manipulation of key enzymes regulating carbon flux medium, the MCM reaction acts like a drain on the methylma-
through the metabolic network of the central carbon metabolism lonyl-CoA pool, but in an oil-based medium, the same reaction
can lead to an increase in the availability of a particular precursor acts to fill the methylmalonyl-CoA pool (Reeves et al., 2006). In
as shown in Figs. 2 and 3. addition, overproduction of erythromycin was accomplished by
duplication of the MCM operon (mutA, mutB, meaB and mutR) and
cultivation of Sac. erythraea in an oil-based fermentation medium
2.1. Carbohydrate metabolism
(Reeves et al., 2007). These experiments led to the conclusion that
the carbon flow under oil-based growth conditions was from
In glucose catabolism the Embden–Meyerhof and pentose
succinyl-CoA to methylmalonyl-CoA (Reeves et al., 2007).
phosphate (PPP) pathways are interlinked to form the metabolic
Antiparasitic avermectins and cell growth inhibitor oligomycin
network shown in Fig. 2. Key enzymes in the individual pathways
are macrocyclic lactones produced by Streptomyces avermitilis
regulate the carbon flux among them. The first intermediate in
(Omura et al., 2001). During the biosynthesis of the avermectins,
glucose catabolism, glucose-6-phosphate (G6P), is used as a
two different starter units are used: isobutyryl-CoA and
common substrate for phosphoglucose isomerase (Pgi), glucose-
2-methylbutyryl-CoA produced by the degradation of branched-
6-phosphate dehydrogenase (Zwf), and phosphoglucomutase
chain amino acids valine and isoleucine. Inactivation of bkdF
(Pgm). Zwf channels glucose to the PPP. Pgm catalyzes the
encoding a branched-chain a-keto acid dehydrogenase leads to
reversible interconversion of G6P and glucose-1-phospate (G1P),
the abolition of avermectins production and in turn to the
potentially leading to the deviation of carbon to the biosynthesis
overproduction of the macrolide oligomycin by enabling addi-
of glucose-based polymers such as glycogen (Ryu et al., 2006) or
tional extender units (malonyl-CoA, methylmalonyl-CoA and
to the biosynthesis of 6-deoxyhexoses (6DOH) characteristic, but
ethylmalonyl-CoA) to enter the biosynthetic pathway of oligomy-
not exclusive, of secondary metabolism and found in many
cin (Cropp et al., 2001; Wei et al., 2006). In Streptomyces
bioactive secondary metabolites (Salas and Méndez, 2005). Pgi
cinnamonensis, inactivation of crotonyl-CoA reductase (CCR)
leads carbohydrate catabolism through the Embden–Meyerhof
involved in the reduction of crotonyl-CoA to butyryl-CoA leads
pathway and subsequently to the tricarboxylic acid cycle, where
to the depletion of ethylmalonyl-CoA synthesized from butyryl-
other precursors for secondary metabolite formation are provided.
CoA and subsequently to the accumulation of monensin B instead
Attempts of genetic manipulation of the initial steps in the
of monensin A. Monensin B requires only methylmalonyl-CoA
Embden–Meyerhof pathway and PPP have been reported for the
while monensin A requires both methylmalonyl-CoA and ethyl-
enhanced production of clavulanic acid, actinorhodin and un-
malonyl-CoA during its biosynthesis (Cropp et al., 2001). Actinor-
decylprodigiosin. The glycolytic pathway in Streptomyces clavuli-
hodin production by Streptomyces coelicolor is another example of
gerus was genetically engineered by disruption of gap1 and gap2
enhanced production of a polyketide by modification of its
genes coding for glyceraldehyde-3-phosphate dehydrogenases
precursor supplies. In this case, the overexpression of the genes
involved in the conversion of D-glyceraldehyde-3-phosphate
accA2, accB and accE, coding for the different subunits of the
(G3P) into 1,3-diphosphoglicerate (1,3-BPG). Since G3P and
enzyme acetyl-CoA carboxylase (ACC) in S. coelicolor, was
L-arginine are precursors in the biosynthesis of clavulanic acid,
sufficient to enhance carbon flux to malonyl-CoA, which is a
the accumulation of G3P in the gap mutants correlated with a
precursor of actinorhodin together with acetyl-CoA, leading to a
significant increase in the production of the antibiotic, increase
six-fold increase in actinorhodin production (Ryu et al., 2006).
that reached 3.1-fold when L-arginine was fed to the cultures
(Li and Townsend, 2006). PPP was engineered by independently
removing two different sets of genes: zwf1 and zwf2 coding for 2.3. Cofactors
isoenzymes of glucose-6-phosphate dehydrogenase, and devB
coding for a 6-phosphoglucolactonase in Streptomyces lividans. Additional elements of the central carbon metabolism take
These deletions lead to channeling the precursors flux through the part in the biosynthesis of secondary metabolites as precursors
Embden–Meyerhof instead of PPP, which in turn lead to the and are also targets for metabolic engineering approaches. This is
ARTICLE IN PRESS

C. Olano et al. / Metabolic Engineering 10 (2008) 281–292 283

Table 1
Increase in production of secondary metabolites produced by actinomycetes achieved by metabolic engineering

Compound Strain Engineering approach Increase (fold)

Actinomycin S. antibioticus Ribosome engineering 5.25

Actinorhodin S. coelicolor Fatty acid precursors 6


Up-regulation 2.6–40
Ribosome engineering 1.6–180

S. lividans Carbohydrate metabolism 4–5


Cofactors 4
Up-regulation 470
Down-regulation 3.5–5

C-1027 S. globisporus Biosynthetic structural genes 2–4


Cephamycin C S. clavuligerus Biosynthetic structural genes 2–4
Up-regulation 2–3
Chromomycin S. griseus Down-regulation 3
Clavulanic acid S. clavuligerus Carbohydrate metabolism 2.1–3.1
Biosynthetic structural genes 1.6–5
Up-regulation 1.5–23.8
Clorobiocin S. coelicolor Heterologous expression 1
Daunorubicin S. peucetius Biosynthetic structural genes 8–9
Up-regulation 2.4–10
6-dEB S. coelicolor Heterologous expression and fatty acid precursors 4
E. coli Heterologous expression and fatty acid precursors 1.8
Desosaminyl tylactone S. venezuelae Heterologous expression, PKS deletion and up-regulation 17.1
Doramectin S. avermitilis Biosynthetic structural genes 4–23
Doxorubicin S. peucetius Biosynthetic structural genes 3–74
Up-regulation 4
Erythromycin A. erythreum Fatty acid precursors 2–4
Sac. erythraea Fatty acid precursors 1.25–1.5
Expression of heterologous genes 2–2.5
Plasmid integration 2–5
Expression in industrial strains 50
Fredericamycin S. chattanoogensis Ribosome engineering 26
Formycin S. lavendulae Ribosome engineering 5.2
GE2270 P. rosea Ribosome engineering 1.8
Hydroxycitric acid Streptomyces U121 Genome shuffling 5
Kanamycin S. kanamyceticus Self-resistance 3.5
Megalomycin Sac. erythraea Heterologous expression and 6DOH metabolism 3.4
15-Methyl-6-dEB S. coelicolor Heterologous expression and plasmid co-integration 4–25
Mithramycin S. argillaceus Up-regulation 2–16
Monensin B S. cinnamonensis Fatty acid precursors 1.76
Nanchangmycin S. nanchangensis Biosynthetic structural genes 3
Neomycin S. fradiae Self-resistance 6

Nikkomycin X S. ansochromogenes Biosynthetic structural genes 1.8–2


Up-regulation 2

Novclobiocin 122 S. coelicolor Heterologous expression and 6DOH metabolism 8–26

Novobiocin S. coelicolor Heterologous expression 1


S. coelicolor Heterologous expression and up-regulation 3

Nystatin S. noursei Up-regulation 3.25


Biosynthetic structural genes 1.6

Oligomycin A S. avermitilis Fatty acid precursors 23


Pikromycin S. venezuelae Up-regulation 1.6–2.6

Pimaricin S. nataliensis Up-regulation 2.4


Down-regulation 1.8

Pristinamycin IIA S. pristinaespiralis Biosynthetic structural genes 1.25


Rapamycin S. hygroscopicus Up-regulation 1.2–1.4
e-Rhodomycinone S. peucetius Up-regulation 7–100
Salinomycin S. albus Ribosome engineering 1.5–2.3
Shengjimycin S. spiramyceticus Biosynthetic structural genes 2
Spinosyn Sac. spinosa Carbohydrate metabolism 3
Tetracenomycin D3 S. glauscescens Biosynthetic structural genes 20–30
Tylactone S. venezuelae Heterologous expression, PKS deletion and up-regulation 2.7
Tylosin S. fradiae Up-regulation 1.2–4.9
Down-regulation 1.5
Genome shuffling 6–8
Undecylprodigiosin S. coelicolor Up-regulation 31
S. lividans Carbohydrate metabolism 4
Down-regulation 11–12
Ribosome engineering 1.9–2.9
ARTICLE IN PRESS

284 C. Olano et al. / Metabolic Engineering 10 (2008) 281–292

the case of the cofactor S-adenosyl-methionine (SAM). Actinorho-


din production was reported to be enhanced both in S. lividans and
in S. coelicolor by overexpression of Streptomyces spectabilis metK
gene coding for S-adenosyl-L-methionine synthetase that cata-
lyzes the biosynthesis of SAM from ATP and L-methionine. The
same effect was obtained by addition of SAM to the culture
medium. The increase in actinorhodin production is the conse-
quence of inducing the expression of pathway-specific transcrip-
tional activator actII-orf4 (Kim et al., 2003; Okamoto et al., 2003).
A similar effect on the production of pikromycin in Streptomyces
venezuelae was reported by expression of metK1-sp gene from
Streptomyces peucetius. In this case an increase in transcripts of the
pathway-specific transcriptional activator pikD and a keto-
synthase gene were observed (Maharjan et al., 2008).

3. Engineering regulatory networks

Genes for the biosynthesis of secondary metabolism pathways


are commonly grouped together in clusters on the chromosome
including their pathway-specific regulatory genes. Pathway-speci-
fic regulators can have either positive (activators) or negative
Fig. 1. Scheme representing the different approaches used for improvement of (repressors) effects on the expression of gene cluster elements.
secondary metabolite production. There are clusters containing different number of pathway-specific

Fig. 2. Glucose catabolism and engineered steps involved in the biosynthesis of several secondary metabolites produced by actinomycetes. 1,3-BPG, 1,3-
diphosphoglycerate; Cas2, clavaminate synthase; Cvm1, enzyme involved in the biosynthesis of antipodal clavams; DevB, phosphoglucolactonase; 6DOH, 6-deoxyhexose
pathways; F6P, fructose-6-phosphate; Gap1 and Gap2, glyceraldehyde-3-phosphate dehydrogenases; Gdh, NDP-glucose dehydratase; G1P, glucose-1-phosphate; G6P,
glucose-6-phosphate; G3P, glyceraldehyde-3-phosphate; Gtt, NDP-glucose synthase; LAT, lysine e-aminotransferase; MetK, S-adenosyl-L-methionine synthetase; Pgi,
phosphoglucose isomerase; 6PG, 6-phosphogluconate; 6PGL, 6-phosphoglucolactone; Pgm, phosphoglucomutase; Pah, proclavaminate amidino hydrolase; PPP, pentose
phosphate pathway; SAM, S-adenosyl-L-methionine; SanO, non-ribosomal peptide synthetase; SanU and SanV, glutamate mutase; Zwf1 and Zwf2, glucose-6-phosphate
dehydrogenases.
ARTICLE IN PRESS

C. Olano et al. / Metabolic Engineering 10 (2008) 281–292 285

Fig. 3. Fatty acid precursors and engineered steps involved in the biosynthesis of several secondary metabolites produced by actinomycetes. ACC, acetyl-CoA carboxylase;
BkdF, branched-chain a-keto acid dehydrogenase; ICM, isobutyryl-CoA mutase; IST, 400 -O-acyltransferase; MMT, methylmalonyl-CoA transcarboxylase; MutB,
methylmalonyl-CoA mutase; PPC, propionyl-CoA carboxylase.

positive regulatory genes ranging from one in actinorhodin path- Protein family (SARP) characterized by the presence of a winged
way (Fernández-Moreno et al., 1991) to three in daunorubicin helix-turn-helix (HTH) motif towards the N-termini (Wietzorrek
pathway (Stutzman-Engwall et al., 1992; Otten et al., 1995, 2000). and Bibb, 1997). Overexpression of SARP positive regulators has
In addition, some clusters contain both activators and repressors been reported to increase the production of different secondary
such as in the tylosin pathway that contains two activator and two metabolites such as actinorhodin and undecylprodigiosin in S.
repressor-coding genes (Stratigopoulos et al., 2004). However, in coelicolor by actII-orf4 and redD (Narva and Feitelson, 1990;
some pathways no regulatory genes have been identified, as is the Fernández-Moreno et al., 1991), undecylprodigiosin in S. lividans
case of the erythromycin pathway (Rawlings, 2001). Moreover, and S. parvulus by redD (Malpartida et al., 1990), nikkomycin in S.
other regulatory genes generally located outside the biosynthetic ansochromogenes by sanG (Liu et al., 2005) and clavulanic acid in S.
gene cluster may play a regulatory role in the cluster, in many cases clavuligerus by ccaR (Pérez-Llarena et al., 1997; Hung et al., 2007).
showing pleiotropic effects on the production of multiple second- The production of other secondary metabolites such as tylosin,
ary metabolites. The best-known example is S. coelicolor that daunorubicin and mithramycin was reported to be up-regulated
produces several antibiotics (actinorhodin, calcium-dependent either using SARP activator coding genes or, in addition, other
antibiotic, undecylprodigiosin and methylenomycin) and where pathway-specific positive regulators. The production of tylosin by
the onset of their biosynthesis is controlled by specific regulators Streptomyces fradiae was reported to be increased by overexpres-
(actII-orf4, cdaR, redD and redZ), while there are several pleiotropic sion of tylS or tylR encoding a SARP regulator and a protein
genes (i.e. afs, abs and bld) affecting antibiotic production and, in containing a transposase domain, respectively. In both cases an
addition, the morphological development of the bacteria (Huang et increase in tylosin production was obtained in the wild-type
al., 2005). Taking into account all previous observations it seems strain and in a tylosin overproducing strain (Stratigopoulos et al.,
obvious that deregulation of the expression of secondary metabo- 2004). Overexpression of locus dnrR1, encoding SARP regulator
lite pathways, by overexpression of pathway-specific positive DnrI, or dnrR2, encoding positive regulators DnrN and DnrO
regulators or by inactivation of pathway repressors, is the most containing LuxR and ArsR type HTH motifs, led to a 10–100-fold
intuitive approach for the improvement of their production (Fig. 1). increase in the production of e-rhodomycinone and 25-fold
increase in daunorubicin (Stutzman-Engwall et al., 1992; Otten
et al., 1995). Mithramycin production by Streptomyces argillaceus
3.1. Up-regulation has been improved by overexpression of two different regulatory
genes in multicopy plasmids: mtrY encoding a protein with a
The vast majority of the pathway-specific activators in PadR-like HTH motif (Garcı́a-Bernardo et al., 2000) and mtmR
actinomycetes belong to the Streptomyces Antibiotic Regulatory encoding a SARP family regulatory protein (Lombó et al., 1999;
ARTICLE IN PRESS

286 C. Olano et al. / Metabolic Engineering 10 (2008) 281–292

Wohlert et al., 1999). In addition, mtmR was also able to activate of the pathway-specific regulator actII-orf4 mRNA increased
actinorhodin biosynthesis and complement actII-orf4 mutation in (Li et al., 2006). Other modifications, such as the inactivation of
S. coelicolor JF1 (Lombó et al., 1999; Blanco et al., 2000). Other polyphosphate kinase gene ppk, have been reported to induce the
secondary metabolite biosynthetic pathways lacking SARP activa- expression of actII-orf4 activating the actinorhodin pathway
tors have been up-regulated using other kind of pathway-specific (Chouayekh and Virolle, 2002).
regulators such as pimM encoding for a PAS/LuxR regulator that
enhance pimaricin production in Streptomyces nataliensis (Antón
et al., 2007) or rapH and rapG, encoding proteins with LuxR and 4. Engineering antibiotic resistance
AraC-like HTH motifs respectively, that increase rapamycin
production in Streptomyces hygroscopicus (Kuscer et al., 2007). Antibiotic biosynthetic gene clusters typically include one or
SARP family proteins generally appear to function as pathway- more genes encoding different resistance mechanisms for self-
specific regulators but, in some cases, they can work as pleiotropic protection to overcome the toxic effects of their products. These
regulatory proteins that control the production of multiple systems include enzymes to modify the antibiotic target site,
secondary metabolites and morphological differentiation. Such antibiotic inactivating enzymes and transport systems (Cundliffe,
is the case of the afsR gene from S. coelicolor able to increase 1989; Méndez and Salas, 2001). In some cases these resistance
actinorhodin and undecylprodigiosin production through its systems are involved not only in self-protection but also in
overexpression in S. lividans (Horinouchi et al., 1983). Others antibiotic biosynthesis (Olano et al., 1995; Menéndez et al., 2007).
SARP pleiotropic activators homologue to AfsR such as SsmA and Consequently, it is not surprising that increased antibiotic
AfsR-p from Streptomyces noursei and S. peucetius have been resistance has often been used to select for mutants with
shown to enhance nystatin and doxorubicin biosynthesis, respec- increased levels of antibiotic production (Yanai et al., 2006).
tively (Sekurova et al., 1999; Parajuli et al., 2005). In addition,
overexpression of afsR-p in S. lividans, S. clavuligerus, S. griseus and 4.1. Ribosome engineering
S. venezuelae leads to overproduction of actinorhodin, clavulanic
acid, streptomycin and pikromycin, respectively (Parajuli et al., A dramatic activation of antibiotic production was observed in
2005; Maharjan et al., 2008). Different pleiotropic activators have S. lividans and S. coelicolor containing a mutation in rpsL gene,
been shown to increase production of actinorhodin such as afsS encoding the ribosomal protein S12, which confers resistance to
(Vögtli et al., 1994), the two-component regulatory system afsQ1 streptomycin (Shima et al., 1996). This led to rationally improve
and afsQ2 (Ishizuka et al., 1992), abaA (Fernández-Moreno et al., the production of actinorhodin and undecylprodigiosin by
1992) and ptpA (Umeyama et al., 1996). The last gene, pptA, introducing point mutations in chromosomal rpsL gene or by
encodes a phosphotyrosine protein phosphatase presumably a using a single-copy-number plasmid to express rpsL mutant
member of the signal transduction network that controls the versions (Shima et al., 1996; Okamoto-Hosoya et al., 2003). Since
production of actinorhodin and undecylprodigiosin (Umeyama then, ribosome engineering has been used as a rational approach
et al., 1996). to enhance antibiotic production in different Streptomyces spp. by
conferring resistance to several antibiotics mediated by mutant
ribosomes (Fig. 1). The beneficial effect on antibiotic production of
3.2. Down-regulation mutations in rpsL gene conferring high-level resistance to
streptomycin have been confirmed by the isolation of mutants
The inactivation of pathway-specific or pleiotropic repressors in other Streptomyces spp. such as S. chattanoogensis, S. anti-
can also lead to overproduction of secondary metabolites. This is bioticus, S. lavendulae and S. albus producers of fredericamycin,
the case of chromomycin that is overproduced when a pathway- actinomycin, formycin and salinomycin, respectively (Hosoya
specific transcriptional repressor cmmRII is inactivated in S. griseus et al., 1998; Tamehiro et al., 2003). Overexpression in S. coelicolor
subsp. griseus (Menéndez et al., 2007). A similar effect was of genes such as frr encoding a ribosome-recycling factor whose
accomplished in tylosin biosynthesis by disruption of tylP, gene overproduction has been detected in streptomycin resistant rpsL
that encode a g-butyrolactone receptor that negatively affects the mutants, also leads to actinorhodin overproduction up to 10-fold
production of the antibiotic and, in addition, influences morpho- (Hosaka et al., 2006). On the other hand, low-level resistance to
logical differentiation in S. fradiae (Stratigopoulos et al., 2002). An streptomycin by mutation or deletion of rsmG gene, coding for a
additional pathway-specific transcriptional repressor, tylQ, is SAM-dependent 16S rRNA methyltransferase, has been reported to
present in tylosin gene cluster but, in this case, its inactivation raise actinorhodin production in S. coelicolor (Nishimura et al.,
does not produce an increase in antibiotic production but to an 2007).
early onset of tylosin production (Stratigopoulos and Cundliffe, Apart from mutations conferring resistance to streptomycin,
2002). In actinorhodin gene cluster there is also a gene, actVB- enhancement of antibiotic production can be also achieved by
orf10, encoding a LysR-type transcriptional regulator that has been mutations conferring resistance to other antibiotics such as
used for antibiotic overproduction through its inactivation in gentamicin, rifampin, paromomycin, geneticin, fusidic acid,
S. lividans (Martı́nez-Costa et al., 1999). thiostrepton and lincomycin. The simultaneous introduction of
A well-known pleiotropic repressor system related with several resistant mutations to streptomycin, gentamicin and
phosphate regulation of secondary metabolite production is the rifampin has a cumulative effect on antibiotic production leading
two-component phoR–phoP system. Disruption of phoR or simul- to improvements in actinorhodin, salinomycin and thiazolylpep-
taneous deletion of both phoR and phoP has been recently shown tide GE2270 production in S. coelicolor, S. albus and Planobispora
to increase pimaricin production in S. nataliensis (Mendes et al., rosea strains, respectively (Hu and Ochi, 2001; Tamehiro et al.,
2007). Deletion of the same system in S. lividans was reported to 2003; Beltrametti et al., 2006). In addition, an increase in
boost actinorhodin and undecylprodigiosin production, 5- and actinorhodin production up to 180-fold has been recently
12-fold increase, respectively (Sola-Landa et al., 2003). In described in a S. coelicolor strain resistant to seven of the eight
S. coelicolor there is an additional pleiotropic gene, nsdA, that antibiotics mentioned above, depending on the culture medium
negatively affects antibiotic production and whose disruption used (Wang et al., 2008). These mutations produce ribosomes
leads to an increase in actinorhodin, calcium-dependent antibiotic with aberrant protein and hyperphosphorylated guanosine
and methylenomycin biosynthesis. In the nsdA mutant the levels nucleotide (ppGpp) synthesis activities (Okamoto-Hosoya et al.,
ARTICLE IN PRESS

C. Olano et al. / Metabolic Engineering 10 (2008) 281–292 287

2003; Wang et al., 2008). There is a positive correlation between belonging to 6DOH pathways are good candidates for this kind
ppGpp and antibiotic biosynthesis since ppGpp triggers bacterial of manipulation. This was the case of gene sgcA1 encoding a NDP-
secondary metabolism when cells enter into stationary phase glucose synthase, involved in the biosynthesis of 6DOH 4-deoxy-
(Chakraburtty and Bibb, 1997). On the other hand, in all these 4-(dimethylamino)-5,5-dimethyl-D-ribopyranose, that has been
mutant strains overproduction of actinorhodin correlates to used to engineer the production of the enediyne antitumor
higher expression of the actII-orf4 gene probably due to the antibiotic C-1027 in S. globisporus. The overexpression of sgcA1
enhanced protein synthesis (Hu and Ochi, 2001; Okamoto-Hosoya alone resulted in a two-fold improvement for C-1027 production,
et al., 2003; Wang et al., 2008). and up to four-fold if cagA gene coding for C-1027 apoprotein was
coexpressed (Murrell et al., 2004). It is noteworthy that genes
4.2. Self-resistance improvement coding for enzymes involved in the biosynthesis of 6DOH, such as
sgcA1, are usually clustered together with the rest of secondary
metabolite biosynthetic genes involved in the biosynthesis of the
Increasing self-resistance levels in producing organisms has
aglycone (Salas and Méndez, 2005). However, there are a few
been also used as an approach for increasing production yields.
examples where genes involved in the biosynthesis of 6DOH are
This strategy was used in the producers of two aminoglycosides,
found outside of the biosynthetic gene cluster and shared
kanamycin and neomycin, and consisted in the introduction
between primary and secondary metabolism. This is the case for
in the producer strains of the gene encoding an aminoglycoside
genes involved in the biosynthesis of L-rhamnose usually found
60 -N-acetyltransferase derived from Streptomyces kanamyceticus
outside of the cluster in all cases reported (Madduri et al., 2001;
that confers resistance to aminoglycoside antibiotics. This re-
Gullón et al., 2006; Luzhetskyy et al., 2007; Ramos et al., 2008).
sulted in a substantial increase in production of these antibiotics,
The duplication of L-rhamnose biosynthetic genes gtt and gdh
especially neomycin (Crameri and Davies, 1986). Similar results
encoding NDP-glucose synthase and NDP-glucose dehydratase
were described in several antibiotic overproducing organisms
respectively, have been used for the improvement of spinosyn
such as S. aureofaciens, 6-demethylchlortetracycline producer,
production in Sac. spinosa. These enzymes are the two first
through the overexpression of a self-defense gene involved in
activities responsible for channeling G1P to the biosynthesis of L-
drug efflux (Dairi et al., 1995), or the S. kanamyceticus industrial
rhamnose and L-dimethyl-forosamine, 6DOHs present in spinosyn
kanamycin producer strain where duplication of the entire
(Madduri et al., 2001).
kanamycin gene cluster, including its three self-resistance genes,
Other primary metabolism precursors such as G3P, lysine or
leads to enhanced kanamycin resistance in addition to antibiotic
2-oxoglutarate can be channeled to enhance the production of
overproduction (Yanai et al., 2006).
several secondary metabolites by using specific biosynthetic
structural genes (Fig. 2). Overexpression or integration into the
chromosome of clavaminate synthase gene cas2, resulted in up to
5. Engineering biosynthetic structural genes
five-fold increase in clavulanic acid production in S. clavuligerus.
An additional improvement, up to 23-fold, was achieved by the
Usually structural genes involved in the biosynthesis of
simultaneous integration into the chromosome of pathway-
secondary metabolites have been targeted for the generation of
specific activator ccaR and the clavaminate synthase gene cas2
new compounds through their inactivation or deletion. In
(Hung et al., 2007). In addition, the duplication of pah2 gene,
addition, these mutants can be used as hosts for the expression
encoding a proclavaminate amidino hydrolase, has been recently
of heterologous genes that can lead to the generation of new
reported to improve clavulanic acid production (Song et al., 2008).
compounds. There are a number of examples in the literature
By expressing lat gene, encoding a lysine e-aminotransferase, in a
where gene dose alteration, modification and heterologous
high-copy-number plasmid in S. clavuligerus, cephamycin C
expression of biosynthetic structural genes have been used to
production was greatly enhanced (Malmberg et al., 1993, 1995).
boost production of secondary metabolites (Fig. 1).
A similar approach has been used in order to increase nikkomycin
X production in S. ansochromogenes by the overexpression of sanU
5.1. Gene dose increase and sanV genes coding for the two subunits of a glutamate mutase
involved in the conversion of glutamate into nikkomycin pre-
S. pristinaespiralis is the producer of pristinamycin II, strepto- cursor 3-methylaspartate (Li et al., 2005). Duplication of sanO
gramin antibiotic produced as a mixture of two compounds PIIA gene coding for a non-ribosomal peptide synthetase (NRPS)
and PIIB in a ratio 80:20. The complete conversion of compound responsible for the formation of the 4-formyl-4-imidazolin-2-
PIIB into PIIA was achieved by integration into the chromosome of one moiety present in nikkomycin X leads to an equivalent
an additional copy of snaA and snaB genes coding for a increment in antibiotic production (Wang and Tan, 2004).
heterodimeric monooxygenase catalyzing this conversion. The
duplication of snaA/snaB gene dosage did not increase the total
amount of pristinamycins but enhanced the ratio of pristinamycin 5.2. Gene inactivation or deletion
PIIA leading to a better production of this compound (Sezonov
et al., 1997). A similar approach has been used to boost the A different approach for improving metabolite production is to
production of tetracenomycin D3, intermediate in the biosynth- remove genes coding for activities that transform the metabolite
esis of tetracenomycin C by Streptomyces glaucescens. In this case into a different one. Such is the case of dnrX and dnrH genes
overexpression of tcmM, coding for an acyl carrier protein (ACP) of identified in S. peucetius and involved in the transformation of
a type II polyketide synthase (PKS), cloned into a high copy vector daunorubicin and doxorubicin into their polyglycosylated forms
leads to increase up to 30-fold the production of this biosynthesis known as baumycins. Inactivation of dnrX and dnrH represent
intermediate (Decker et al., 1994). improvements of 3- and 8.5-fold in the biosynthesis of doxor-
Carbon flux from glucose can be channeled to enhance the ubicin and daunorubicin, respectively (Lomovskaya et al., 1998;
production of secondary metabolites by manipulating glucose Scotti and Hutchinson, 1996). In addition, disruption of dnrU,
catabolism as showed in Section 2.1 and, in addition, by altering involved in the transformation of daunorubicin into 1,3-dihydro-
the expression of structural genes specifically involved in the daunorubicin, leads to a similar improvement of doxorubicin
biosynthesis of a particular metabolite. Among these, genes production (Lomovskaya et al., 1999). Production of doxorubicin
ARTICLE IN PRESS

288 C. Olano et al. / Metabolic Engineering 10 (2008) 281–292

was further improved by disruption of several genes in the same erythraea industrial strain by a chromosomally integrated copy of
strain, which lead to increases up to seven-fold in the dnrU/dnrX a bacterial hemoglobin gene vhb, originally isolated from
double mutant and to 26-fold in the dnrU/dnrX/dnrH triple Vitreoscilla spp. (Brünker et al., 1998; Minas et al., 1998). This
mutant. An additional 1.3–2.8-fold raising was obtained by system was previously used for the enhancement of cephalospor-
overexpressing dnrV and doxA genes, involved in late steps during in C in the fungus Acremonium chrysogenum (DeModena et al.,
doxorubicin biosynthesis, in the double or triple mutants 1993). Improvement of erythromycin production in that strain
(Lomovskaya et al., 1999). Following a similar approach, over- may be the consequence of an increase in erythromycin
expressing dnmT, involved in the biosynthesis of daunorubicin biosynthetic flux as a result of the increased activity of an
deoxysugar L-daunosamine, in the dnrH mutant led to an oxygen-dependent step in erythromycin synthesis, most likely
improvement in daunorubicin production (Scotti and Hutchinson, hydroxylation steps (Brünker et al., 1998). However, it is uncertain
1996). how the presence of the oxygen-binding heme protein affects
Channeling of carbon flux by modification of amino acid- erythromycin production (Minas et al., 1998).
related precursor supplies has been applied to increase clavulanic
acid production in S. clavuligerus, strain that also produces
cephamycin C and other clavams. Production of cephamycin C
6. Genome shuffling
can be abolished by inactivation of lat gene coding for a lysine e-
aminotransferase. This inactivation leads in turn to an increase in
Genome shuffling has been described and demonstrated
clavulanic acid production (Paradkar et al., 2001). A similar effect
as a new method for rapid enhancement of secondary metabolite
can be achieved if production of non-clavulanic acid clavams is
production. Using this approach six- to eight-fold increase in
depleted by inactivation of cvm1 (Mosher et al., 1999). Both
tylosin production in S. fradiae were obtained by two rounds of
approaches, inactivation of lat and cvm1 applied at the same
genome shuffling over a population of classically improved
organism (Fig. 2), in this case an industrial strain of S. clavuligerus,
strains. Similar titers were obtained by classical improvement
resulted in a significant improvement of clavulanic acid produc-
methods but along 20 rounds of mutagenesis and screening
tion, several orders of magnitude higher than in the wild-type
(Zhang et al., 2002). Following the same method Streptomyces spp.
strain (Paradkar et al., 2001).
U121 has been engineered for the overproduction of (2S,3R)-
Deletion of biosynthetic genes can lead occasionally to raise
hydroxycitric acid by three rounds of genome shuffling after
secondary metabolite production. Such is the case of nysF
one round of mutagenesis using nitrosoguanidine (Hida et al.,
inactivation that increases nystatin production in S. noursei. nysF
2007).
encodes a putative 40 -phosphopantetheinyl transferase that is
supposed to be involved in post-translational modification of the
ACP domains in nystatin type I PKS, but it was shown to negatively
regulate nystatin production (Volokhan et al., 2005). In the case of 7. Heterologous expression of entire gene clusters
nanchangmycin production by Streptomyces nanchangensis, this
can be improved by selective deletion of other PKS-containing Advances in developing DNA manipulation tools and the
clusters found in the same organism. Eight of these clusters were improvement in genome sequencing technologies have proved
found in S. nanchangensis and the inactivation of only one of them fruitful for the isolation of many gene clusters involved in natural
led to a three-fold increase in nanchangmycin yield. This deletion products biosynthesis. However, in some cases production titers
probably affects precursor supply for the rest of PKSs (Sun et al., of the encoded compound are low, and there are no genetic tools
2002). optimized for metabolic engineering in the particular producer
strain. The consequence of these problems is an increased interest
5.3. Gene modification for transferring secondary metabolite pathways to new hosts
where genetic tools are available or they have been previously
engineered for the heterologous production of bioactive com-
Doramectin (also named CHC-B1) is an antihelmintic polyke-
pounds.
tide compound produced by a S. avermitilis mutant strain, which
lacks a branched-chain a-keto acid dehydrogenase activity coded
by bkdF. This strain produces two related compounds, doramectin
7.1. Plasmids for heterologous expression
and CHC-B2. The conversion of CHC-B2 into doramectin is a
process that remains to be clarified but it involves in some way
Heterologous production of secondary metabolites in S.
aveC, a gene with unknown function. Improvement of production
coelicolor can be significantly enhanced by using specific plasmids
ratios for doramectin was achieved by generating several modified
designed for that purpose. This is the case of plasmid pSMALL that
versions of aveC by site-directed mutagenesis and error-prone
was obtained by co-integration of plasmids pJRJ2 and SCP2@ both
PCR, screening for the best variants and chromosomal insertion of
derived from plasmid SCP2*. The utilization of pSMALL for the
the refined gene in the producer strain where the wild-type aveC
production of 15-methyl-6-deoxyerythronolide B (15-methyl-6-
gene was previously inactivated (Stutzman-Engwall et al., 2003).
dEB) resulted in four-fold increase in both S. coelicolor and S.
Further enhancements of doramectin production were accom-
lividans, probably due to increased gene dosage and higher
plished by aveC semi-synthetic DNA shuffling and chromosomal
plasmid stability. This production can be further enhanced until
insertion of the best versions (Stutzman-Engwall et al., 2005).
25-fold if plasmid pBOOST is used for co-integration instead of
SCP2@ (Hu et al., 2003). Other plasmids have been developed for
5.4. Expression of heterologous genes its use in the erythromycin producer Sac. erythraea and other
actinomycetes. Such is the case of pCJR24, integrative vector
In S. spiramyceticus F21 the production of shengjimycin (also containing the pathway-specific activator actII-orf4 and promoters
named 400 -isovalerylspiramycin) was improved by inserting in its from actI and actIII, that lead to increments in erythromycin
chromosome a copy of the 400 -O-acyltransferase gene, ist, from S. production up to five-fold compared to wild-type strain after
mycarofaciens 1748 (Guangdong et al., 2001) (Fig. 3). On the other placing the entire PKS under the control of the actI promoter
hand, productivity of erythromycin has been improved in a Sac. (Rowe et al., 1998).
ARTICLE IN PRESS

C. Olano et al. / Metabolic Engineering 10 (2008) 281–292 289

7.2. Using Streptomyces hosts CoA from exogenous malonate and methylmalonate, in a S.
coelicolor strain expressing erythromycin PKS from Sac. erythraea
In some cases heterologous expression has being used as a (Lombó et al., 2001).
genetic tool for the identification of a particular gene cluster that
resulted in the improvement of the production yields. This was 7.3. Using industrially optimized strains
the case of the anthracycline tetracenomycin C produced by S.
glaucescens and the indolocarbazole rebeccamycin produced by
A further increase in production can be achieved by expression
Lechevaleria (formerly Saccharothrix) aerocolonigenes. Tetraceno-
of the newly engineered pathways into industrially optimized
mycin C cluster was identified by its expression in S. lividans,
strains. Using this approach, high erythromycin titers were
which resulted in the overproduction of pigmented intermediates
obtained in the industrial Sac. erythraea K41-135 strain after
of the biosynthetic pathway (Motamedi and Hutchinson, 1987),
expressing a genetically modified PKS. This is possible because it
while rebeccamycin gene cluster was identified through the
has been demonstrated that the overproduction phenotype is due
expression of L. aerocolonigenes DNA containing cosmids into S.
to mutations in non-PKS genes (Rodrı́guez et al., 2003).
albus, which led to production levels several folds higher that
those of the original strain (Sánchez et al., 2002).
In other cases, a set of genes was heterologously expressed. An 7.4. Using other microorganisms as hosts
example is the production of megalomycin in Sac. erythraea.
Megalomycin was originally produced by Micromonospora mega- In the production of secondary metabolites not only important
lomicea and it is structurally very similar to erythromycin but is the final yield but also the fermentation process. Heterologous
differs in the presence of an additional deoxysugar, megosamine, expression of gene clusters in Escherichia coli to overproduce
attached at the C-6 hydroxyl group of the aglycone. Since the desired compounds would first reduce fermentation times and
production of megalomycin in M. megalomicea is quite poor, the would allow industrial scale fermentations since there are well-
expression of the megosamine pathway in Sac. erythraea was established scalable protocols. In addition, the genetic tools for
attempted. This led to a significant increment in megalomycin engineering E. coli strains are highly developed. Following these
production and opened the possibility for further improvement criteria during the last years important efforts have been made for
both in production and in new compound development (Volche- the production of erythromycin in E. coli. These efforts started by
gursky et al., 2000). In a similar way, production of the tylosin expressing erythromycin PKS in an E. coli strain genetically
aglycone tylactone and its glycosylated derivative desosaminyl modified to facilitate (i) post-translational PKS modifications by
tylactone was achieved by expressing the tylosin PKS genes into a expressing sfp phosphopantetheinyl transferase gene from Bacillus
S. venezuelae strain where the pikromycin PKS genes were subtilis and (ii) methylmalonyl-CoA production by expressing PCC
previously deleted to avoid competition for the acyl-CoA pre- genes (pccA and pccB) from S. coelicolor. The activity of the
cursors (Jung et al., 2006). Furthermore, 2.7- and 17.1-fold biotinylated subunit PccA was enhanced by overexpression of E.
increments in production of these compounds were accomplished coli birA biotin ligase gene. The resultant strain was able to
by introducing an additional copy of the pikromycin pathway- produce 6-dEB at levels equivalent to a Sac. erythaea industrial
specific positive regulator gene pikD (Jung et al., 2008). strain (Pfeifer et al., 2001). Further metabolic engineering of that
Expression of different secondary metabolite gene clusters in strain such as the expression of metK (Fig. 2) from S. spectabilis
heterologous hosts usually is accompanied by approaches pre- encoding a SAM synthetase, led to a two-fold improvement in
viously described in this review such as overexpression of genes 6-dEB production (Wang et al., 2007). All accumulated knowledge
involved in 6DOH metabolism, fatty acid precursors supply or has led to the final production of erythromycin C and D in E. coli by
pathway-specific regulatory genes. In the case of the aminocou- expression of erythromycin PKS and 17 additional genes from
marins novobiocin and clorobiocin produced by S. spheroides and megosamine pathway coding for enzymes involved in 6DOH
S. roseochromogenes, respectively, the entire gene clusters were biosynthesis and other tailoring modifications (Peirú et al., 2005).
expressed independently in S. coelicolor M512 leading to equiva- Similar approaches are underway for the production of polyke-
lent production titers of the respective compounds as in the tides in other heterologous hosts such as Saccharomyces cerevisiae,
original strains (Eustáquio et al., 2005a). However, in the where different pathways, propionyl-CoA ‘‘dependent and inde-
heterologous host these pathways can be more easily engineered pendent’’ for the production of methylmalonyl-CoA have been
than in their wild-type strains. Metabolic engineering of these introduced (Mutka et al., 2006).
pathways in S. coelicolor has led to enhance novobiocin production In addition to polyketides, other secondary metabolites such as
by introducing the regulatory gene novG in a multicopy plasmid non-ribosomal peptides are produced in E. coli. In these cases, the
leading to a three-fold overproduction (Eustáquio et al., 2005b) strain also has to be engineered by introducing sfp gene for post-
and to increase 8–26-fold the production of novclobiocin 122 by translational modification of the NRPS. In that way antitumor
the introduction of genes involved in the biosynthesis of echinomycin and its intermediate triostin A were produced in
L-rhamnose from the elloramycin pathway (Freitag et al., 2006a). E. coli by expressing the entire gene cluster isolated from
Production of novclobiocin 122 was attempted before in Streptomyces lasaliensis (Watanabe et al., 2006).
S. coelicolor expressing the novobiocin gene cluster, by inactivation
of methyltransferase gene cloU, but only small amounts of the
desired L-rhamnoside derivative were obtained (Freitag et al., 8. Conclusions
2006b).
A good example on how to engineer fatty acid precursor supply The future success of the pharmaceutical industry depends on
in order to enhance the production of a secondary metabolite in a the identification or development of new compounds with novel
heterologous host is the production of 6-deoxy-erythronolide B activities or directed to more specific targets. The rapidly growing
(6-dEB) in S. coelicolor. This compound is produced from amount of secondary metabolite gene clusters identified and
propionyl-CoA and methylmalonyl-CoA. Production titers of this characterized provides new genetic tools for the generation of
compound were increased four-fold by the heterologous expres- novel compounds by combinatorial biosynthesis. In addition,
sion of genes matB and matC genes from Rhizobium trifolii, these clusters contain elements that can be used to increase
involved in the generation of malonyl-CoA and methylmalonyl- the production yields. As shown in this review, significant
ARTICLE IN PRESS

290 C. Olano et al. / Metabolic Engineering 10 (2008) 281–292

improvement of secondary metabolite production has been Acremonium chrysogenum is improved by the intracellular expression of a
obtained during recent years applying different methods of bacterial hemoglobin. Biotechnology 11, 926–929.
Eustáquio, A.S., Gust, B., Galm, U., Li, S.M., Chater, K.F., Heide, L., 2005a.
genetic and metabolic engineering both of wild-type producer Heterologous expression of novobiocin and clorobiocin biosynthetic gene
organisms and heterologous host after expression of biosynthetic clusters. Appl. Environ. Microbiol. 71, 2452–2459.
gene clusters. These genetic methods include the deletion and Eustáquio, A.S., Li, S.M., Heide, L., 2005b. NovG, a DNA-binding protein acting as a
positive regulator of novobiocin biosynthesis. Microbiology 151, 1949–1961.
overexpression of genes directed to up- or down-regulate the Fernández-Moreno, M.A., Caballero, J.L., Hopwood, D.A., Malpartida, F., 1991.
expression of secondary metabolite gene clusters, or to override The act cluster contains regulatory and antibiotic export genes, direct targets
bottlenecks in the biosynthesis of these compounds. Particular for translational control by the bldA tRNA gene of Streptomyces. Cell 66,
769–780.
attention has been drawn to alter carbon flux by redirecting the
Fernández-Moreno, M.A., Martı́n-Triana, A.J., Martı́nez, E., Niemi, J., Kieser, H.M.,
central primary metabolic networks toward specific parts of the Hopwood, D.A., Malpartida, F., 1992. abaA, a new pleiotropic regulatory locus
metabolism. Following these approaches, further improvements for antibiotic production in Streptomyces coelicolor. J. Bacteriol. 174, 2958–2967.
in the production of the compounds reviewed here and new ones Freitag, A., Li, S.M., Heide, L., 2006a. Biosynthesis of the unusual 5,5-gem-dimethyl-
deoxysugar noviose: investigation of the C-methyltransferase gene cloU.
should be expected in the future, with special attention to the Microbiology 152, 2433–2442.
efforts in the development of heterologous hosts for the expres- Freitag, A., Méndez, C., Salas, J.A., Kammerer, B., Li, S.M., Heide, L., 2006b. Metabolic
sion of secondary metabolites. engineering of the heterologous production of clorobiocin derivatives and
elloramycin in Streptomyces coelicolor M512. Metab. Eng. 8, 653–661.
Garcı́a-Bernardo, J., Braña, A.F., Méndez, C., Salas, J.A., 2000. Insertional inactivation
of mtrX and mtrY genes from the mithramycin gene cluster affects production
Acknowledgments and growth of the producer organism Streptomyces argillaceus. FEMS Microbiol.
Lett. 186, 61–65.
Guangdong, S., Jianlu, D., Yiguang, W., 2001. Construction and physiological studies
Research in the authors’ laboratory has been supported by on a stable bioengineered strain of shengjimycin. J. Antibiot. 54, 66–73.
grants from the Spanish Ministry of Education and Science Gullón, S., Olano, C., Abdelfattah, M.S., Braña, A.F., Rohr, J., Méndez, C., Salas, J.A.,
(BFU2006-00404 to J.A.S. and BIO2005-04115 to C.M.), the Red 2006. Isolation, characterization, and heterologous expression of the biosynth-
esis gene cluster for the antitumor anthracycline steffimycin. Appl. Environ.
Temática de Investigación Cooperativa de Centros de Cáncer to Microbiol. 72, 4172–4183.
J.A.S. (Ministry of Health, Spain; ISCIII-RETIC RD06/0020/0026) Hida, H., Yamada, T., Yamada, Y., 2007. Genome shuffling of Streptomyces sp. U121
and from the UE FP6 (Integrated Project no. 005224). We thank for improved production of hydroxycitric acid. Appl. Microbiol. Biotechnol. 73,
1387–1393.
Obra Social Cajastur for financial support to Carlos Olano and
Horinouchi, S., Hara, O., Beppu, T., 1983. Cloning of a pleiotropic gene that
Felipe Lombó. positively controls biosynthesis of A-factor, actinorhodin, and prodigiosin in
Streptomyces coelicolor A3(2) and Streptomyces lividans. J. Bacteriol. 155,
1238–1248.
References
Hosaka, T., Xu, J., Ochi, K., 2006. Increased expression of ribosome recycling factor
is responsible for the enhanced protein synthesis during the late growth phase
Adrio, J.L., Demain, A.L., 2006. Genetic improvement of processes yielding in an antibiotic-overproducing Streptomyces coelicolor ribosomal rpsL mutant.
microbial products. FEMS Microbiol. Rev. 30, 187–214. Mol. Microbiol. 61, 883–897.
Antón, N., Santos-Aberturas, J., Mendes, M.V., Guerra, S.M., Martı́n, J.F., Aparicio, J.F., Hosoya, Y., Okamoto, S., Muramatsu, H., Ochi, K., 1998. Acquisition of certain
2007. PimM, a PAS domain positive regulator of pimaricin biosynthesis in streptomycin-resistant (str) mutations enhances antibiotic production in
Streptomyces natalensis. Microbiology 153, 3174–3183. bacteria. Antimicrob. Agents. Chemother. 42, 2041–2047.
Beltrametti, F., Rossi, R., Selva, E., Marinelli, F., 2006. Antibiotic production Hu, H., Ochi, K., 2001. Novel approach for improving the productivity of antibiotic-
improvement in the rare actinomycete Planobispora rosea by selection of producing strains by inducing combined resistant mutations. Appl. Environ.
mutants resistant to the aminoglycosides streptomycin and gentamycin and to Microbiol. 67, 1885–1892.
rifamycin. J. Ind. Microbiol. Biotechnol. 33, 283–288. Hu, Z., Hopwood, D.A., Hutchinson, C.R., 2003. Enhanced heterologous polyketide
Bérdy, J., 2005. Bioactive microbial metabolites. J. Antibiot. 58, 1–26. production in Streptomyces by exploiting plasmid co-integration. J. Ind.
Blanco, G., Fernández, E., Fernández, M.J., Braña, A.F., Weissbach, U., Künzel, E., Microbiol. Biotechnol. 30, 516–522.
Rohr, J., Méndez, C., Salas, J.A., 2000. Characterization of two glycosyltrans- Huang, J., Shi, J., Molle, V., Sohlberg, B., Weaver, D., Bibb, M.J., Karoonuthaisiri, N.,
ferases involved in early glycosylation steps during biosynthesis of the Lih, C.J., Kao, C.M., Buttner, M.J., Cohen, S.N., 2005. Cross-regulation among
antitumor polyketide mithramycin by Streptomyces argillaceus. Mol. Gen. disparate antibiotic biosynthetic pathways of Streptomyces coelicolor. Mol.
Genet. 262, 991–1000. Microbiol. 58, 1276–1287.
Bro, C., Nielsen, J., 2004. Impact of ‘ome’ analyses on inverse metabolic Hung, T.V., Malla, S., Park, B.C., Liou, K., Lee, H.C., Sohng, J.K., 2007. Enhancement of
engineering. Metab. Eng. 6, 204–211. clavulanic acid by replicative and integrative expression of ccaR and cas2 in
Brünker, P., Minas, W., Kallio, P.T., Bailey, J.E., 1998. Genetic engineering of an Streptomyces clavuligerus NRRL3585. J. Microbiol. Biotechnol. 17, 1538–1545.
industrial strain of Saccharopolyspora erythraea for stable expression of the Ishizuka, H., Horinouchi, S., Kieser, H.M., Hopwood, D.A., Beppu, T., 1992. A putative
Vitreoscilla haemoglobin gene (vhb). Microbiology 144, 2441–2448. two-component regulatory system involved in secondary metabolism in
Butler, M.J., Bruheim, P., Jovetic, S., Marinelli, F., Postma, P.W., Bibb, M.J., 2002. Streptomyces spp. J. Bacteriol. 173, 2311–2318.
Engineering of primary carbon metabolism for improved antibiotic production Jung, W.S., Lee, S.K., Hong, J.S., Park, S.R., Jeong, S.J., Han, A.R., Sohng, J.K., Kim, B.G.,
in Streptomyces lividans. Appl. Environ. Microbiol. 68, 4731–4739. Choi, C.Y., Sherman, D.H., Yoon, Y.J., 2006. Heterologous expression of tylosin
Chakraburtty, R., Bibb, M., 1997. The ppGpp synthetase gene (relA) of Streptomyces polyketide synthase and production of a hybrid bioactive macrolide in
coelicolor A3(2) plays a conditional role in antibiotic production and Streptomyces venezuelae. Appl. Microbiol. Biotechnol. 72, 763–769.
morphological differentiation. J. Bacteriol. 179, 5854–5861. Jung, W.S., Jeong, S.J., Park, S.R., Choi, C.Y., Park, B.C., Park, J.W., Yoon, Y.J., 2008.
Chouayekh, H., Virolle, M.J., 2002. The polyphosphate kinase plays a negative role Enhanced heterologous production of desosaminyl macrolides and their
in the control of antibiotic production in Streptomyces lividans. Mol. Microbiol. hydroxylated derivatives by overexpression of the pikD regulatory gene in
43, 919–930. Streptomyces venezuelae. Appl. Environ. Microbiol. 74, 1972–1979.
Crameri, R., Davies, J.E., 1986. Increased production of aminoglycosides associated Kim, D.J., Huh, J.H., Yang, Y.Y., Kang, C.M., Lee, I.H., Hyun, C.G., Hong, S.K., Suh, J.W.,
with amplified antibiotic resistance genes. J. Antibiot. 39, 128–135. 2003. Accumulation of S-adenosyl-L-methionine enhances production of
Cropp, A., Chen, S., Liu, H., Zhang, W., Reynolds, K.A., 2001. Genetic approaches for actinorhodin but inhibits sporulation in Streptomyces lividans TK23. J. Bacteriol.
controlling ratios of related polyketide products in fermentation processes. 185, 592–600.
J. Ind. Microbiol. Biotechnol. 27, 368–377. Kuscer, E., Coates, N., Challis, I., Gregory, M., Wilkinson, B., Sheridan, R., Petković,
Cundliffe, E., 1989. How antibiotic-producing organisms avoid suicide. Annu. Rev. H., 2007. Roles of rapH and rapG in positive regulation of rapamycin
Microbiol. 43, 207–233. biosynthesis in Streptomyces hygroscopicus. J. Bacteriol. 189, 4756–4763.
Dairi, T., Aisaka, K., Katsumata, R., Hasegawa, M., 1995. A self-defense gene Li, R., Townsend, C.A., 2006. Rational strain improvement for enhanced clavulanic
homologous to tetracycline effluxing gene essential for antibiotic production in acid production by genetic engineering of the glycolytic pathway in
Streptomyces aureofaciens. Biosci. Biotechnol. Biochem. 59, 1835–1841. Streptomyces clavuligerus. Metab. Eng. 8, 240–252.
Decker, H., Summers, R.G., Hutchinson, C.R., 1994. Overproduction of the acyl Li, Y., Ling, H., Li, W., Tan, H., 2005. Improvement of nikkomycin production by
carrier protein component of a type II polyketide synthase stimulates enhanced copy of sanU and sanV in Streptomyces ansochromogenes and
production of tetracenomycin biosynthetic intermediates in Streptomyces characterization of a novel glutamate mutase encoded by sanU and sanV.
glaucescens. J. Antibiot. 47, 54–63. Metab. Eng. 7, 165–173.
Demain, A.L., 2006. From natural products discovery to commercialization: a Li, W., Ying, X., Guo, Y., Yu, Z., Zhou, X., Deng, Z., Kieser, H., Chater, K.F., Tao, M.,
success story. J. Ind. Microbiol. Biotechnol. 33, 486–495. 2006. Identification of a gene negatively affecting antibiotic production and
DeModena, J.A., Gutiérrez, S., Velasco, J., Fernández, F.J., Fachini, R.A., Galazzo, J.L., morphological differentiation in Streptomyces coelicolor A3(2). J. Bacteriol. 188,
Hughes, D.E., Martı́n, J.F., 1993. The production of cephalosporin C by 8368–8375.
ARTICLE IN PRESS

C. Olano et al. / Metabolic Engineering 10 (2008) 281–292 291

Liu, G., Tian, Y., Yang, H., Tan, H., 2005. A pathway-specific transcriptional Olano, C., Rodrı́guez, A.M., Méndez, C., Salas, J.A., 1995. A second ABC transporter is
regulatory gene for nikkomycin biosynthesis in Streptomyces ansochromogenes involved in oleandomycin resistance and its secretion by Streptomyces
that also influences colony development. Mol. Microbiol. 55, 1855–1866. antibioticus. Mol. Microbiol. 16, 333–343.
Lombó, F., Braña, A.F., Méndez, C., Salas, J.A., 1999. The mithramycin gene cluster of Omura, S., Ikeda, H., Ishikawa, J., Hanamoto, A., Takahashi, C., Shinose, M.,
Streptomyces argillaceus contains a positive regulatory gene and two repeated Takahashi, Y., Horikawa, H., Nakazawa, H., Osonoe, T., Kikuchi, H., Shiba, T.,
DNA sequences that are located at both ends of the cluster. J. Bacteriol. 181, Sakaki, Y., Hattori, M., 2001. Genome sequence of an industrial microorganism
642–647. Streptomyces avermitilis: deducing the ability of producing secondary metabo-
Lombó, F., Pfeifer, B., Leaf, T., Ou, S., Kim, Y.S., Cane, D.E., Licari, P., Khosla, C., 2001. lites. Proc. Natl. Acad. Sci. USA 98, 12215–12220.
Enhancing the atom economy of polyketide biosynthetic processes through Otten, S.L., Ferguson, J., Hutchinson, C.R., 1995. Regulation of daunorubicin
metabolic engineering. Biotechnol. Prog. 17, 612–617. production in Streptomyces peucetius by the dnrR2 locus. J. Bacteriol. 177,
Lomovskaya, N., Doi-Katayama, Y., Filippini, S., Nastro, C., Fonstein, L., Gallo, M., 1216–1224.
Colombo, A.L., Hutchinson, C.R., 1998. The Streptomyces peucetius dpsY and dnrX Otten, S.L., Olano, C., Hutchinson, C.R., 2000. The dnrO gene encodes a DNA-binding
genes govern early and late steps of daunorubicin and doxorubicin biosynth- protein that regulates daunorubicin production in Streptomyces peucetius by
esis. J. Bacteriol. 180, 2379–2386. controlling expression of the dnrN pseudo response regulator gene. Micro-
Lomovskaya, N., Otten, S.L., Doi-Katayama, Y., Fonstein, L., Liu, X.C., Takatsu, T., biology 146, 1457–1468.
Inventi-Solari, A., Filippini, S., Torti, F., Colombo, A.L., Hutchinson, C.R., 1999. Paradkar, A.S., Mosher, R.H., Anders, C., Griffin, A., Griffin, J., Hughes, C., Greaves, P.,
Doxorubicin overproduction in Streptomyces peucetius: cloning and character- Barton, B., Jensen, S.E., 2001. Applications of gene replacement technology to
ization of the dnrU ketoreductase and dnrV genes and the doxA cytochrome Streptomyces clavuligerus strain development for clavulanic acid production.
P-450 hydroxylase gene. J. Bacteriol. 181, 305–318. Appl. Environ. Microbiol. 67, 2292–2297.
Luzhetskyy, A., Mayer, A., Hoffmann, J., Pelzer, S., Holzenkämper, M., Schmitt, B., Parajuli, N., Viet, H.T., Ishida, K., Tong, H.T., Lee, H.C., Liou, K., Sohng, J.K., 2005.
Wohlert, S.E., Vente, A., Bechthold, A., 2007. Cloning and heterologous Identification and characterization of the afsR homologue regulatory gene from
expression of the aranciamycin biosynthetic gene cluster revealed a new Streptomyces peucetius ATCC 27952. Res. Microbiol. 156, 707–712.
flexible glycosyltransferase. Chembiochem 16, 599–602. Peirú, S., Menzella, H.G., Rodrı́guez, E., Carney, J., Gramajo, H., 2005. Production of
Madduri, K., Waldron, C., Merlo, D.J., 2001. Rhamnose biosynthesis pathway the potent antibacterial polyketide erythromycin C in Escherichia coli. Appl.
supplies precursors for primary and secondary metabolism in Saccharopoly- Environ. Microbiol. 71, 2539–2547.
spora spinosa. J. Bacteriol. 183, 5632–5638. Pérez-Llarena, F.J., Liras, P., Rodrı́guez-Garcı́a, A., Martı́n, J.F., 1997. A regulatory
Maharjan, S., Oh, T.J., Lee, H.C., Sohng, J.K., 2008. Heterologous expression of gene (ccaR) required for cephamycin and clavulanic acid production in
metK1-sp and afsR-sp in Streptomyces venezuelae for the production of Streptomyces clavuligerus: amplification results in overproduction of both
pikromycin. Biotechnol. Lett. [Epub ahead of print]. beta-lactam compounds. J. Bacteriol. 179, 2053–2059.
Malmberg, L.H., Hu, W.S., Sherman, D.H., 1993. Precursor flux control through Pfeifer, B.A., Admiraal, S.J., Gramajo, H., Cane, D.E., Khosla, C., 2001. Biosynthesis of
targeted chromosomal insertion of the lysine epsilon-aminotransferase (lat) complex polyketides in a metabolically engineered strain of E. coli. Science 291,
gene in cephamycin C biosynthesis. J. Bacteriol. 175, 6916–6924. 1790–1792.
Malmberg, L.H., Hu, W.S., Sherman, D.H., 1995. Effects of enhanced lysine epsilon- Ramos, A., Lombó, F., Braña, A.F., Rohr, J., Méndez, C., Salas, J.A., 2008. Biosynthesis
aminotransferase activity on cephamycin biosynthesis in Streptomyces clavu- of elloramycin in Streptomyces olivaceus requires glycosylation by enzymes
ligerus. Appl. Microbiol. Biotechnol. 44, 198–205. encoded outside the aglycon cluster. Microbiology 154, 781–788.
Malpartida, F., Niemi, J., Navarrete, R., Hopwood, D.A., 1990. Cloning and Rawlings, B.J., 2001. Type I polyketide biosynthesis in bacteria (Part A—erythro-
expression in a heterologous host of the complete set of genes for mycin biosynthesis). Nat. Prod. Rep. 18, 190–227.
biosynthesis of the Streptomyces coelicolor antibiotic undecylprodigiosin. Gene Reeves, A.R., Cernota, W.H., Brikun, I.A., Wesley, R.K., Weber, J.M., 2004.
93, 91–99. Engineering precursor flow for increased erythromycin production in Aero-
Martı́nez-Costa, O.H., Martı́n-Triana, A.J., Martı́nez, E., Fernández-Moreno, M.A., microbium erythreum. Metab. Eng. 6, 300–312.
Malpartida, F., 1999. An additional regulatory gene for actinorhodin production Reeves, A.R., Brikun, I.A., Cernota, W.H., Leach, B.I., González, M.C., Weber, J.M.,
in Streptomyces lividans involves a LysR-type transcriptional regulator. 2006. Effects of methylmalonyl-CoA mutase gene knockouts on erythromycin
J. Bacteriol. 181, 4353–4364. production in carbohydrate-based and oil-based fermentations of Saccharopo-
Mendes, M.V., Tunca, S., Antón, N., Recio, E., Sola-Landa, A., Aparicio, J.F., Martı́n, lyspora erythraea. J. Ind. Microbiol. Biotechnol. 33, 600–609.
J.F., 2007. The two-component phoR–phoP system of Streptomyces natalensis: Reeves, A.R., Brikun, I.A., Cernota, W.H., Leach, B.I., González, M.C., Weber, J.M.,
Inactivation or deletion of phoP reduces the negative phosphate regulation of 2007. Engineering of the methylmalonyl-CoA metabolite node of Saccharopo-
pimaricin biosynthesis. Metab. Eng. 9, 217–227. lyspora erythraea for increased erythromycin production. Metab. Eng. 9,
Méndez, C., Salas, J.A., 2001. The role of ABC transporters in antibiotic-producing 293–303.
organisms: drug secretion and resistance mechanisms. Res. Microbiol. 152, Rodrı́guez, E., Hu, Z., Ou, S., Volchegursky, Y., Hutchinson, C.R., McDaniel, R., 2003.
341–350. Rapid engineering of polyketide overproduction by gene transfer to indust-
Menéndez, N., Braña, A.F., Salas, J.A., Méndez, C., 2007. Involvement of a rially optimized strains. J. Ind. Microbiol. Biotechnol. 30, 480–488.
chromomycin ABC transporter system in secretion of a deacetylated precursor Rowe, C.J., Cortés, J., Gaisser, S., Staunton, J., Leadlay, P.F., 1998. Construction of new
during chromomycin biosynthesis. Microbiology 153, 3061–3070. vectors for high-level expression in actinomycetes. Gene 216, 215–223.
Minas, W., Brünker, P., Kallio, P.T., Bailey, J.E., 1998. Improved erythromycin Ryu, Y.G., Butler, M.J., Chater, K.F., Lee, K.J., 2006. Engineering of primary
production in a genetically engineered industrial strain of Saccharopolyspora carbohydrate metabolism for increased production of actinorhodin in
erythraea. Biotechnol. Prog. 14, 561–566. Streptomyces coelicolor. Appl. Environ. Microbiol. 72, 7132–7139.
Mosher, R.H., Paradkar, A.S., Anders, C., Barton, B., Jensen, S.E., 1999. Genes specific Salas, J.A., Méndez, C., 2005. Biosynthesis pathways for deoxysugars in antibiotic-
for the biosynthesis of clavam metabolites antipodal to clavulanic acid are producing actinomycetes: isolation, characterization and generation of novel
clustered with the gene for clavaminate synthase 1 in Streptomyces glycosylated derivatives. J. Mol. Microbiol. Biotechnol. 9, 77–85.
clavuligerus. Antimicrob. Agents. Chemother. 43, 1215–1224. Sánchez, C., Butovich, I.A., Braña, A.F., Rohr, J., Méndez, C., Salas, J.A., 2002. The
Motamedi, H., Hutchinson, C.R., 1987. Cloning and heterologous expression of a biosynthetic gene cluster for the antitumor rebeccamycin: characterization
gene cluster for the biosynthesis of tetracenomycin C, the anthracycline and generation of indolocarbazole derivatives. Chem. Biol. 9, 519–531.
antitumor antibiotic of Streptomyces glaucescens. Proc. Natl. Acad. Sci. USA 84, Scotti, C., Hutchinson, C.R., 1996. Enhanced antibiotic production by manipulation
4445–4449. of the Streptomyces peucetius dnrH and dnmT genes involved in doxorubicin
Murrell, J.M., Liu, W., Shen, B., 2004. Biochemical characterization of the SgcA1 (adriamycin) biosynthesis. J. Bacteriol. 178, 7316–7321.
alpha-D-glucopyranosyl-1-phosphate thymidylyltransferase from the enediyne Sekurova, O., Sletta, H., Ellingsen, T.E., Valla, S., Zotchev, S., 1999. Molecular cloning
antitumor antibiotic C-1027 biosynthetic pathway and overexpression of sgcA1 and analysis of a pleiotropic regulatory gene locus from the nystatin producer
in Streptomyces globisporus to improve C-1027 production. J. Nat. Prod. 67, Streptomyces noursei ATCC11455. FEMS Microbiol. Lett. 177, 297–304.
206–213. Sezonov, G., Blanc, V., Bamas-Jacques, N., Friedmann, A., Pernodet, J.L., Guérineau,
Mutka, S.C., Bondi, S.M., Carney, J.R., Da Silva, N.A., Kealey, J.T., 2006. Metabolic M., 1997. Complete conversion of antibiotic precursor to pristinamycin IIA by
pathway engineering for complex polyketide biosynthesis in Saccharomyces overexpression of Streptomyces pristinaespiralis biosynthetic genes. Nat.
cerevisiae. FEMS Yeast Res. 6, 40–47. Biotechnol. 15, 349–353.
Narva, K.E., Feitelson, J.S., 1990. Nucleotide sequence and transcriptional analysis of Shima, J., Hesketh, A., Okamoto, S., Kawamoto, S., Ochi, K., 1996. Induction of
the redD locus of Streptomyces coelicolor A3(2). J. Bacteriol. 172, 326–333. actinorhodin production by rpsL (encoding ribosomal protein S12) mutations
Nielsen, J., 1998. The role of metabolic engineering in the production of secondary that confer streptomycin resistance in Streptomyces lividans and Streptomyces
metabolites. Curr. Opin. Microbiol. 1, 330–336. coelicolor A3(2). J. Bacteriol. 178, 7276–7284.
Nishimura, K., Hosaka, T., Tokuyama, S., Okamoto, S., Ochi, K., 2007. Mutations in Sola-Landa, A., Moura, R.S., Martı́n, J.F., 2003. The two-component PhoR–PhoP
rsmG, encoding a 16S rRNA methyltransferase, result in low-level streptomycin system controls both primary metabolism and secondary metabolite biosynth-
resistance and antibiotic overproduction in Streptomyces coelicolor A3(2). J. esis in Streptomyces lividans. Proc. Natl. Acad. Sci. USA 100, 6133–61338.
Bacteriol. 189, 3876–3883. Song, J.Y., Kim, E.S., Kim, D.W., Jensen, S.E., Lee, K.J., 2008. Functional effects of
Okamoto, S., Lezhava, A., Hosaka, T., Okamoto-Hosoya, Y., Ochi, K., 2003. Enhanced increased copy number of the gene encoding proclavaminate amidino
expression of S-adenosylmethionine synthetase causes overproduction of hydrolase on clavulanic acid production in Streptomyces clavuligerus ATCC
actinorhodin in Streptomyces coelicolor A3(2). J. Bacteriol. 185, 601–609. 27064. J. Microbiol. Biotechnol. 18, 417–426.
Okamoto-Hosoya, Y., Okamoto, S., Ochi, K., 2003. Development of antibiotic- Stratigopoulos, G., Cundliffe, E., 2002. Expression analysis of the tylosin-
overproducing strains by site-directed mutagenesis of the rpsL gene in biosynthetic gene cluster: pivotal regulatory role of the tylQ product. Chem.
Streptomyces lividans. Appl. Environ. Microbiol. 69, 256–259. Biol. 9, 71–78.
ARTICLE IN PRESS

292 C. Olano et al. / Metabolic Engineering 10 (2008) 281–292

Stratigopoulos, G., Gandecha, A.R., Cundliffe, E., 2002. Regulation of tylosin Volchegursky, Y., Hu, Z., Katz, L., McDaniel, R., 2000. Biosynthesis of the anti-
production and morphological differentiation in Streptomyces fradiae parasitic agent megalomicin: transformation of erythromycin to megalomicin
by TylP, a deduced gamma-butyrolactone receptor. Mol. Microbiol. 45, in Saccharopolyspora erythraea. Mol. Microbiol. 37, 752–762.
735–744. Volokhan, O., Sletta, H., Sekurova, O.N., Ellingsen, T.E., Zotchev, S.B., 2005. An
Stratigopoulos, G., Bate, N., Cundliffe, E., 2004. Positive control of tylosin unexpected role for the putative 40 -phosphopantetheinyl transferase-encoding
biosynthesis: pivotal role of TylR. Mol. Microbiol. 54, 1326–1334. gene nysF in the regulation of nystatin biosynthesis in Streptomyces noursei
Stutzman-Engwall, K.J., Otten, S.L., Hutchinson, C.R., 1992. Regulation of secondary ATCC 11455. FEMS Microbiol. Lett. 249, 57–64.
metabolism in Streptomyces spp. and overproduction of daunorubicin in Wang, G., Tan, H., 2004. Enhanced production of nikkomycin X by over-expression
Streptomyces peucetius. J. Bacteriol. 174, 144–154. of SanO, a non-ribosomal peptide synthetase in Streptomyces ansochromogenes.
Stutzman-Engwall, K., Conlon, S., Fedechko, R., Kaczmarek, F., McArthur, H., Biotechnol. Lett. 26, 229–233.
Krebber, A., Chen, Y., Minshull, J., Raillard, S.A., Gustafsson, C., 2003. Wang, Y., Boghigian, B.A., Pfeifer, B.A., 2007. Improving heterologous polyketide
Engineering the aveC gene to enhance the ratio of doramectin to its CHC-B2 production in Escherichia coli by overexpression of an S-adenosylmethionine
analogue produced in Streptomyces avermitilis. Biotechnol. Bioeng. 82, synthetase gene. Appl. Microbiol. Biotechnol. 77, 367–373.
359–369. Wang, G., Hosaka, T., Ochi, K., 2008. Dramatic activation of antibiotic production in
Stutzman-Engwall, K., Conlon, S., Fedechko, R., McArthur, H., Pekrun, K., Chen, Y., Streptomyces coelicolor by cumulative drug-resistance mutations. Appl.
Jenne, S., La, C., Trinh, N., Kim, S., Zhang, Y.X., Fox, R., Gustafsson, C., Krebber, A., Environ. Microbiol. 74, 2834–2840.
2005. Semi-synthetic DNA shuffling of aveC leads to improved industrial Watanabe, K., Hotta, K., Praseuth, A.P., Koketsu, K., Migita, A., Boddy, C.N., Wang,
scale production of doramectin by Streptomyces avermitilis. Metab. Eng. 7, C.C., Oguri, H., Oikawa, H., 2006. Total biosynthesis of antitumor nonribosomal
27–37. peptides in Escherichia coli. Nat. Chem. Biol. 2, 423–428.
Sun, Y., Zhou, X., Liu, J., Bao, K., Zhang, G., Tu, G., Kieser, T., Deng, Z., 2002. Wei, X., Yunxiang, L., Yinghua, Z., 2006. Enhancement and selective production of
Streptomyces nanchangensis, a producer of the insecticidal polyether antibiotic oligomycin through inactivation of avermectin’s starter unit in Streptomyces
nanchangmycin and the antiparasitic macrolide meilingmycin, contains avermitilis. Biotechnol. Lett. 28, 911–916.
multiple polyketide gene clusters. Microbiology 148, 361–371. Wietzorrek, A., Bibb, M., 1997. A novel family of proteins that regulates antibiotic
Tamehiro, N., Hosaka, T., Xu, J., Hu, H., Otake, N., Ochi, K., 2003. Innovative production in streptomycetes appears to contain an OmpR-like DNA-binding
approach for improvement of an antibiotic-overproducing industrial strain of fold. Mol. Microbiol. 25, 1181–1184.
Streptomyces albus. Appl. Environ. Microbiol. 69, 6412–6417. Wohlert, S.E., Künzel, E., Machinek, R., Méndez, C., Salas, J.A., Rohr, J., 1999. The
Umeyama, T., Tanabe, Y., Aigle, B.D., Horinouchi, S., 1996. Expression of the structure of mithramycin reinvestigated. J. Nat. Prod. 62, 119–121.
Streptomyces coelicolor A3(2) ptpA gene encoding a phosphotyrosine protein Yanai, K., Murakami, T., Bibb, M., 2006. Amplification of the entire kanamycin
phosphatase leads to overproduction of secondary metabolites in S. lividans. biosynthetic gene cluster during empirical strain improvement of Streptomyces
FEMS Microbiol. Lett. 144, 177–184. kanamyceticus. Proc. Natl. Acad. Sci. USA 103, 9661–9666.
Vögtli, M., Chang, P.C., Cohen, S.N., 1994. afsR2: a previously undetected gene Zhang, Y.X., Perry, K., Vinci, V.A., Powell, K., Stemmer, W.P., del Cardayré, S.B., 2002.
encoding a 63-amino-acid protein that stimulates antibiotic production in Genome shuffling leads to rapid phenotypic improvement in bacteria. Nature
Streptomyces lividans. Mol. Microbiol. 14, 643–653. 415, 644–646.

You might also like