altering the glycosylation pattern of bioactive compounds

8
Carmen Méndez José A. Salas* Dept Biologia Funcional e Instituto Universitario de Oncología del Principado de Asturias (IUOPA), Universidad de Oviedo, 33006 Oviedo, Spain. *e-mail: jasf@ sauron.quimica.uniovi.es Nature produces an amazing number of products that have useful biological activities, including compounds with clinical application, such as antibiotics, antitumor or immunosuppressive agents, and also compounds with veterinary or agricultural applications, such as growth promoters, insecticides, herbicides and antiparasitic agents. These bioactive compounds are the end products of complex multistep biosynthetic pathways. Among microorganisms, approximately three-quarters of all known bioactive compounds are produced by members of the actinomycete family, mainly belonging to the genus Streptomyces 1 . Polyketides are probably the most interesting subgroup of bioactive natural products, because they have a great diversity of chemical structures and biological activities 2 . Commercially important polyketides include antibiotics (e.g. erythromycin, rifamycin, tetracyclins and oleandomycin), antitumor drugs (e.g. doxorubicin, aclacinomycin), cholesterol-lowering agents (e.g. lovastatin), immunosuppressive agents (e.g. rapamycin), insecticides (e.g. spinosyn) and antiparasitic agents (e.g. avermectins). During the past few years, there has been a great increase in our knowledge of the biosynthetic pathways by which microorganisms produce antibiotics and other bioactive secondary metabolites. This increase has prompted researchers in this field to manipulate antibiotic gene clusters to alter the structure of natural compounds. This new technology, named combinatorial biosynthesis 3 , manipulates genes in natural-product biosynthesis pathways as a way of producing natural-product analogs and thus forming novel derivatives with potential bioactivity 3–8 . An important source of structural biodiversity in natural products are the sugars that are attached to specific positions on the aglycon core. These sugar components usually participate in the molecular recognition of its cellular target 9 . Their presence is Many bioactive natural products are glycosylated compounds in which the sugars are important or essential for biological activity.The isolation of several sugar biosynthesis gene clusters and glycosyltransferases from different antibiotic-producing organisms,and the increasing knowledge about these biosynthetic pathways opens up the possibility of generating novel bioactive compounds through combinatorial biosynthesis in the near future. Recent advances in this area indicate that antibiotic glycosyltransferases show some substrate flexibility that might allow us to alter the types of sugar transferred to the different aglycons or, less frequently,to change the position of its attachment. Altering the glycosylation pattern of bioactive compounds Carmen Méndez and José A. Salas TRENDS in Biotechnology Vol.19 No.11 November 2001 http://tibtech.trends.com 0167-7799/01/$ – see front matter © 2001 Elsevier Science Ltd. All rights reserved. PII: S0167-7799(01)01765-6 449 Review 55 Cornils, B. and Wiebus, E. (1995) Aqueous catalysts for organic reactions. CHEMTECH 25, 33–38 56 Wan, K.T. and Davies, M.E. (1994) Asymmetric synthesis of naproxen by supported aqueous- phase catalysis. J. Catal. 148, 1–8 57 Ley, S.V. et al. (2000) Multi-step organic synthesis using solid-supported reagents and scavengers: a new paradigm in chemical library generation. J. Chem. Soc., Perkin Trans. 3815–4195 58 Kirschning, A. et al. (2001) Functionalized polymers – emerging versatile tools for solution-phase chemistry and automated parallel synthesis. Angew. Chem., Int. Ed. Engl. 40, 650–679 59 Wasserscheid, P. and Keim, W. (2000) Ionic liquids. Angew. Chem., Int. Ed. Engl. 39, 3772–3789 60 Bergbreiter, D.E. et al. (2000) Palladium- catalyzed C–C coupling under thermomorphic conditions. J. Am. Chem. Soc. 122, 9058–9064 61 Tas, D. et al. (1997) Bifunctional catalytic membrane containing Brønsted acid and sites for enantioselective hydrogenation. Chem. Commun. 2323–2324 62 Guijarro, D. et al. (1998) Enantioselective addition of diethylzinc reagents to N-(diphenylphosphinoyl) imines promoted by 2-azanorbornylmethanols. J. Org. Chem. 63, 2530–2535 63 Shvo, Y. et al. (1994) Sol-gel glass with enantioselective catalytic activity. J. Chem. Soc. Chem. Commun. 2179–2720 64 Belokon, B. et al. (1997) Asymmetric trimethylsilylcyanation of benzaldehyde catalyzed by (salen)Ti(IV) complexes derived from (R)- and/or (S)-4-hydroxy-5- formyl[2.2]paracyclophane and diamines. Tetrahedron Asymmetry 8, 3245–3250 65 Sellner, H. et al. (2000) Immobilization of BINOL by cross-linking copolymerization of styryl derivatives with styrene, and applications in enantioselective Ti and Al Lewis acid mediated additions of Et 2 Zn and Me 3 SiCN to aldehydes and of diphenyl nitrone to enol ethers. Chem. Eur. J. 6, 3692–3705 66 Vankelecom, I.F.J. et al. (1996) Chiral catalytic membranes. Angew. Chem., Int. Ed. Engl. 35, 1346–1348 67 Reger, S.R. and Janda, K.D. (2000) Polymer- supported (salen)Mn complexes for asymmetric epoxidation: a comparison between soluble and insoluble matrices. J. Am. Chem. Soc. 122, 6929–6934 68 Flach, H.N. et al. (1994) Polymeric surfactant systems in the asymmetric hydrogenation of amino acid precursors with a rhodium complex. Makromol. Chem. Phys. 195, 3289–3301 69 Baiker, A. and Blaser, H.U. (1997) Enantioselective catalysts and reactions. In Handbook of Heterogeneous Catalysis (Ertl, G. et al., eds), pp. 2422–2436, VCH 70 Blaser, H.U. and Spindler, F. (1999) The chiral switch of metolachlor. In Comprehensive Asymmetric Catalysis (Jacobsen, E.N. et al., eds), pp. 1427–1437, Springer 71 Schmid, R. and Scalone, M. (1999) Process R&D of pharmaceuticals, vitamins and fine chemicals. In Comprehensive Asymmetric Catalysis (Jacobsen, E.N. et al., eds), pp. 1439–1451, Springer 72 Akutagawa, S. (1999) Asymmetric isomerization of olefins. In Comprehensive Asymmetric Catalysis (Jacobsen, E.N. et al., eds), pp. 1461–1473, Springer 73 Liese, A. et al. (2000) Industrial biotransformations, Wiley-VCH Review

Upload: carmen-mendez

Post on 16-Sep-2016

218 views

Category:

Documents


3 download

TRANSCRIPT

Page 1: Altering the glycosylation pattern of bioactive compounds

Carmen Méndez

José A. Salas*

Dept Biologia Funcional eInstituto Universitario deOncología del Principadode Asturias (IUOPA),Universidad de Oviedo,33006 Oviedo, Spain.*e-mail: [email protected]

Nature produces an amazing number of products

that have useful biological activities, including

compounds with clinical application, such as

antibiotics, antitumor or immunosuppressive

agents, and also compounds with veterinary or

agricultural applications, such as growth promoters,

insecticides, herbicides and antiparasitic agents.

These bioactive compounds are the end products of

complex multistep biosynthetic pathways. Among

microorganisms, approximately three-quarters of

all known bioactive compounds are produced by

members of the actinomycete family, mainly

belonging to the genus Streptomyces1.

Polyketides are probably the most interesting

subgroup of bioactive natural products, because they

have a great diversity of chemical structures and

biological activities2. Commercially important

polyketides include antibiotics (e.g. erythromycin,

rifamycin, tetracyclins and oleandomycin),

antitumor drugs (e.g. doxorubicin, aclacinomycin),

cholesterol-lowering agents (e.g. lovastatin),

immunosuppressive agents (e.g. rapamycin),

insecticides (e.g. spinosyn) and antiparasitic agents

(e.g. avermectins). During the past few years, there

has been a great increase in our knowledge of the

biosynthetic pathways by which microorganisms

produce antibiotics and other bioactive secondary

metabolites. This increase has prompted

researchers in this field to manipulate antibiotic

gene clusters to alter the structure of natural

compounds. This new technology, named

combinatorial biosynthesis3, manipulates genes in

natural-product biosynthesis pathways as a way of

producing natural-product analogs and thus

forming novel derivatives with potential

bioactivity3–8.

An important source of structural biodiversity in

natural products are the sugars that are attached to

specific positions on the aglycon core. These sugar

components usually participate in the molecular

recognition of its cellular target9. Their presence is

Many bioactive natural products are glycosylated compounds in which the

sugars are important or essential for biological activity. The isolation of

several sugar biosynthesis gene clusters and glycosyltransferases from

different antibiotic-producing organisms, and the increasing knowledge

about these biosynthetic pathways opens up the possibility of generating

novel bioactive compounds through combinatorial biosynthesis in the near

future. Recent advances in this area indicate that antibiotic

glycosyltransferases show some substrate flexibility that might allow us to

alter the types of sugar transferred to the different aglycons or, less

frequently, to change the position of its attachment.

Altering the glycosylation pattern of

bioactive compounds

Carmen Méndez and José A. Salas

TRENDS in Biotechnology Vol.19 No.11 November 2001

http://tibtech.trends.com 0167-7799/01/$ – see front matter © 2001 Elsevier Science Ltd. All rights reserved. PII: S0167-7799(01)01765-6

449Review

55 Cornils, B. and Wiebus, E. (1995) Aqueous

catalysts for organic reactions. CHEMTECH

25, 33–38

56 Wan, K.T. and Davies, M.E. (1994) Asymmetric

synthesis of naproxen by supported aqueous-

phase catalysis. J. Catal. 148, 1–8

57 Ley, S.V. et al. (2000) Multi-step organic

synthesis using solid-supported reagents and

scavengers: a new paradigm in chemical

library generation. J. Chem. Soc., Perkin Trans.

3815–4195

58 Kirschning, A. et al. (2001) Functionalized

polymers – emerging versatile tools for

solution-phase chemistry and automated

parallel synthesis. Angew. Chem., Int. Ed. Engl.

40, 650–679

59 Wasserscheid, P. and Keim, W. (2000) Ionic liquids.

Angew. Chem., Int. Ed. Engl. 39, 3772–3789

60 Bergbreiter, D.E. et al. (2000) Palladium-

catalyzed C–C coupling under thermomorphic

conditions. J. Am. Chem. Soc. 122, 9058–9064

61 Tas, D. et al. (1997) Bifunctional catalytic

membrane containing Brønsted acid and sites for

enantioselective hydrogenation. Chem. Commun.

2323–2324

62 Guijarro, D. et al. (1998) Enantioselective

addition of diethylzinc reagents to

N-(diphenylphosphinoyl) imines promoted by

2-azanorbornylmethanols. J. Org. Chem. 63,

2530–2535

63 Shvo, Y. et al. (1994) Sol-gel glass with

enantioselective catalytic activity. J. Chem. Soc.

Chem. Commun. 2179–2720

64 Belokon, B. et al. (1997) Asymmetric

trimethylsilylcyanation of benzaldehyde

catalyzed by (salen)Ti(IV) complexes derived

from (R)- and/or (S)-4-hydroxy-5-

formyl[2.2]paracyclophane and diamines.

Tetrahedron Asymmetry 8, 3245–3250

65 Sellner, H. et al. (2000) Immobilization of BINOLby

cross-linking copolymerization of styryl derivatives

with styrene, and applications in enantioselective

Ti and Al Lewis acid mediated additions of Et2Zn

and Me3SiCN to aldehydes and of diphenyl nitrone

to enol ethers. Chem. Eur. J. 6, 3692–3705

66 Vankelecom, I.F.J. et al. (1996) Chiral catalytic

membranes. Angew. Chem., Int. Ed. Engl. 35,

1346–1348

67 Reger, S.R. and Janda, K.D. (2000) Polymer-

supported (salen)Mn complexes for asymmetric

epoxidation: a comparison between soluble and

insoluble matrices. J. Am. Chem. Soc. 122,

6929–6934

68 Flach, H.N. et al. (1994) Polymeric surfactant

systems in the asymmetric hydrogenation of

amino acid precursors with a rhodium complex.

Makromol. Chem. Phys. 195, 3289–3301

69 Baiker, A. and Blaser, H.U. (1997)

Enantioselective catalysts and reactions. In

Handbook of Heterogeneous Catalysis (Ertl, G.

et al., eds), pp. 2422–2436, VCH

70 Blaser, H.U. and Spindler, F. (1999) The chiral

switch of metolachlor. In Comprehensive

Asymmetric Catalysis (Jacobsen, E.N. et al., eds),

pp. 1427–1437, Springer

71 Schmid, R. and Scalone, M. (1999) Process R&D

of pharmaceuticals, vitamins and fine chemicals.

In Comprehensive Asymmetric Catalysis

(Jacobsen, E.N. et al., eds), pp. 1439–1451, Springer

72 Akutagawa, S. (1999) Asymmetric isomerization

of olefins. In Comprehensive Asymmetric

Catalysis (Jacobsen, E.N. et al., eds),

pp. 1461–1473, Springer

73 Liese, A. et al. (2000) Industrial

biotransformations, Wiley-VCH

Review

Page 2: Altering the glycosylation pattern of bioactive compounds

therefore important for the biological activity of the

compounds, even, in many cases, essential9. These

sugars can be linked to the aglycon as

monosaccharides, disaccharides or oligosaccharides

of variable sugar length through C-, N- or (most

often) O-glycosylation (Fig. 1). These sugars are

transferred to the corresponding aglycon by

glycosyltransferases, which are generally sugar,

aglycon and site specific. The importance of these

sugar moieties in biological activity has led to the

idea of using combinatorial biosynthesis to

incorporate different sugars into an aglycon or to

glycosylate a different position of the aglycon. This

requires an understanding of how these sugars are

synthesized and knowledge about the

glycosyltransferases responsible for sugar transfer

and their specificity. Great progress has been made

in the identification and characterization of genes

involved in the biosynthesis of these sugar moieties

and in the isolation of genes coding for

glycosyltransferases from many different

antibiotic-producing microorganisms.

Sugar biosynthesis genes

The most abundant and diverse group of secondary

carbohydrate moieties in bioactive compounds are

the 6-deoxyhexoses (6DOHs), which occur in a wide

range of plants, fungi and bacteria; 70 different

6DOHs form part of bioactive compounds10–12

(Table 1). The 6DOHs are formed from nucleoside

diphosphate-activated hexoses (mainly D-glucose)

via a 4-keto-6-deoxy intermediate. The two common

enzymatic steps leading to the biosynthesis of this

intermediate are catalyzed by a dNDP–D-hexose

synthase and a dNDP–D-hexose-4,6-dehydratase,

and structural diversification into either D or L forms

is the result of the action of a 5- or a 3,5-epimerase.

The different 6DOHs will also vary depending on the

substituents and/or the stereochemistry of carbon

atoms at positions 2–5 of the hexose carbon

chain13–15. These typically include further

deoxygenations, transaminations and C-, N- or

O-methylations, thus generating the large family of

6DOHs (Fig. 2).

More and more genes involved in the

biosynthesis of various 6DOHs from antibiotic-

producing actinomycetes are being reported

(Table 1). The involvement of these genes in sugar

biosynthesis has usually been proved by the

insertional inactivation of individual genes and the

isolation of biosynthetic intermediates that either

lack the sugar or possess a modified sugar linked to

the aglycon. For most of them, the assignment of a

role for their gene products in 6DOH biosynthesis

has been made by comparison with similar proteins

in databases. So far, evidence has been provided in

only a few cases to show which set of genes are

necessary to direct the biosynthesis of the sugar.

This is the case for L-daunosamine in the

daunorubicin producer Streptomyces peucetius16,

TRENDS in Biotechnology Vol.19 No.11 November 2001

http://tibtech.trends.com

450 Review

TRENDS in Biotechnology

OH O

OHO

O

O

OH

OCH3

CH3

OCH3

OHNH2

OH OOH

H3CO

O

CH3

OHO

OH

H

CH3

O OOOO

OHOHCH3

CH3CH3CH3

OH

HO

OOO

CH3OH

O

CH3OH

OH

O

O

O

CH3

CH3

O

CH3

OH

CH3

O

OHCH3

OHCH3

CH3

O CH3

N(CH3)2OH

OCH3

OHCH3

OMe

Doxorubicin

Mithramycin

Erythromycin A

Urdamycin AO

OHOH

CH3

O

O

O

CH3

OOHCH3

OOH

O CH3

OH

OH

OO

OCH3

OH

O OH

OH

OMe

OH

Rebeccamycin

NH

N NH

O O

Cl Cl

Fig. 1.The chemicalstructures of someglycosylated bioactivecompounds produced byactinomycetes. Examples areshown that differ in thenumber of sugars(monosaccharides such aserythromycin A, doxorubicin,urdamycin A andrebeccamycin, disaccharidessuch as mithramycin, andtrisaccharides such asmithramycin and urdamycinA [mithramycin contains botha di- and a trisaccharideattached to the aglycon andurdamycin A contains amono- and a trisaccharideattached to the aglycon]) andthe linkage to the aglycon(O-glycosylation as inerythromycin A, mithramycin,doxorubicin and urdamycin A,C-glycosylation as inurdamycin A, andN-glycosylation as inrebeccamycin).

Table 1. 6DOH gene clusters from bioactive compounds that have been totally

or partially identified and characterized

6DOH Bioactive compound Producer organisma Refs

D-Desosamine Erythromycin Sacc. erythraea 29,40,50

Oleandomycin S. antibioticus 17,26,39

Pikromycin S. venezuelae 51

Megalomicin M. megalomicea 52

D-Olivose Mithramycin S. argillaceus 20,31

Urdamycin S. fradiae 32

Landomycin S. cyanogenus 53

D-Oliose Mithramycin S. argillaceus 20,31

D-Mycarose Mithramycin S. argillaceus 20,31

D-Mycaminose Tylosin S. fradiae 19,41

D-Mycinose Tylosin S. fradiae 54

D-Mycosamine Nystatin S. noursei 55

L-Dihydrostreptose Streptomycin S. griseus 14

L-Oleandrose Oleandomycin S. antibioticus 17,39

Avermectin S. avermitilis 56

L-Mycarose Erythromycin Sacc. erythraea 29,49,50

Megalomicin M. megalomicea 52

Tylosin S. fradiae 19,41

L-Noviose Novobiocin S. spheroides 57

L-Rhodinose Urdamycin S. fradiae 32

Landomycin S. cyanogenus 53

Granaticin S. violaceoruber 58

L-Daunosamine Daunorubicin S. peucetius 59,60

L-Nogalose Nogalamycin S. nogalater 61

L-Megosamineb Megalomicin M. megalomicea 52

L-Rhodosamine Aclacinomycins S. galilaeus 62

L-Epivancosamine Chloroeremomycin A. orientalis 63

2-Deoxy-L-fucose Aclacinomycins S. galilaeus 62aAbbreviations: A., Amycolatopsis; M., Micromonospora; Sacc., Saccharopolyspora; S., Streptomyces.bAlso named L-rhodosamine.

Page 3: Altering the glycosylation pattern of bioactive compounds

L-olivose and L-oleandrose in the oleandomycin

producer Streptomyces antibioticus17, and

L-oleandrose in the avermectin producer

Streptomyces avermitilis18. Plasmid constructs were

generated containing all the genes involved in the

biosynthesis of these 6DOHs and shown to

incorporate these sugars into the corresponding

aglycon in the presence of the respective

glycosyltransferases. Some sugar gene clusters have

been cloned and characterized from more than one

producer organism, including the amino sugar

D-desosamine, whose biosynthesis gene cluster has

been cloned from erythromycin-, oleandomycin-,

pikromycin- and megalomicin-producing organisms.

In a few (but increasing) cases, the activity of the

gene products on different sugar intermediates has

been shown using in vitro enzymatic assays19–23.

Recently, all enzymes involved in the biosynthesis

of L-epivancosamine from the chloroeremomycin

pathway have been expressed in Escherichia coli,

purified and characterized for their enzymatic

activities23. This is the first report of the

reconstitution of TDP–deoxysugar biosynthesis

from the individual enzymes from an antibiotic

biosynthetic pathway.

Glycosyltransferase genes

With the increasing knowledge of the organization of

antibiotic-biosynthesis gene clusters, many genes

encoding glycosyltransferases are being reported.

More than 50 glycosyltransferase genes have been

reported from antibiotic-producing actinomycetes.

All of these belong to glycosyltransferase family 1 in

the Henrissat classification24 (Fig. 3). Most

glycosyltransferases share a conserved motif region

(glycine-rich) close to the C terminus that is

characteristic of UDP–glycosyl and

UDP–glucuronosyl transferases. Four different

desosaminyl glycosyltransferases from various

antibiotic-producing organisms have been identified.

These have amino acid identities ranging from 50.1%

to 53.8% (with the exception of EryCIII and MegCIII,

which share 83.4% identical amino acids but

recognize the same aglycon). It is worth mentioning

that two urdamycin-A glycosyltransferases

(UrdGT1b and UrdGT1c) show a very high degree of

similarity: they share 91% identical amino acids and

only differ in a few amino acids that are mostly

located within a 31-amino acid region25.

Interestingly, these two glycosyltransferases

transfer different 6DOHs: UrdGT1c incorporates

L-rhodinose and UrdTG1b transfers D-olivose.

Three other glycosyltransferases have been

described in antibiotic producers: OleD and OleI

from the oleandomycin producer S. antibioticus26

and MgtA from the spiramycin producer

Streptomyces ambofaciens27. These transferases do

not incorporate any sugar present in the final

antibiotic molecule but rather glycosylate the

antibiotic (or a biosynthetic intermediate), leading

to its inactivation, and therefore behave as self-

resistance mechanisms.

Recently, the glucosyltransferase GtfB has been

crystallized and its structure has been determined by

X-ray analysis at 1.8 Å resolution28. This

glucosyltransferase transfers the glucose residue

from UDP–glucose to the vancomycin aglycon. The

enzyme has been shown to have a two-domain

structure, with a deep interdomain cleft that

probably contains the UDP–glucose-binding site.

Determining the structure of this transferase will

allow us to get further insight into the specificity of

the glycosyltransferases and will facilitate

engineering of antibiotics through the incorporation

of novel sugars.

Using producer organisms as cell factories

Microorganisms producing glycosylated compounds

can be used as cell factories to generate novel

glycosylated compounds. This strategy takes

advantage of the existence of an intracellular pool of

nucleotide-activated sugars and intermediates in

these microorganisms. Evidence of the ability of

some antibiotic biosynthesis glycosyltransferases to

incorporate different sugars (‘sugar flexibility’)

came from insertional inactivation experiments of

genes that encode enzymes involved in sugar

biosynthesis. To assign functions to putative sugar

biosynthesis genes, knockout mutants are

frequently generated either by inserting an

antibiotic resistance marker within the coding

region of a selected gene or by ‘in-frame’ deletion of

part of the gene. This approach usually leads to the

TRENDS in Biotechnology Vol.19 No.11 November 2001

http://tibtech.trends.com

451Review

TRENDS in Biotechnology

O

OH

OH

CH3

O

OH

OH

CH3

NH2

L-Rhodinose L-Daunosamine

O

OH

OH

CH3

N(CH3)2

L-Rhodosamine

O

OH

OHCH3OH

CH3

L-Mycarose

O

OH

OH

CH3

NH2 OH

O

OH

NH2

CH3

OH OH

O

OH

OH

CH3

OMe OMe

D-MycosamineD-Perosamine

D-Mycinose

O

OH

OH

CH3

N(CH3)2

OH

D-Mycaminose

O

OH

CH3

N(CH3)2

OH

D-Desosamine

O

OH

OH

CH3

OH OH

CH3

D-Mycarose

O

OH

OH

CH3

OH

D-Olivose

O

OH

OHCH3

OH

D-Oliose

O

OH

N(CH3)2

CH3

D-Forosamine

O

OH

CH3

OH

OMe

D-Chalcose

D-Deoxyhexoses

L-Deoxyhexoses

O

OH

OHCH3

OMe

L-Oleandrose

O

OH

OHCH3

CH3

OMe

L-Cladinose

O

OH

OHCH3

OH OH

CH3

L-Nogalose

O

OH

OH

CH3

OH

L-Oliose

O

OH

OHCH3

OH OH

CH3

L-Noviose

O

OH

OHCH3CH3

NH2

L-Epivancosamine

Fig. 2. The structures ofsome D- and L-6-deoxyhexoses that areused to glycosylatebioactive compounds.

Page 4: Altering the glycosylation pattern of bioactive compounds

generation of nonproducing mutants, many of which

accumulate biosynthesis intermediates that lack

the sugar owing to the inability of the

glycosyltransferase to recognize and transfer the

synthesized sugar biosynthetic intermediate.

However, in some cases, the glycosyltransferases

are able to transfer (although usually less

efficiently) some sugar biosynthesis intermediates

to the aglycon. This type of mutant has been

generated in erythromycin29,30, mithramycin31,

urdamycin32 and methymycin and/or

pikromycin33–35 producers, and the new compounds

accumulated by these mutants have been purified

and their structures elucidated (Fig. 4).

Mutants affected in specific sugar biosynthesis

genes can also be used as hosts for the expression of

genes involved in sugar biosynthesis in other

pathways. These recombinant strains can synthesize

novel sugar derivatives through the joint action of the

host genes and the newly acquired genes. An

outstanding and ingenious strategy was designed for

the production of the antitumor drugs epirubicin

(4′-epidoxorubicin) and 4′-epidaunorubicin36. These

are important cancer chemotherapy drugs that are

usually produced by low-yielding semisynthetic

processes in which some chemical modification steps

follow the fermentative production of doxorubicin and

daunorubicin by S. peucetius. Combining gene

inactivation and gene expression led to a less time-

consuming and less expensive procedure to produce

these two compounds. Doxorubicin and daunorubicin

differ from epirubicin and 4′-epidaunorubicin only in

the configuration at C4 of the sugar moiety,

L-daunosamine. A mutant in dnmV (which encodes a

4-ketoreductase) produces a 4-keto intermediate. One

of two other genes from different sugar pathways

(eryBIV or avrE from the L-mycarose and L-oleandrose

sugar pathways, respectively), both encoding

4-ketoreductases with different diastereoselectivity

at C4 from that of the daunosamine ketoreductase,

were then expressed independently in this mutant. In

this way, the 4′-epimeric anthracycline derivatives

were formed and recovered from the fermentation

broths with satisfactory yield. One of the implications

of this experiment was the demonstration that the

daunosaminyl glycosyltransferase could recognize

both C4 epimers of this sugar (L-daunosamine and

L-epidaunosamine).

Transfer of glycosyltransferases between species

of antibiotic producers can be another strategy to

promote the incorporation of different sugars into

an aglycon. This is usually achieved by transferring

glycosyltransferases between organisms producing

structurally related bioactive compounds. The host

strain will act as a cell factory providing the source

of nucleotide-activated sugars for the glycosylation

events. The aglycon can be synthesized

endogenously by the host strain, either governing its

biosynthesis by chromosomal genes from the host or

by exogenous genes acquired by the incorporation of

an appropriate replicative or integrative vector

harboring the coding genes. Alternatively, the

aglycon can be fed to the engineered strain. The first

example of this type of strategy was the production

of a hybrid glycopeptide antibiotic by expression of

the glycosyltransferase gene gtfE′ from the

vancomycin producer Amycolatopsis orientalis

C329.4 into the producer of the nonglycosylated

glycopeptide A47934 Streptomyces toyocaensis37.

This resulted in the formation of a new glucosyl

A47934 derivative.

TRENDS in Biotechnology Vol.19 No.11 November 2001

http://tibtech.trends.com

452 Review

SpcFSpcGNysDIPimKBgtfAGtfABgtfBGtfBGtfEBgtfCGtfCGtfDTylNElmGTSnogDLanGT2UrdGT2Gra-orf14CumHNovNMtmGIILanGT4UrdGT1aDauHDnrHAknKEryBVMegBVAknSSnogEDnrSDesVIITylMIIEryCIIIMegCIIIMegDIOleG1OleG2MtmGIIISnogZUrdGT1bUrdGT1cLanGT1LanGT3MtmGIMtmGIVGimAOleDOleIAveBITylCVNgtAMitBArd-orf5

TRENDS in Biotechnology

Fig. 3. A dendrogram showing the relationships among differentglycosyltransferases from antibiotic-producing actinomycetes. Theseglycosyltransferases are involved in the biosynthesis of the differentbioactive compounds mentioned (glycosyltransferases, products):SpcF and SpcG, spectinomycin (accession number U70376); NysDI,nystatin55; PimK, pimaricin64; BgtfA, BgtfB and BgtfC, balhimycin65;GtfA, GtfB and GtfC, chloroeremomycin63; GtfD and GtfE, vancomycin37;TylN, TylM2 and TylCV, tylosin41,54,66; ElmGT, elloramycin42; SnogD,SnogE and SnogZ, nogalamycin (accession number AF187532);LanGT1, LanGT2, LanGT3 and LanGT4, landomycin53; UrdGT1a,UrdGT1b, UrdGT1c and UrdGT2, urdamycin25,67; Gra-orf14,granaticin58; CumH, coumermycin A1 (Ref. 68); NovN, novobiocin57;MtmGI, MtmGII, MtmGIII and MtmGIV, mithramycin46,69; DauH,daunomycin70; DnrS, daunorubicin/doxorubicin71; DnrH, baumycin59;AknK and AknS, aclacinomycin62; EryBV and EryCIII, erythromycin50;DesVII, pikromycin51; MegDI, MegBV and MegCIII, megalomycin52;OleG1 and OleG2, oleandomycin39; GimA, spiramycin72; OleD and OleI,oleandomycin26; AveBI, avermectin56; Ngt, rebeccamycin73; Ard-orf5,antibiotic A201A (Ref. 74); MitB, mitomycin75.

Page 5: Altering the glycosylation pattern of bioactive compounds

Other interesting examples of the formation of

novel compounds by heterologous expression of

glycosyltransferases are the production of novel

macrolide derivatives. In this case, the

glycosyltransferases transferred different sugars

from those they usually did. One example is the

formation of a new erythromycin derivative by

heterologous expression of a glycosyltransferase

from the oleandomycin producer38 (Fig. 5a). The

oleGII gene from S. antibioticus (oleandomycin

producer) codes for an oleandrosyl

glycosyltransferase responsible for the transfer of

this sugar into the 3-hydroxy group of the

8,8a-deoxyoleandolide aglycon17,39. In the

erythromycin producer Saccharopolyspora

erythraea, the EryBV transferase mediates the

transfer of L-mycarose into the same position of the

6-deoxyerythronolide aglycon. Expression of oleGII

in an eryBV deletion mutant produced, with a good

yield, several new active compounds in which the

neutral sugar L-rhamnose was attached to position

C3 of the aglycon. One of these new compounds was

identified as 3-rhamnosyl-6-deoxyerythromycin B,

which showed antibiotic activity38 (Fig. 5a).

The second example involves the formation of a

tylosin derivative in an erythromycin mutant40

(Fig. 5b). The tylM2 glycosyltransferase gene from

Streptomyces fradiae is responsible for the

incorporation of mycaminose during tylosin

biosynthesis41. The ability of this transferase to use

a different amino sugar was examined by

integrating the tylM2 glycosyltransferase gene into

the chromosome of a S. erythraea triple mutant

SGT2, in which the genes coding for the polyketide

synthase and the mycarosyl and desosaminyl

transferases were deleted40. This mutant still

retains the ability to synthesize L-mycarose and

D-desosamine. By feeding this recombinant strain

with the 16-membered aglycone tylactone (the

TylM2 transferase substrate), a new hybrid

glycoside was produced and identified as

5-O-desosaminyl-tylactone (Fig. 5b), indicating

that TylM2 can recognize and transfer a different

amino sugar40. These two examples illustrate the

existence of some sugar flexibility in antibiotic

glycosyltransferases.

The ElmGT glycosyltransferase from S. olivaceus

is an interesting example of a glycosyltransferase

that can transfer different deoxysugars. This

transferase incorporates an L-rhamnose moiety

(which is later permethylated) to the elloramycinone

aglycon during the biosynthesis of the anthracycline-

like antitumor drug elloramycin42. ElmGT has a

broader substrate specificity range than any other

glycosyltransferase from antibiotic biosynthesis

pathways; it can transfer to its natural substrate

aglycon (8-demethyl-tetracenomycin C) L-sugars

(L-rhamnose, L-olivose or L-rhodinose), D-sugars

(D-olivose or D-mycarose) and even a D-olivose

disaccharide42–45 (Fig. 6). The transfer of the

D-diolivosyl disaccharide was achieved by expression

of elmGT in the producer of the antitumor drug

mithramycin (S. argillaceus) fed with the precursor

8-demethyl-tetracenomycin C (Ref. 42). Formation of

this diglycosylated derivative did not occur in a

mutant of the mithramycin producer in which all four

mithramycin glycosyltransferases were

simultaneously deleted. However, interestingly, the

diolivosyl derivative was formed through the

simultaneous expression of the genes encoding two

TRENDS in Biotechnology Vol.19 No.11 November 2001

http://tibtech.trends.com

453Review

TRENDS in Biotechnology

O

CH3

OH

CH3

O

O

CH3

CH3

OCH3

O

CH3

OH OH

O CH3

CH3

O

OH N

CH3

CH3CH3

CH3

OH

O

O

OH

OCH3

OH OH

OH

OO

CH3CH3

OOOOH

O

CH3CH3OH

O

Sacc. erythraea eryBIV-minus mutant

OCH3

OH

OOH

O

OHCH3

O CH3

OH

OH

OO

S. fradiae urdR-minus mutantS. argillaceus mtmC-minus mutant

O

O

O

CH3

CH3CH3

CH3CH3

O OOH

CH3OH

OH

S. venezuelae desI-minus mutant

Fig. 4. Someglycosylated derivativesgenerated by inactivationof sugar biosynthesisgenes in the erythromycinproducerSaccharopolysporaerythraea, in themethymycin/pikromycinproducer Streptomycesvenezuelae, in themithramycin producerStreptomyces argillaceusand in the urdamycinproducer Streptomycesfradiae. The proposedfunctions for the deducedproducts of theinactivated genes are:eryBIV, 4-ketoreductase30;desI, 4-dehydrase33;mtmC, 3-C-methyltransferase31;urdR, 4-ketoreductase32.

TRENDS in Biotechnology

S. erythraea BV88 (eryBV–)(+oleGII)

Sugars

oleGIIGTF

3-O-rhamnosyl-6-deoxyerythromycin B

S. erythraea SGT2(eryA–, eryBV–, eryCIII–)

(+tylM2)

Sugars

tylM2

GTF

5-O-desosaminyl-tylactone

O

O

CH3

CH3 O

OH

CH3

CH3

CH3

CH3OH

O

O

CH3

CH3 O

OH

CH3

CH3

CH3

CH3

O OOH N(CH3)2

CH3

O

CH3

OH

CH3

O

O

CH3

CH3

CH3

CH3

CH3OH

OH

(a) (b)

O

CH3

OH

CH3

O

O

CH3

CH3

OCH3

CH3

OOH N(CH3)2

CH3CH3

O

O CH3

OHOH

OH

Tylactone

S. argillaceus M3∆MG(+elmGT, mtmGI, mtmGII)

GTFs

D-Diolivosyl-tetracenomycin C

(c)

CH3 OH O O

O

OH

OH

H3COOC

OCH3

OHHO

elmGT

8-Demethyl-tetracenomycin C

D-olivose

GII GI

D-diolivose

OCH3OHOCH3

OHOH O

CH3 OH O O

OOH

H3COOC

OHO

OH

OCH3

Fig. 5. Generation of glycosylated derivatives by interspecific expression of glycosyltransferases.(a) Formation of an erythromycin derivative in a glycosyltransferase-deficient mutant ofSaccharopolyspora erythraea by expression of the oleandomycin oleGII glycosyltransferase gene38.(b) Formation of a tylosin derivative in a polyketide-synthase- and glycosyltransferase-deficientmutant of S. erythraea by expression of the tylosin tylM2 glycosyltransferase gene40. (c) Formation ofa diglycosylated tetracenomycin C derivative in a glycosyltransferase-deficient mutant ofStreptomyces argillaceus by the simultaneous expression of the mithramycin mtmGI and mtmGIIglycosyltransferase genes, and the elloramycin elmGT glycosyltransferase gene42. The aglycon isendogenously synthesized (a) or exogenously supplied by feeding (b,c).

Page 6: Altering the glycosylation pattern of bioactive compounds

mithramycin transferases (MtmGI and MtmGII) and

the elloramycin transferase (ElmGT) (Fig. 5c). These

two mithramycin glycosyltransferases cooperate in

the formation and transfer of a diolivosyl disaccharide

during the biosynthesis of mithramycin46. Production

of this diglycosylated derivative is the first example of

the use of combinatorial biosynthesis of

glycosyltransferases to create novel glycoside

antibiotics through the cooperative and joint action of

three glycosyltransferases42.

The different sugars transferred by the

glycosyltransferases described in all the examples

shown were incorporated at the position of the

aglycon at which the glycosyltransferase usually

acts. It is an attractive idea to use genetic

engineering to alter the position of sugar

attachment. This has been recently achieved by

making novel hybrid compounds between the

antitumor drugs urdamycin and mithramycin.

UrdGT2 is a glycosyltransferase from the urdamycin

producer S. fradiae Tü2717, which catalyzes the

transfer of D-olivose to the urdamycin aglycon

through a C-glycoside linkage47. The urdGT2 gene

was expressed from a multicopy plasmid in

a mithramycin-non-producing mutant of

S. argillaceus. This mutant, M3∆MG, lacks all the

mithramycin glycosyltransferases48. However, it

contains all the enzymatic machinery necessary for

the biosynthesis of the sugars forming part of the

mithramycin molecule (D-olivose, D-oliose and

D-mycarose) and the ability to synthesize the

mithramycin tetracyclic aglycon. Introduction of the

urdGT2 gene in this mutant caused the formation of

two novel compounds containing the mithramycin

aglycon 4-demethyl-premithramycinone and either

D-olivose or D-mycarose attached through a

C-glycoside linkage to a position other than that at

which mithramycin is usually glycosylated

(A. Trefzer et al., unpublished). The fact that a

different aglycon (4-demethyl-premithramycinone)

and a different 6DOH (D-mycarose) were substrates

for UrdGT2 is another example of the flexibility of

antibiotic glycosyltransferases.

Using nonproducing hosts as cell factories

Appropriate recombinant strains can be constructed

using genetic manipulation to give them the ability

to produce the necessary antibiotic building blocks.

Consequently, they can be used as alternative,

useful cell factories for producing novel compounds.

In these recombinant strains, the aglycon can be

synthesized by the host strain through the

incorporation of a plasmid harboring the genes

necessary for aglycon biosynthesis. Alternatively,

the recombinant strain can be used as a

biotransformation host and the aglycon added to the

culture medium. Selected glycosyltransferases and

the ability to synthesize sugars can be supplied to

the host strains by transformation with one or more

plasmids harboring the corresponding gene sets.

Some plasmids directing the biosynthesis of

different deoxyhexoses have been developed during

the past few years16,17,45 as important tools for the

generation of new glycosylated compounds.

Several outstanding approaches have been

undertaken in this area. The macrolide antibiotic

oleandomycin contains the neutral sugar

L-oleandrose in its molecule and all the genes

involved in its biosynthesis have been identified17,39.

Using different subsets of L-oleandrose biosynthesis

genes, three plasmids were constructed, each

containing the genes encoding the enzymatic

functions required to direct the biosynthesis of a

different sugar, L-rhamnose45, L-olivose or

L-oleandrose17. A biotransformation host was

constructed using Streptomyces albus in which the

oleGII glycosyltransferase gene was stably integrated

into the chromosome. When the plasmids coding for

TRENDS in Biotechnology Vol.19 No.11 November 2001

http://tibtech.trends.com

454 Review

TRENDS in Biotechnology

CH3 OH O O

O

OR1

OH

H3COOC

OCH3

OH

OR2

OCH3

OH

OHOH

OCH3

OH

OCH3

OH

OH

OCH3

OH

MeO

OCH3

MeO

MeO

OCH3

OHOH

OCH3

OH

OH

CH3

OCH3

OHOCH3

OHOH

O

R2 R2

L-Rhamnose

L-Rhodinose

L-Olivose

L-Oleandrose

3,4-Dimethoxy-L-olivose

D-Olivose

D-Mycarose

D-Diolivose

Fig. 6. Structures of8-demethyl-tetracenomycin Cderivatives generated withthe participation of theElmGT glycosyltransferasefrom Streptomycesolivaceus. The ElmGTglycosyltransferase hasbeen shown to be able torecognize and transferdifferent monosaccharidesand a disaccharide to theelloramycin aglycon42,43,45.

TRENDS in Biotechnology

S. albus(+oleGII, pOLE)

L-Oleandrose

GTF

pOLEoleGII

L-Oleandrosyl-erythronolide B

O

CH3

OH

CH3

O

O

CH3

CH3

CH3

CH3

CH3OH

OH

OH

(a)

Erythronolide B

CH3 OH O O

O

OH

OH

H3COOC

OCH3OH

OH

S. albus(+elmGT, pOLV)

L-Olivose

GTF

pOLVelmGT

L-Olivosyl-tetracenomycin C

(b)

8-Demethyl-tetracenomycin C

O

CH3

OH

CH3

O

O

CH3

CH3

CH3

CH3

CH3O

OH

O CH3OHOMe

OOH

CH3 OH O O

OOH

H3COOC

OHO

OH

OCH3

OH

CH3

Fig. 7. The use of genetically engineered biotransformation hosts for the generation of novelglycosylated derivatives. (a) Formation of a glycosylated derivative from erythronolide B in aStreptomyces albus strain harboring the oleandomycin oleGII glycosyltransferase gene and aplasmid (pOLE) directing the biosynthesis of L-oleandrose17. (b) Formation of a tetracenomycin Cderivative in an S. albus strain harboring the elloramycin elmGT glycosyltransferase gene and aplasmid (pOLV) directing the biosynthesis of L-olivose42. In both cases, the precursor was suppliedby feeding.

Page 7: Altering the glycosylation pattern of bioactive compounds

References

1 Demain, A.L. (1999) Pharmaceutically active

secondary metabolites of microorganisms. Appl.

Microbiol. Biotechnol. 52, 455–463

2 Hopwood, D.A. (1997) Genetic contribution to

understanding polyketide synthases. Chem. Rev.

97, 2465–2497

3 Khosla, C. and Zawada, R.J.X. (1996)

Generation of polyketide libraries via

combinatorial biosynthesis. Trends Biotechnol.

14, 335–341

4 Leadlay, P.F. (1997) Combinatorial approaches to

polyketide biosynthesis. Curr. Opin. Chem. Biol.

1, 162–168

5 Hutchinson, C.R. (1998) Combinatorial

biosynthesis for new drug discovery. Curr. Opin.

Microbiol. 1, 319–329

6 Salas, J.A. and Méndez, C. (1998) Genetic

manipulation of antitumor-agents biosynthesis:

a potential method for developing novel drugs.

Trends Biotechnol. 16, 475–482

7 Bechthold, A. and Salas, J.A. (1999)

Combinatorial biosynthesis of microbial

metabolites. In Combinatorial Organic Chemistry

(Jung, G., ed.), pp. 381–407, Wiley-VCH

8 Méndez, C. et al. (2001) Structure alteration

of polyketides by recombinant DNA

technology in producer organisms –

prospects for the generation of novel

pharmaceutical drugs. Curr. Pharm. Biotechnol.

1, 355–395

9 Weymouth-Wilson, A.C. (1997) The role of

carbohydrates in biologically active natural

products. Nat. Prod. Rep. 14, 99–110

10 Kirschning, A. et al. (1997) Chemical and

biochemical aspects of deoxysugars and

therefrom derived oligosaccharides. Top. Curr.

Chem. 188, 1–84

11 Stockmann, M. and Piepersberg, W. (1992)

Gene probes for the detection of 6-deoxyhexose

metabolism in secondary metabolite-producing

streptomycetes. FEMS Microbiol. Lett. 90,

185–190

12 Trefzer, A. et al. (1999) Genes and enzymes of

deoxysugar biosyntheses. Nat. Prod. Rep. 16,

283–299

13 Liu, H-W. and Thorson, J.S. (1994) Pathways and

mechanisms in the biogenesis of novel

deoxysugars by bacteria. Annu. Rev. Microbiol.

48, 223–256

14 Piepersberg, W. (1994) Pathway engineering in

secondary metabolite-producing actinomycetes.

Crit. Rev. Biotechnol. 14, 251–285

15 Hallis, T.M. and Liu, H-W. (1999) Learning

nature’s strategies for making deoxy sugars:

pathways, mechanisms and combinatorial

applications. Acc. Chem. Res. 32, 579–588

16 Olano, C. et al. (1999) A two-plasmid system for

the glycosylation of polyketide antibiotics:

bioconversion of epsilon-rhodomycinone to

rhodomycin D. Chem. Biol. 6, 845–855

17 Aguirrezabalaga, I. et al. (2000) Identification and

expression of genes involved in biosynthesis of

L-oleandrose and its intermediate L-olivose in the

oleandomycin producer Streptomyces

antibioticus. Antimicrob. Agents Chemother. 44,

1266–1275

18 Wohlert, S-E. et al. (2001) Insights about the

biosynthesis of the avermectin deoxysugar

L-oleandrose through heterologous expression of

Streptomyces avermitilis deoxysugar genes in

Streptomyces lividans. Chem. Biol. 8, 681–700

19 Merson-Davies, L.A. and Cundliffe, E. (1994)

Analysis of five tylosin biosynthetic genes from

the tyllBA region of the Streptomyces fradiae

genome. Mol. Microbiol. 13, 349–355

20 Lombó, F. et al. (1997) Cloning and insertional

inactivation of Streptomyces argillaceus genes

involved in earliest steps of sugar biosynthesis

of the antitumor polyketide mithramycin.

J. Bacteriol. 179, 3354–3357

21 Draeger, G. et al. (1999) Mechanism of the

2-deoxygenation step in the biosynthesis of the

deoxyhexose moieties of the antibiotics

granaticin and oleandomycin. J. Am. Chem Soc.

121, 2611–2612

22 Chen, H. et al. (1999) Biosynthesis of mycarose:

isolation and characterization of enzymes

involved in the C-2 deoxygenation. J. Am. Chem.

Soc. 121, 8124–8125

23 Chen, H. et al. (2000) Deoxysugars in

glycopeptide antibiotics: enzymatic synthesis of

TDP–L-epivancosamine in chloroeremomycin

biosynthesis. Proc. Natl. Acad. Sci. U. S. A. 97,

11942–11947

24 Campbell, J.A. et al. (1997) A classification of

nucleotide–diphospho sugar

glycosyltransferases based on amino acid

sequence similarities. Biochem. J. 326,

929–942

25 Hoffmeister, D. et al. (2001) Two sequence

elements of glycosyltransferases involved in

urdamycin biosynthesis are responsible for

substrate specificity and enzymatic activity.

Chem. Biol. 8, 557–567

26 Quirós, L.M. et al. (1998) Two

glycosyltransferases and a glycosidase are

involved in oleandomycin modification during its

biosynthesis by Streptomyces antibioticus. Mol.

Microbiol. 28, 1177–1186

27 Gourmelen, A. et al. (1998) Characterization of

a glycosyl transferase inactivating macrolides,

encoded by gimA from Streptomyces

ambofaciens. Antimicrob. Agents Chemother. 42,

2612–2619

28 Mulichak, A.M. et al. (2001) Structure of the

UDP–glucosyltransferase GtfB that modifies

the heptapeptide aglycone in the biosynthesis of

vancomycin group antibiotics. Structure 9,

547–557

29 Gaisser, S. et al. (1997) Analysis of seven genes

from the eryAI–eryK region of the erythromycin

biosynthetic gene cluster in Saccharopolyspora

erythraea. Mol. Gen. Genet. 256, 239–251

30 Salah-Bey, K. et al. (1998) Targeted gene

inactivation for the elucidation of deoxysugar

biosynthesis in the erythromycin producer

Saccharopolyspora erythraea. Mol. Gen. Genet.

257, 542–553

TRENDS in Biotechnology Vol.19 No.11 November 2001

http://tibtech.trends.com

455Review

the biosynthesis of L-olivose and L-oleandrose were

introduced into this recombinant host, the

erythromycin aglycon erythronolide B was converted

into monoglycosylated derivatives after feeding this

compound to the recombinant strains (Fig. 7a).

The three plasmids in combination with the

ElmGT glycosyltransferase described above were

also used to produce novel hybrid elloramycin

analogs45. In this case, coexpression in Streptomyces

lividans of cosmid 16F4 (coding for the biosynthesis

of the elloramycin intermediate 8-demethyl-

tetracenomycin and containing the elmGT gene) and

any of these three sugar plasmids produced three

new compounds: 8-demethyl-8-α-L-olivosyl-

tetracenomycin C, 8-demethyl-8-α-L-oleandrosyl-

tetracenomycin C and 2′-demethoxy-elloramycin45,

in which the sugar moieties were encoded by the

macrolide oleandomycin genes and the aglycon and

sugar transfer were encoded by the elloramycin

cluster. The ElmGT glycosyltransferase described

above is responsible for the incorporation of the

different sugars, as was confirmed by the formation

of L-olivosyl-tetracenomycin C in a S. albus

recombinant strain containing an L-olivose-

synthesizing plasmid and expressing elmGT when

fed with 8-demethyl-tetracenomycin C (Fig. 7b).

Conclusions

In the near future, glycosylation could be an

attractive target for generating novel antibiotics

because the specific attachment of one or more sugars

is required for the activity of many bioactive

compounds. Understanding the functionality of the

different domains of glycosyltransferases might allow

engineering of the domains that contribute to

substrate specificity, thus making it possible to alter

the specificity of the glycosyltransferases. Recent

advances in the understanding of antibiotic

glycosyltransferases show that substrate flexibility

(mainly related to the sugar transferred) is fairly

common in antibiotic glycosyltransferases from

producer organisms. The availability of in vitro

enzymatic systems for the assay of

glycosyltransferases will also contribute to the

possible enzymatic synthesis of a wide variety of

glycosylated derivatives.

Acknowledgements

Work in our laboratorywas supported by grantsof the Spanish Ministry ofEducation and Sciencethrough the Plan Nacionalen Biotecnologia(BIO97-0771), Plan FEDER(1FD97-0040) and theEuropean Union(BIO4-CT96-0080 andQLK3-CT-1999-00095).We thank all the peoplethat have worked in ourlaboratory for helpfulsuggestions anddiscussions, especiallythose contributing toresearch on sugarbiosynthesis andglycosylation.

Page 8: Altering the glycosylation pattern of bioactive compounds

31 González, A. et al. (2001) The mtmVUC genes of

the mithramycin gene cluster in Streptomyces

argillaceus are involved in the biosynthesis of the

sugar moieties. Mol. Gen. Genet. 264, 827–835

32 Hoffmeister, D. et al. (2000) The NDP–sugar

co-substrate concentration and the enzyme

expression level influence the substrate specificity

of glycosyltransferases: cloning and

characterization of deoxysugar biosynthetic

genes of the urdamycin biosynthetic gene cluster.

Chem. Biol. 7, 821–831

33 Zhao, L. et al. (1998) Mechanistic studies of

desosamine biosynthesis: C-4 deoxygenation

precedes C-3 transamination. J. Am. Chem. Soc.

120, 12159–12160

34 Zhao, L. et al. (1999) Engineering a

methymycin/pikromycin–calicheamicin hybrid:

construction of two new macrolides carrying a

designed sugar moiety. J. Am. Chem. Soc. 121,

9881–9882

35 Borisova, S.A. et al. (1999) Biosynthesis of

desosamine: construction of a new macrolide

carrying a genetically designed sugar moiety.

Org. Chem. 15, 133–136

36 Madduri, K. et al. (1998) Production of the

antitumor drug epirubicin (4′-epidoxorubicin)

and its precursor by a genetically engineered

strain of Streptomyces peucetius. Nat. Biotechnol.

16, 69–74

37 Solenberg, P.J. et al. (1997) Production of

hybrid glycopeptide antibiotics in vitro and in

Streptomyces toyocaensis. Chem. Biol. 4,

195–202

38 Doumith, M. et al. (1999) Interspecies

complementation in Saccharopolyspora

erythraea: elucidation of the function of oleP1,

oleG1 and oleG2 from the oleandomycin

biosynthetic gene cluster of Streptomyces

antibioticus and generation of new erythromycin

derivatives. Mol. Microbiol. 34, 1039–1048

39 Olano, C. et al. (1998) Analysis of a Streptomyces

antibioticus chromosomal region involved in

oleandomycin biosynthesis that contains two

glycosyltransferases responsible for

glycosylation of the macrolactone ring. Mol. Gen.

Genet. 259, 299–308

40 Gaisser, S. et al. (2000) A defined system for

hybrid macrolide biosynthesis in

Saccharopolyspora erythraea. Mol. Microbiol.

36, 391–401

41 Gandecha, A.R. et al. (1997) Analysis of four

tylosin biosynthetic genes from the tylLM region

of the Streptomyces fradiae genome. Gene 184,

197–203

42 Blanco, G. et al. (2001) Identification of a

sugar flexible glycosyltransferase from

Streptomyces olivaceus, the producer of the

antitumor polyketide elloramycin. Chem. Biol.

8, 253–263

43 Decker, H. et al. (1995) Novel genetically

engineered tetracenomycins. Angew. Chem., Int.

Ed. Engl. 34, 1107–1110

44 Wohlert, S.E. et al. (1998) Novel hybrid

tetracenomycins through combinatorial

biosynthesis using a glycosyltransferase encoded

by the elm-genes in cosmid 16F4 which shows a

broad sugar substrate specificity. J. Am. Chem.

Soc. 120, 10596–10601

45 Rodriguez, L. et al. (2000) Generation of hybrid

elloramycin analogs by combinatorial

biosynthesis using genes from anthracycline-

type and macrolide biosynthetic pathways.

J. Mol. Microbiol. Biotechnol. 2, 271–276

46 Fernández, E. et al. (1998) Identification of two

genes from Streptomyces argillaceus encoding

two glycosyltransferases involved in the transfer

of a disaccharide during the biosynthesis of the

antitumor drug mithramycin. J. Bacteriol. 180,

4929–4937

47 Faust, B. et al. (2000) Two new tailoring enzymes,

a glycosyltransferase and an oxygenase, involved

in biosynthesis of the angucycline antibiotic

urdamycin A in Streptomyces fradiae Tü2717.

Microbiol. UK 146, 147–154

48 Prado, L. et al. (1999) Oxidative cleavage of

premithramycin B is one of the last steps in the

biosynthesis of the antitumor drug mithramycin.

Chem. Biol. 6, 19–30

49 Haydock, S.F. et al. (1991) Cloning and sequence

analysis of genes involved in erythromycin

biosynthesis in Saccharopolyspora erythraea:

sequence similarities between EryG and a

family of S-adenosylmethionine-dependent

methyltransferases. Mol. Gen. Genet. 230,

120–128

50 Summers, R.G. et al. (1997) Sequencing and

mutagenesis of genes from the erythromycin

biosynthetic gene cluster of Saccharopolyspora

erythraea that are involved in L-mycarose and

D-desosamine production. Microbiology 143,

3251–3262

51 Xue, Y. et al. (1998) A gene cluster for macrolide

antibiotic biosynthesis in Streptomyces

venezuelae: architecture of metabolic diversity.

Proc. Natl. Acad. Sci. U. S. A. 95, 12111–12116

52 Volchegursky, Y. et al. (2000) Biosynthesis of the

anti-parasitic agent megalomicin:

transformation of erythromycin to megalomicin

in Saccharopolyspora erythraea. Mol. Microbiol.

37, 752–762

53 Westrich, L. et al. (1999) Cloning and

characterization of a gene cluster from

Streptomyces cyanogenus S136 probably involved

in landomycin biosynthesis. FEMS Microbiol.

Lett. 15, 381–387

54 Fouces, R. et al. (1999) The tylosin biosynthetic

cluster from Streptomyces fradiae: genetic

organization of the left region. Microbiology

145, 855–868

55 Brautaset, T. et al. (2000) Biosynthesis of the

polyene antifungal antibiotic nystatin in

Streptomyces nourseiATCC 11455: analysis of the

gene cluster and deduction of the biosynthetic

pathway. Chem. Biol. 7, 395–403

56 Ikeda, H. et al. (1999) Organization of the

biosynthetic gene cluster for the polyketide

anthelmintic macrolide avermectin in

Streptomyces avermitilis. Proc. Natl. Acad. Sci.

U. S. A. 96, 9509–9514

57 Steffensky, M. et al. (2000) Identification of the

novobiocin biosynthetic gene cluster of

Streptomyces spheroides NCIB 11891.

Antimicrob. Agents Chemother. 44, 1214–1222

58 Ichinose, K. et al. (1998) The granaticin

biosynthetic gene cluster of Streptomyces

violaceoruber Tu22: sequence analysis and

expression in a heterologous host. Chem. Biol.

5, 647–659

59 Lomovskaya, N. et al. (1999) Doxorubicin

overproduction in Streptomyces peucetius: cloning

and characterization of the dnrU ketoreductase

and dnrV genes and the doxA cytochrome P-450

hydroxylase gene. J. Bacteriol. 181, 305–318

60 Otten, S.L. et al. (1997) Cloning and

characterization of the Streptomyces peucetius

dnmZUV genes encoding three enzymes required

for biosynthesis of the daunorubicin precursor

thymidine diphospho-L-daunosamine.

J. Bacteriol. 179, 4446–4450

61 Torkkell, S. et al. (1997) Characterization of

Streptomyces nogalater genes encoding enzymes

involved in glycosylation steps in nogalamycin

biosynthesis. Mol. Gen. Genet. 256, 203–209

62 Räty, K. et al. (2000) A gene cluster from

Streptomyces galilaeus involved in glycosylation

of aclarubicin. Mol. Gen. Genet. 264, 164–172

63 van Wageningen, A.M. et al. (1998) Sequencing

and analysis of genes involved in the

biosynthesis of a vancomycin group antibiotic.

Chem. Biol. 5, 155–162

64 Aparicio, J.F. et al. (2000) A complex

multienzyme system encoded by five polyketide

synthase genes is involved in the biosynthesis of

the 26-membered polyene macrolide pimaricin

in Streptomyces natalensis. Chem. Biol. 7,

895–905

65 Pelzer, S. et al. (1999) Identification and analysis

of the balhimycin biosynthetic gene cluster and

its use for manipulating glycopeptide

biosynthesis in Amycolatopsis mediterranei

DSM5908. Antimicrob. Agents Chemother. 43,

1565–1573

66 Bate, N. et al. (2000) The mycarose-biosynthetic

genes of Streptomyces fradiae, producer of tylosin.

Microbiology 146, 139–146

67 Trefzer, A. et al. (2000) Function of

glycosyltransferase genes involved in urdamycin

A biosynthesis. Chem. Biol. 7, 33–42

68 Wang, Z.X. et al. (2000) Identification of the

coumermycin A1 biosynthetic gene cluster of

Streptomyces rishiriensis DSM 40489.

Antimicrob. Agents Chemother. 44, 3040–3048

69 Blanco, G. et al. (2000) Characterization of two

glycosyltransferases involved in early

glycosylation steps during biosynthesis of the

antitumor polyketide mithramycin by

Streptomyces argillaceus. Mol. Gen. Genet. 262,

991–1000

70 Dickens, M.L. et al. (1996) Cloning, sequencing,

and analysis of aklaviketone reductase from

Streptomyces sp. strain C5. J. Bacteriol. 178,

3384–3388

71 Otten, S.L. et al. (1995) Cloning and

characterization of the Streptomyces peucetius

dnrQS genes encoding a daunosamine

biosynthesis enzyme and a glycosyl transferase

involved in daunorubicin biosynthesis.

J. Bacteriol. 177, 6688–6692

72 Gourmelen, A. et al. (1998) Characterization of a

glycosyl transferase inactivating macrolides,

encoded by gimA from Streptomyces

ambofaciens. Antimicrob. Agents Chemother. 42,

2612–2619

73 Ohuchi, T. et al. (2000) Cloning and expression of

a gene encoding N-glycosyltransferase (ngt) from

Saccharothrix aerocolonigenesATCC39243.

J. Antibiot. 53, 393–403

74 Barrasa, M.I. et al. (1997) The aminonucleoside

antibiotic A201A is inactivated by a

phosphotransferase activity from Streptomyces

capreolus NRRL 3817, the producing organism.

Isolation and molecular characterization of the

relevant encoding gene and its DNA flanking

regions. Eur. J. Biochem. 245, 54–63

75 Mao, Y. et al. (1999) Genetic localization and

molecular characterization of two key genes

(mitAB) required for biosynthesis of the

antitumor antibiotic mitomycin C. J. Bacteriol.

181, 2199–2208

TRENDS in Biotechnology Vol.19 No.11 November 2001

http://tibtech.trends.com

456 Review