Chemistry & Biology Article to central pyridine moiety formation. Substi-tution of the endogenous...

10
Chemistry & Biology Article Genome Mining Reveals a Minimum Gene Set for the Biosynthesis of 32-Membered Macrocyclic Thiopeptides Lactazoles Shohei Hayashi, 1 Taro Ozaki, 1 Shumpei Asamizu, 1 Haruo Ikeda, 2 Satoshi Omura, 3 Naoya Oku, 4 Yasuhiro Igarashi, 4 Hiroshi Tomoda, 5 and Hiroyasu Onaka 1,4, * 1 Department of Biotechnology, Graduate School of Agricultural and Life Sciences, The University of Tokyo, Tokyo 113-8657, Japan 2 Kitasato Institute for Life Sciences, Kitasato University, Sagamihara 228-8555, Japan 3 The Kitasato Institute, Kitasato University, Tokyo 108-8642, Japan 4 Biotechnology Research Center, Toyama Prefectural University, Imizu 939-0398, Japan 5 Graduate School of Pharmaceutical Sciences, Kitasato University, Tokyo 108-8641, Japan *Correspondence: [email protected] http://dx.doi.org/10.1016/j.chembiol.2014.03.008 SUMMARY Although >100 thiopeptides have been discovered, the number of validated gene clusters involved in their biosynthesis is lagging. We use genome mining to identify a silent thiopeptide biosynthetic gene cluster responsible for biosynthesis of lactazoles. Lactazoles are structurally unique thiopeptides with a 32-membered macrocycle and a 2-oxazolyl-6-thia- zolyl pyridine core. We demonstrate that lactazoles originate from the simplest cluster, containing only six unidirectional genes (lazA to lazF). We show that lazC is involved in the macrocyclization process, leading to central pyridine moiety formation. Substi- tution of the endogenous promoter with a strong promoter results in an approximately 30-fold in- crease in lactazole A production and mutagenesis of lazC precursor gene in production of two analogs. Lactazoles do not exhibit antimicrobial activity but may modulate signaling cascades triggered by bone morphogenetic protein. Our approach facili- tates the production of a more diverse set of thio- peptide structures, increasing the semisynthetic repertoire for use in drug development. INTRODUCTION Microorganisms produce a large variety of ribosomally synthe- sized and posttranslationally modified peptides (RiPPs). This structurally diverse set of natural product macromolecules in- cludes thiopeptides, linear azole/azoline-containing peptides (LAPs), lanthipeptides, cyanobactins, and microcins (Arnison et al., 2013). In RiPP biosynthesis, precursor peptide genes are transcribed and translated into precursor peptides, which then undergo posttranslational modification to form disulfides, dehy- droamino acids, or heterocycles. In most cases, a leader peptide in the N terminus of the precursor peptide is required for the recruitment of modifying enzymes; this peptide is removed in the final stage of biosynthesis. RiPPs exhibit a wide spectrum of biological properties, including antimicrobial, anticancer, and antiviral activities (Arnison et al., 2013). The modifications of the peptide backbone confer bioactivity and stability and also protect the peptides from proteolysis (Walsh and Nolan, 2008). The genetic encoding of RiPP precursor peptides can be exploited by altering triplet sequences of codons in order to make novel amino acid substitutions. This can be incorporated into drug development strategies that use peptide compounds. The Streptomyces genus is a rich source of RiPPs such as thiopeptides, LAPs, lanthipeptides, and sactipeptides. Goad- sporin, which is the only LAP isolated from actinomycetes, is produced by Streptomyces sp. TP-A0584. This peptide, which consists of 19 amino acids containing six azole rings and two de- hydroalanines, promotes secondary metabolism and morpho- genesis (Onaka et al., 2001). We previously identified the goadsporin biosynthetic gene cluster of Streptomyces sp. TP- A0584 (Onaka et al., 2005). This was the first example of an acti- nomycete biosynthetic gene cluster for RiPPs that was capable of producing both azole and dehydroamino acids. The god clus- ter contains a class I lanthipeptide dehydratase homolog for de- hydroalanine formation that is also found in all known thiopeptide biosynthetic gene clusters. Thiopeptides are produced mainly by actinomycetes and typi- cally contain highly modified sulfur-containing peptides, which have a characteristic macrocycle bearing a six-membered nitro- gen-containing ring (e.g., pyridine or piperidine). During biosyn- thesis of thiopeptides such as goadsporin, the precursor peptide is modified to form dehydroalanines/dehydrobutyrines and azole/azoline rings. Subsequently, the nitrogen-containing ring is formed from two dehydroalanines in the core peptide via cycloaddition (Bycroft and Gowland, 1978). Further tailoring modifications lead to structural diversification of thiopeptides. The majority of thiopeptides inhibit the growth of Gram-positive bacteria, including methicillin-resistant Staphylococcus aureus and vancomycin-resistant enterococci (Walsh et al., 2010). To date, more than 100 thiopeptides have been discovered, and several biosynthetic gene clusters have been identified (Bagley et al., 2005; Li and Kelly, 2010). In addition, the progress of genome sequencing techniques has triggered exponential growth of genome databases, and numerous additional cryptic Chemistry & Biology 21, 679–688, May 22, 2014 ª2014 Elsevier Ltd All rights reserved 679

Transcript of Chemistry & Biology Article to central pyridine moiety formation. Substi-tution of the endogenous...

Chemistry & Biology

Article

Genome Mining Reveals a Minimum Gene Setfor the Biosynthesis of 32-MemberedMacrocyclic Thiopeptides LactazolesShohei Hayashi,1 Taro Ozaki,1 Shumpei Asamizu,1 Haruo Ikeda,2 Satoshi �Omura,3 Naoya Oku,4 Yasuhiro Igarashi,4

Hiroshi Tomoda,5 and Hiroyasu Onaka1,4,*1Department of Biotechnology, Graduate School of Agricultural and Life Sciences, The University of Tokyo, Tokyo 113-8657, Japan2Kitasato Institute for Life Sciences, Kitasato University, Sagamihara 228-8555, Japan3The Kitasato Institute, Kitasato University, Tokyo 108-8642, Japan4Biotechnology Research Center, Toyama Prefectural University, Imizu 939-0398, Japan5Graduate School of Pharmaceutical Sciences, Kitasato University, Tokyo 108-8641, Japan

*Correspondence: [email protected]://dx.doi.org/10.1016/j.chembiol.2014.03.008

SUMMARY

Although >100 thiopeptides have been discovered,the number of validated gene clusters involved intheir biosynthesis is lagging. We use genome miningto identify a silent thiopeptide biosynthetic genecluster responsible for biosynthesis of lactazoles.Lactazoles are structurally unique thiopeptides witha 32-membered macrocycle and a 2-oxazolyl-6-thia-zolyl pyridine core. We demonstrate that lactazolesoriginate from the simplest cluster, containing onlysix unidirectional genes (lazA to lazF). We showthat lazC is involved in the macrocyclization process,leading to central pyridine moiety formation. Substi-tution of the endogenous promoter with a strongpromoter results in an approximately 30-fold in-crease in lactazole A production and mutagenesisof lazC precursor gene in production of two analogs.Lactazoles do not exhibit antimicrobial activitybut may modulate signaling cascades triggered bybone morphogenetic protein. Our approach facili-tates the production of a more diverse set of thio-peptide structures, increasing the semisyntheticrepertoire for use in drug development.

INTRODUCTION

Microorganisms produce a large variety of ribosomally synthe-

sized and posttranslationally modified peptides (RiPPs). This

structurally diverse set of natural product macromolecules in-

cludes thiopeptides, linear azole/azoline-containing peptides

(LAPs), lanthipeptides, cyanobactins, and microcins (Arnison

et al., 2013). In RiPP biosynthesis, precursor peptide genes are

transcribed and translated into precursor peptides, which then

undergo posttranslational modification to form disulfides, dehy-

droamino acids, or heterocycles. Inmost cases, a leader peptide

in the N terminus of the precursor peptide is required for the

recruitment of modifying enzymes; this peptide is removed in

Chemistry & Biology 21,

the final stage of biosynthesis. RiPPs exhibit a wide spectrum

of biological properties, including antimicrobial, anticancer,

and antiviral activities (Arnison et al., 2013). The modifications

of the peptide backbone confer bioactivity and stability and

also protect the peptides from proteolysis (Walsh and Nolan,

2008). The genetic encoding of RiPP precursor peptides can

be exploited by altering triplet sequences of codons in order to

make novel amino acid substitutions. This can be incorporated

into drug development strategies that use peptide compounds.

The Streptomyces genus is a rich source of RiPPs such

as thiopeptides, LAPs, lanthipeptides, and sactipeptides. Goad-

sporin, which is the only LAP isolated from actinomycetes, is

produced by Streptomyces sp. TP-A0584. This peptide, which

consists of 19 amino acids containing six azole rings and two de-

hydroalanines, promotes secondary metabolism and morpho-

genesis (Onaka et al., 2001). We previously identified the

goadsporin biosynthetic gene cluster of Streptomyces sp. TP-

A0584 (Onaka et al., 2005). This was the first example of an acti-

nomycete biosynthetic gene cluster for RiPPs that was capable

of producing both azole and dehydroamino acids. The god clus-

ter contains a class I lanthipeptide dehydratase homolog for de-

hydroalanine formation that is also found in all known thiopeptide

biosynthetic gene clusters.

Thiopeptides are produced mainly by actinomycetes and typi-

cally contain highly modified sulfur-containing peptides, which

have a characteristic macrocycle bearing a six-membered nitro-

gen-containing ring (e.g., pyridine or piperidine). During biosyn-

thesis of thiopeptides such as goadsporin, the precursor peptide

is modified to form dehydroalanines/dehydrobutyrines and

azole/azoline rings. Subsequently, the nitrogen-containing ring

is formed from two dehydroalanines in the core peptide via

cycloaddition (Bycroft and Gowland, 1978). Further tailoring

modifications lead to structural diversification of thiopeptides.

The majority of thiopeptides inhibit the growth of Gram-positive

bacteria, including methicillin-resistant Staphylococcus aureus

and vancomycin-resistant enterococci (Walsh et al., 2010).

To date, more than 100 thiopeptides have been discovered,

and several biosynthetic gene clusters have been identified

(Bagley et al., 2005; Li and Kelly, 2010). In addition, the progress

of genome sequencing techniques has triggered exponential

growth of genome databases, and numerous additional cryptic

679–688, May 22, 2014 ª2014 Elsevier Ltd All rights reserved 679

hiostrepton(tsr) A B C D E F G H I J K L M N O P Q R S T U

nocathiacin(noc) P S1 B QAS2 S3 S4 S5 S6 T U R V C D E F G H I K L M N O

nosiheptide(nos) A B C D E F G H I J K L M N O P

lactazole (laz) A B C D E F

GE37468 (get) A B C D E F G H I J K L M

GE2270 (tpd-GE) R O P Q S T U A B C D E F G K L M N

thiomuracin (tpd-TM) I J1 J2 R A B C D E F G H

2.5 5.0 7.5 10.0 12.50 15.0 17.5 22.5 25.020.0 30.027.5 32.5kb

thiocillin(tcl) ABC DE-H I J K L M N O P QR S T U V W X

A B C D E F G H Rorf12 3 Igoadsporin

(god)

TP-1161 (tpa) A B C D E F G H I J K LX

cyclothiazomycin(clt) Q P O NM I H G F E B A C D JK L

siomycin(sio) A B C D E F G H I J K L M N O P Q R S T U

Precursor peptide Lantibiotic dehydration Dehydration

Function unknown No relation

Transportation Host resistance DeaminationRegulation Side ring formation Other tailoring Central core ring formation Dehydrogenation CyclodehydrationE1 E2 G1G2 D A B C H I J

berninamycin(ber)

Figure 1. Gene Organization of the lazClus-

ter and Comparison with Other Thiopeptide

Biosynthetic Gene Clusters

Lactazole (laz), TP-1161 (tpa) (Engelhardt et al.,

2010), GE37468 (get) (Young and Walsh, 2011),

GE2270A (tpd_GE) (Morris et al., 2009), thio-

strepton (tsr) (Liao et al., 2009), cyclothiazomycin

(clt) (Wang et al., 2010), thiomuracins (tpd_TM)

(Morris et al., 2009), thiocillin (tcl) (Wieland Brown

et al., 2009), siomycin (sio) (Liao et al., 2009),

berninamycin (ber) (Malcolmson et al., 2013), no-

siheptide (nos) (Yu et al., 2009), nocathiacin (noc)

(Ding et al., 2010), and goadsporin (god) (Onaka

et al., 2005). The deduced functions are labeled

in color.

Chemistry & Biology

A Minimum Gene Set for Thiopeptide Biosynthesis

thiopeptide synthesis gene clusters have been identified in a

wide range of microorganisms (Li et al., 2012; Wieland Brown

et al., 2009). Recently, the development of convenient vectors

and host systems in the Streptomyces genus has facilitated

the identification of metabolites produced from several silent

gene clusters. However, there is no precedent for the identifica-

tion of thiopeptides from such clusters.

In this study, we identified a silent thiopeptide biosynthetic

gene cluster in the genome of Streptomyces lactacystinaeus

OM-6519, a lactacystin producer (Omura et al., 1991), and

isolated lactazole A, B, and C by heterologous expression in

Streptomyces lividans TK23. Lactazoles are structurally unique

thiopeptides that can be biosynthesized using only six unidirec-

tional genes spanning a 9.8-kb region of DNA. We discuss how

this concise biosynthetic gene cluster can be exploited to in-

crease the diversity and yield of thiopeptide analogs.

RESULTS

Genome Mining and Heterologous Expression of aCryptic Biosynthetic Gene Cluster for ThiopeptidesWe used the sequence of godF, a class I lanthipeptide dehydra-

tase encoded by the goadsporin biosynthetic gene cluster, to

identify a putative thiopeptide biosynthetic gene cluster from

a draft genome sequence of S. lactacystinaeus OM-6519. This

cluster also contained the godD, godE, and godG homologs

that encode the enzymes of cyclodehydratase, nitroreductase,

and SpaB homolog, respectively; these enzymes are required

for dehydroalanine and azole synthesis. Additionally, a putative

precursor peptide gene was present in the upstream region.

Analysis of this region using FramePlot revealed that the cluster

contained six open reading frames (ORFs), which we later desig-

nated lazA, lazB, lazC, lazD, lazE, and lazF (Figure 1).

680 Chemistry & Biology 21, 679–688, May 22, 2014 ª2014 Elsevier Ltd All rights reserved

Because S. lactacystinaeus OM-6519

did not produce metabolites related

to thiopeptides, we decided to transfer

the putative thiopeptide biosynthetic

gene cluster into S. lividans TK23 as

a surrogate host. We constructed a

pKU465cos-based cosmid library of the

S. lactacystinaeus OM-6519 genome

and isolated cosmid pKU465-ltc18-6C,

which contains the putative biosynthetic

gene cluster. S. lividans TK23 was transformed with this cosmid

to generate S. lividans ltc18-6C. Comparative high-performance

liquid chromatography (HPLC) analysis of the fermentation

broth of S. lividans ltc18-6C and S. lividans transformed with

pKU465cos revealed a specific peak at 11.0 min that was pre-

sent only in the mutant strain (Figure 2).

Isolation and Structural Elucidation of LactazolesFor large-scale preparation, mycelium collected from 3 l of

fermentation broth was extracted with methanol, followed by

two steps of liquid-liquid separation. Subsequently, size-exclu-

sion chromatography (LH-20), reverse-phase chromatography,

and preparative HPLC were performed to give pure compounds.

Finally, we isolated three thiopeptides, namely, lactazole A

(6.7 mg), lactazole B (2.7 mg), and lactazole C (0.7 mg).

Using high-resolution electrospray ionization mass spectrom-

etry (ESI-MS), 1H nuclear magnetic resonance (NMR), 13C NMR,

H-H correlated spectroscopy, H-H total correlation spectros-

copy, heteronuclear single quantum coherence (HSQC) spec-

troscopy, and heteronuclear multiple bond correlation (HMBC)

spectroscopy (Table S1 and Figure S3 available online), we

determined the chemical structure of lactazoles (Figure 3). The

major compound, lactazole A, showed m/z 1401.3975 [M+H]+,

indicating that the molecular formula is C61H64N18O16S3 (calcu-

lated 1,401.3788 [M+H]+). Abundant amide NH proton signals

(d 7.89–10.34) in 1H NMR spectra and carbonyl carbon signals

(d 158.85–173.91) in the 13C NMR spectra indicated a peptidic

structure. Four specific 1H signals (d 5.57–6.53) of two

methylenes and HMBC correlations to two relevant a carbons

(d 128.84 and 133.39) indicated two dehydroalanine residues.

Abundant 1H (d 8.34–8.71) and 13C (d 125.13–140.14) signals of

methines and HMBC correlations from those signals to relevant

quaternary carbons (d 139.72–173.24) indicated four azole rings.

10 12 14 16 18 20

S11C

W2S

pTYM1ga (control)

E

F

G

4 6 8 10 12 14Time (min)

11.0

14.7 A

B

C

D

ltc18-6C

lazA

KU465 (control)

lazC

11.3

13.1

Time (min)

Figure 2. HPLC Traces of Lactazoles and

the Expression of Related Compounds in

Heterologous Hosts

(A) S. lividans ltc18-6C containing the whole laz

cluster.

(B) S. lividans KU465 harboring pKU465 as a

control.

(C) S. lividans DlazA, the lazA disruptant.

(D) S. lividans DlazC, the lazC disruptant.

(E) S. lividans S11C produces the Ser11Cys

analog.

(F) S. lividans W2S produces the Trp2Ser analog.

(G) S. lividans pTYM1ga containing pTYM1ga as a

control.

HPLC conditions and sample preparation are

described in Experimental Procedures. The

elution peaks with arrows indicate the lactazole A

(A), a compound accumulated in DlazC (D), S11C

(E), and W2S (F), respectively.

Chemistry & Biology

A Minimum Gene Set for Thiopeptide Biosynthesis

Two methine signals (1H d 8.11 and 7.37, 13C d 118.18 and

138.35) and HMBC correlations to quaternary carbons (d

131.80–167.31) and carbonyl carbon (d 166.74) indicated trisub-

stituted pyridine. 1H signal (d 10.82) of NH and HMBC correla-

tions to carbons (d 110.17, 123.65, 127.37, and 136.09) in the

indole of tryptophan were detected. HMBC correlations also

indicated that tryptophan is attached to 3-position of pyridine.

The chemical structure of lactazoles exactly matched the

deduced amino acid sequence of the C-terminal region of the

putative precursor peptide. These spectrum analyses revealed

that lactazole A is an 11-amino acid macrocyclic thiopeptide

containing one oxazole, three thiazoles, two dehydroalanines,

and a 2,3,6-trisubstituted pyridine. Oxazole, tryptophan, and

thiazole attached at the pyridine 2-, 3-, and 6-positions, respec-

tively. Mass spectrometry (MS)/MS fragment patterns were

consistent with the structure elucidated from NMR analysis.

We also detected twominor analogs, lactazole B and C, which

hadsimilarUV-visible spectrumsonHPLC-photodiodearray (Fig-

ure S3). ESI-MS analysis showedm/z 1,529.4586 [M+H]+ for lac-

tazole B and m/z 1,176.2830 [M+H]+ for lactazole C, indicating

molecular formulas of C66H72N20O18S3 (calculated 1,529.4547

[M+H]+) and C51H49N15O13S3 (calculated 1,176.2875 [M+H]+),

respectively (Figure S3). These results indicate that lactazole B

has an additional glutamine residue (128.0 Da) at the C terminus

compared with lactazole A, whereas lactazole C lacks the C-ter-

minal glutamine and proline residues (225.2 Da) that are found in

lactazole A (Figure 3).MS/MSanalysis confirmed the existence of

an additional glutamine for lactazole B and the absence of gluta-

mine and proline in the C terminus of lactazole C (Figure S3). The

Chemical Abstracts Service database did not contain any com-

pounds related to lactazole A, B, and C.

Determination of the Lactazole Biosynthetic GeneCluster RegionTo identify the essential gene set required for lactazole biosyn-

thesis, we conducted subcloning and gene disruption experi-

ments (Figure S2). S. lividans ltc6C07, which harbors the

13.6-kb KpnI and TasI fragment that catalyzes lactazole pro-

duction, was used. Subsequently, ltc74 and ltc66, which are

located in the flanking regions at either end of this cluster,

were disrupted. Lactazoles were still produced by both Dltc74

Chemistry & Biology 21,

and Dltc67 S. lividans mutants, suggesting that these gene

products were not required for the biosynthesis. To confirm

that we had identified the essential gene set, we PCR-cloned

only lazA to lazF and expressed these six genes in S. lividans

TK23. The derivative strain, S. lividans lazAF, produced lacta-

zoles, and we thus conclude that these six genes are sufficient

for lactazole biosynthesis.

lazA encodes a 57 amino acid precursor peptide with no ho-

mology to other published sequences (Figure 4). LazB shares

similarity with GodF (identity 23.5%), which is known as a lanti-

biotic-type dehydratase. LazE is similar to GodD (identity

22.1%) and contains a C-terminal YcaO domain; the protein is

a putative cyclodehydratase required for azoline formation.

LazF is composed of an N-terminal lantibiotic dehydratase

domain and a C-terminal flavin mononucleotide-dependent de-

hydrogenase domain; the N-terminal domain shares homology

with GodG, while the C terminus is similar to that of GodE. This

observation may suggest that LazF functions as a bifunctional

enzyme that can catalyze both dehydroalanine formation and

azoline dehydrogenation. We could not assign any function to

LazC and LazD on the basis of a sequence homology search.

Because all posttranslational modifications except for macro-

cyclization could be assigned to LazB, LazE, and LazF, we

assumed that LazC and/or LazDmight be involved in the macro-

cyclization step (Table 1).

Although we conducted a series of genetic disruption experi-

ments to dissect the lactazole biosynthetic pathway, lactazoles

and related compounds could not be detected in the fermenta-

tion broth when lazA, lazB, lazD, lazE, or lazF was disrupted (Fig-

ure S2). By contrast, a newHPLC peak at 11.3 min was identified

in the extract of the fermentation broth of the lazC deletion strain

(S. lividansDlazC) (Figure 2). MS analysis of the peak showedm/z

1,437.4150 [M+H]+ (Figure S3), indicating a molecular formula of

C61H68N18O18S3 (calculated 1,437.4199 [M+H]+).

Structural Elucidation of the Products from the lazC

Deletion StrainThe product corresponding to m/z 1,437.4150 [M+H]+

(1,437.4194 calculated for C61H69N18O18S3) of lazC disruptant

was purified for structural elucidation. Mycelium from 2.4 l of

fermentation broth was extracted with methanol, followed by

679–688, May 22, 2014 ª2014 Elsevier Ltd All rights reserved 681

N

NO

S

N

NHO

O

NHO

NH

OHN

H2N O

O

O

HN

O

NH

O

HN

HO

N S

OHN

S N

HN

R

HN

N

O

OH2N

OHO

HN

N

O

OH2N

HNO

OH2N

OH

O

OH

O

HN

OH2N

O

NH O

HN N

OO

NH

SN

O

HN N

H

O N

O

ONH2

OHO

S

N

HN

O

S

N

NH

OH

O

HN

ONH

NH

O

O

N

NS

S

N

NHO

O

NHO

NH

OHN

H2N O

O

O

HN

O

NH

O

HN

HO

N S

OHN

S N

HN

HN

OH2N

ON

OHO

N

NO

S

N

NHO

O

NHO

NH

OHN

H2N O

O

O

HN

O

NH

O

HN

HO

N S

HO

OHN

S N

HN

OH2N

ON

OHO

A

B

C

Trp2

Gly3Ser4

Cys5*

Ser6*

Cys7*

Gln8

Ala9

Ser10*

Ser11*

Ser12*

Ser1*Cys13*

Ala14 R =

R =

R =

Lactazole A

Lactazole B

Lactazole C

Me-ketone1

Cys11*

Ser2

S11C

Gln15Pro16

Gln17

W2S

Figure 3. Chemical Structures of the Com-

pounds Isolated in This Study

(A) Lactazole A, B, and C.

(B) A compound that accumulates in the DlazC

mutant.

(C) Lactazole analogs, S11C and W2S.

Asterisks indicate the amino acid residues that are

posttranslationally modified.

Chemistry & Biology

A Minimum Gene Set for Thiopeptide Biosynthesis

lactazole purification, which yielded 2.5 mg of pure compounds.

NMR and MS analyses revealed that this compound was

fully modified, linear, and proteolyzed to 15 residues, as was

observed for lactazole A (Figure 3).

The product lacked the two proton signals derived from the

central pyridine (8.11 and 7.37 ppm in lactazole A; Table S1),

suggesting that this compound did not undergo macrocycliza-

tion and retained a linear structure. The loss of UV absorbance

around 325 nm supported this idea (Figure S3). Similar observa-

tions were reported following studies of thiocillin biosynthesis,

as the linear variants of thiocillin produced by a tclM deletion

strain lacked the absorbance around 350 nm (Bowers et al.,

2010).

Instead of a signal indicative of a central pyridine, three pairs

of signals corresponding to dehydroalanine residues were

observed at 6.24 and 5.75 ppm (Dha6), 5.90 and 5.85 ppm

(Dha10), and 6.18 and 5.75 ppm (Dha12) in the 1H NMR

spectrum. Correlations between HSQC and HMBC spectra

confirmed the presence of three dehydroalanines (Figure S3).

In addition to an additional dehydroalanine, a characteristic

methyl proton at 2.20 ppm was observed as a singlet. In the

HMBC spectrum, this methyl signal was correlation with two

carbonyl signals at 196.9 and 161.3 ppm. These observations

682 Chemistry & Biology 21, 679–688, May 22, 2014 ª2014 Elsevier Ltd All rights reserved

suggested the presence of an N-terminal

methyl ketone moiety. This is likely to

be derived from tautomerization and

hydrolysis of the N-terminal dehydroala-

nine residue, indicating that Ser1, Ser6,

Ser10, and Ser12 had already been

processed to dehydroalanines in the

product that accumulated in the lazC

disruptant.

Signals corresponding to three thia-

zoles and one oxazole could also be as-

signed in one-dimensional and 2D NMR

spectra. MS/MS fragmentation analysis

and the chemical shifts observed in 1H

NMR and 13C NMR are summarized in

Table S1.

Bioactivity of LactazolesLactazoles did not show any antimicro-

bial activities against the strains tested

in this study. In paper disc assays, lacta-

zoles did not inhibit the growth of

S. aureus FDA209P JC-1, Bacillus subtilis

ATCC6633, S. lividans TK23, Micro-

coccus luteus ATCC9341, Escherichia

coli NIHJ JC-2, Saccharomyces cerevi-

siae S100, Candida albicans A9540 and Penicillium chrysoge-

num NBRC6223.

However, during screening of compounds that could be used

for treatment of fibrodysplasia ossificans progressiva (Fukuda

et al., 2012), we found that the alkaline phosphatase activity of

osteoblastic C2C12 (R206H) cells was reduced by treatment

with lactazoles. The half maximal inhibitory concentration values

for lactazole A, B, and C were 101, 116, and 93.5 mM, respec-

tively, and no cytotoxicity was observed at doses up to 200 mg

ml�1. This result suggests that molecular targeting of lactazoles

could modulate signaling cascades triggered by bone morpho-

genetic protein.

Enhanced Lactazole A Production via GeneticEngineeringThe laz cluster is 9.8 kb, and all laz genes are ordered in the same

orientation without apparent internal gaps (Figure 1). This organi-

zation suggests that the lazA promoter drives expression of the

entire laz gene cluster. The yield of lactazole from this cluster

was 12.9 mg l�1, whereas the yield of goadsporin from the clus-

ter driven by the godA promoter was 342.7 mg l�1 (H.O., unpub-

lished data). Thus, we reasoned that lactazole production would

be increased by substitution of the lazA promoter with the godA

core peptide a No. of ring SCT ratio (aa) b (%) c

A T V S T I L C S G G T L S S A G C V 0 47 S W G S C S C Q A S S S C A - Q P - Q - D M 1 (11) 47 (56) S C N C V C G F C C S C S P - S A 1 (10) 63 (64) S C N C V C G V C C S C S P d 1 (10) 64 S T N C F C Y I C C S C S S - N 1 (10) 67 (71) S C N C F C Y I C C S C S S - A 1 (10) 67 (71) S C T T T G C A C S S S S S S T 1 (12) 88 S C T T S G C A C S S S S S S S S d 1 (12) 88 S C T C S G C A C S S S S S S S S d 1 (12) 88 S C T C S G C A C S S S S S S d 1 (12) 87 S C V G T A C A T S S S S S d 1 (9) 71 S C V G T A C A C S S T S S S S d 1 (12) 75 S C T T T S V S T S S S S S S 1 (12) 93 S C T T S S V S T S S S S S S S d 1 (12) 94 S C T T T G C T T S S T S S S S S d 1 (12) 94 S C T T T G C T T S S S S S S S S d 1 (12) 94 S C T T T G C T T S S S S S S S d 1 (12) 94 S N C T S T G T P A S C C S C C C C 2 (10, 12+s) 78 S C T T C V C I C S C C T d 1 (9) 85 S C T T C V C T C S C C T T 1 (9) 93 S C T S C V C T C S C C T T d 1 (9) 93

T S S S S C T T C I C T C S C S S d 2 (9, 7+q) 94 V S S A S C T T C I C T C S C S S 2 (9, 7+q) 82 I A S A S C T T C I C T C S C S S 2 (9, 7+q) 76 V A S A S C T T C I C T C S C S S d 2 (9, 7+q) 76

S C T T C E C C C S C S S 2 (9, 3+I) 92 S C T T C E C C C S C S d 2 (9, 3+I) 92 S C T T C E C S C S C S S 2 (9, 3+I) 92

goadsporin GodAlactazole LazA

GE2270A TpdA_GE amythiamicin

GE37468 GetAthiomuracin TpdA_TM

TP-1161 TpaA Thioplabin

thioxamycin thioactin

promothiocin radamycin

berninamycin BerA geninthiocin

methylsulfomycin sulfomycin

thiotipin cyclothiazomycin CltA

nocardithiocin thiocilin TclE-H

QN3323Y1 Sch40832 siomycin SioH

thiostrepton TsrHthiopeptin

nosiheptide NosMglycothiohexide nocathiacin NocM

0.1

Figure 4. Phylogenetic Analysis of the Core Peptide Amino Acid Sequences and Comparison with Various Thiopeptides and GoadsporinaAlignment of the core peptide amino acid sequences is shown. Color coding: red, serine; blue, cysteine; green, threonine. Amino acid residues that participate in

the formation of the main ring containing the central N-containing heterocyclic domain are underlined. The C-terminal cleavage sites in lactazole, GE37468,

thiomuracin, and GE2270A are indicated by hyphens. The core peptides in goadsporin (GodA) (Onaka et al., 2005) and in thiopeptide precursors, lactazole (LazA),

GE2270A (TpdA_GE) (Morris et al., 2009), GE37468 (GetA) (Young and Walsh, 2011), thiomuracins (TpdA_TM) (Morris et al., 2009), TP-1161 (TpaA) (Engelhardt

et al., 2010), berninamycin (BerA) (Malcolmson et al., 2013), cyclothiazomycin (CltA) (Wang et al., 2010), thiocillin (TclE-H) (Wieland Brown et al., 2009), siomycin

(SioH) (Liao et al., 2009), thiostrepton (TsrH) (Liao et al., 2009), nosiheptide (NosM) (Yu et al., 2009), and nocathiacin (NocM) (Ding et al., 2010), are shown.bThe number of macrocycles in thiopeptides is shown, with the number of amino acid residues that form the macrocycle given in parentheses (submacrocycles

bridged by sulfide, quinaldic acid, and indole are indicated by s, q, and i, respectively).cSer, Thr, and Cys (SCT) ratios in the core peptides are shown. The ratios except for C-terminal cleavage amino acids in lactazole A, GE37468, thiomuracin, and

GE2270A are given in parentheses.dPutative precursor sequences proposed from the chemical structures (Engelhardt et al., 2010).

Chemistry & Biology

A Minimum Gene Set for Thiopeptide Biosynthesis

promoter. As predicted, the amount of lactazole A in S. lividans

gp-lazAF, in which expression of the laz cluster is driven by the

godA promoter, reached 383.8 mg l�1 (Figure 5).

Production of Lactazole A Analogs by Amino AcidSubstitutionTo assess whether the posttranslational modification enzymes

could tolerate analogs, we performed a systematic site-directed

mutagenesis of the lazA precursor gene. For the analog produc-

tion, we constructed the pTYM1gp+lazA plasmid for precursor

peptide (lazA) overexpression and the pTYM19 gp+lazBF

plasmid for expression of the enzymes that mediate posttransla-

tional modifications (lazB to lazF operon). Both plasmids are co-

integrated into chromosomal DNA and are under control of the

godA promoter. We then performed two separate site-directed

mutagenesis reactions of lazA that led to the expression

of Ser11Cys and Trp2Ser analogs in the relevant S. lividans

transformant (Figure 2). The fermentation broth of S. lividans

Ser11Cys contained a specific product corresponding to m/z

Chemistry & Biology 21,

1417.3788 [M+H]+ that had a molecular mass of 16.0 Da more

than that of lactazole A, indicating the replacement of oxygen

with sulfur. The molecular weights and MS fragmentation anal-

ysis (Figure S3) indicated the production of the lactazole analog

Ser11Cys, which has a thiazole instead of an oxazole at the 2-

position of the central pyridine (Figure 3). Trp2Ser showed m/z

1302.3478 [M+H]+ (1302.3510 calculated for C53H60N17O17S3).

The result of MS fragmentation analysis (Figure S3) suggested

that the Trp2Ser analog can still be converted to a macrocyclic

ring, despite the significant change in amino acid structure

(Figure 3).

DISCUSSION

Here, we used a genome mining of the S. lactacystinaeus OM-

6519 sequence to identify a cryptic thiopeptide biosynthetic

gene cluster. Phylogenetic analysis of thiopeptide precursor

peptides revealed that the LazA core peptide did not belong

to an existing clade, suggesting that the lazA locus evolved

679–688, May 22, 2014 ª2014 Elsevier Ltd All rights reserved 683

Table 1. Deduced Genes and Their Proposed Functions in laz Cluster and the Surrounding Region

Gene

Amino

Acid Putative Function Pfam Prediction Homologous Gene (Origin) Identity (%)

Accession

Number

ltc74 857 – ND serine/threonine protein kinase

(S. griseus XylebKG-1)

50 ZP_08233747

lazA 57 precursor peptide – – –

lazB 880 Dha formation lantibiotic dehydratase,

N terminus

lantibiotic dehydratase,

C terminus

godF (Streptomyces sp. TP-A0584) 20 BAE46921

tsrJ (S. laurentii ATCC 31255) 25 ACN80672

lazC 406 [4+2] cycloaddition – tclM (B. cereus ATCC 14579) 12 AE016877

lazD 565 unknown – – – –

lazE 688 azoline formation YcaO-like family godD (Streptomyces sp. TP-A0584) 21 BAE46919

sioO (S. sioyaensis ATCC 13989) 28 ACN80652

lazF 557 Dha formation (11-288) SpaB C-terminal domain godG (Streptomyces sp. TP-A0584) 19 BAE46922

azole formation (362-547) nitroreductase family godE (Streptomyces sp. TP-A0584) 26 BAE46920

ltc66 117 – ND peptidase S24/S26A/S26B, conserved

region (Niastella koreensis GR20-10)

33 AEW02165

ltc65 275 – ND creatininase (Polymorphum gilvum

SL003B-26A1)

46 YP_004304855

ND, not determined.

Chemistry & Biology

A Minimum Gene Set for Thiopeptide Biosynthesis

independently from other known thiopeptide-encoding genes

(Figure 4). Indeed, in LazA, the proportion of posttranslationally

modified serine/threonine/cysteine residues is the lowest

(56%) among the known thiopeptide families. Furthermore, the

primary sequence of the LazA core peptide contains tryptophan

and glutamine residues, which to date have never been found in

natural thiopeptides. The unique genomic context and associ-

ated peptide sequence drove us to characterize the metabolites

derived from this cryptic biosynthetic gene cluster. Strikingly,

this cluster only produces a unique class of thiopeptides, the

lactazoles, when expressed in an alternate host bacterial strain,

S. lividans TK23.

Unexpectedly, lactazoles were composed of a macrocyclic

ring of 11 amino acids. This is unusual, because all other macro-

cycles found in thiopeptides are 9, 10, or 12 amino acids in

length. Thus, the lactazole ring addsmore diversity to themacro-

cycle framework. In addition, the 2-oxazolyl-6-thiazolyl pyridine

core, in which the 3-position is connected to the bulky trypto-

phan side chain by an amide linkage, is also a structure unique

to lactazole. Typically, 2,3,6-trithiazolylpyridine core is known

in thiopeptides. Although some atypical substitution patterns,

such as 2-oxazolyl-3-thiazolylpyridine in berninamycin and TP-

1161 and 6-thiazolylpyridine in cyclothiazomycin, have also

been isolated, 2-oxazolyl-6-thiazolyl pyridine moiety was clearly

different from known examples. Additionally, we found that the

length of the C terminus varies between the lactazole B andC an-

alogs. This C-terminal chain length diversity is also not observed

in other thiopeptides, and we suggest that it may be due to pep-

tidases present in the heterologous host.

Most thiopeptides have growth-inhibitory activity against

Gram-positive bacteria (Bagley et al., 2005). However, lactazoles

did not exert such effects against any of the microorganisms we

tested. We found that the alkaline phosphatase activity of oste-

oblastic C2C12 (R206H) cells was reduced following lactazole

treatment, suggesting that the molecular target(s) of lactazoles

684 Chemistry & Biology 21, 679–688, May 22, 2014 ª2014 Elsevier

may be involved in bone morphogenetic protein signaling cas-

cades in vivo. Identification of the precise targets will shed light

on the potential antibiotic utility of this novel thiopeptide family.

The compact genomic organization of the lactazole biosyn-

thetic cluster makes it highly amenable to genetic manipulation;

indeed, the whole cluster can be amplified by a single round of

PCR. This conciseness can be explained by the lack of apparent

intergenic regions. Among the six genes, lazB, lazC, lazD, and

lazE are translationally coupled, while the lengths of the inter-

genic regions between lazA and lazB and between lazE and

lazF are 221 and 80 bp, respectively. We aimed to improve lac-

tazole production by replacing the single promoter upstream

of lazA with the godA promoter, because we have previously

shown that this promoter drives high levels of goadsporin pro-

duction (342.7 mg l�1) (H.O., unpublished data) in S. lividans

TK23. Indeed, the strain with the godA promoter substitution

produced 383.8 mg 1�1 of lactazole A upon culture, which is a

29.8-fold increase compared with the original strain.

Because promoter substitution gave a high yield of production

in S. lividans TK23, we further examined whether analogs could

be produced using this system. To evaluate the substrate spec-

ificities of tailoring enzymes, we focused on the substitution of

neighboring residues of the central pyridine core, which led to

the production of the two analogs, S11C andW2S. The chemical

structure of the S11C substitution was transformed from oxazole

to thiazole; however, the W2S substitution was not converted to

oxazole. Together, these data suggest that the enzyme required

for azole ring formation is regiospecific.

The macrocyclization step during thiopeptide biosynthesis

is challenging to reproduce biochemically, and the associated

mechanistic steps remain. Our gene disruption experiments re-

vealed that lazC is responsible for the pyridine ring formation,

because the product of lazC disruptant is linear, and the N-termi-

nal serine residue was replacedwith amethyl ketone. The forma-

tion of the methyl ketone would likely occur through a primary

Ltd All rights reserved

S. lividans ltc6C07 S. lividans gp-lazAF

lact

azol

e A

(mg/

l)

GS

A-3M

500

400

300

200

100

0

Figure 5. Lactazole Production by Promoter Substitution of laz

Cluster in GS and A-3M Media

S. lividans Ltc6C07: laz cluster expressed under original promoter. S. lividans

gp-lazAF: laz cluster expressed under godA promoter. Seven day culture broth

was measured. Each value is the mean ± SD of triplicate values from a

representative experiment.

Chemistry & Biology

A Minimum Gene Set for Thiopeptide Biosynthesis

enamine form of the first dehydrated amino acid (Dha1) through

tautomerization and subsequent spontaneous hydrolysis.

Similar linear peptides have been detected following gene

disruption within other biosynthetic clusters, such as those that

produce thiocillin, GE2270. The onlymechanistic insight into pyr-

idine ring formation in thiopeptide biosynthesis was provided

following disruption of tclM in the thiocillin biosynthetic gene

cluster (Bowers et al., 2010). TclM, a putative Diels-Alderase, is

conserved in the gene cluster for macrocyclic thiopeptide and

does not exist in the acyclic goadsporin gene cluster. Disruption

of tclM resulted in the accumulation of a linear compound con-

taining an N-terminal methyl ketone moiety, suggesting that

TclM catalyzes [4+2] cycloaddition between the two dehydroala-

nines. Despite very low similarity between lazC and tclM (Fig-

ure S1), disruption of lazC also led to the accumulation of a linear

compound that has properties similar that isolated from the tclM

mutant. Similar structures, namely, linear polypeptides with a

methyl ketone moiety at the N terminus, were observed in

biosynthetic intermediate derivatives of GE2270 from Planobis-

pora rosea (Tocchetti et al., 2013). In this latter case, the authors

concluded that insufficient maturation of the C terminus pre-

vented completion of the macrocyclization step, which led to

the accumulation of linear compounds. However, our current

study revealed that the products that accumulate in the lazC

deletion strains were mature but simply had not undergone

macrocyclization. On the basis of these results, we conclude

that LazC catalyzes the Diels-Alder condensation at the last

step of posttranslational modification in a manner similar to

TclM. Linear compounds with highly matured structures were

also isolated in berninamycin biosynthesis by precursor peptides

mutations. This also provided insight into late-stage macrocycli-

zation event (Malcolmson et al., 2013).

Consistent with earlier studies of thiopeptide biosynthesis

(Ding et al., 2010; Liao et al., 2009), disruption of laz genes abol-

ishes lactazole production and prevents product accumulation.

Specifically, our disruption studies confirmed the involvement

of lazB, lazD, lazE, and lazF in the lactazole biosynthesis. Bio-

informatic analyses predicted that LazB, LazE, and LazF prod-

ucts were responsible for the formation of dehydroalanine and

thiazole/oxazole. On the other hand, the lack of sequence simi-

Chemistry & Biology 21,

larity of LazDwith any of the known enzymes involved in thiopep-

tide biosyntheses precluded the assignment of a function for

this protein (Table 1). It is possible that this enzyme cleaves the

leader peptide of the matured precursor; in this scenario, the

resulting linear core peptide would be a suitable substrate for a

macrocyclization reaction possibly catalyzed by LazC.

On the basis of the chemical structures of lactazoles, the gene

annotation, and our gene disruption experiments, we propose

the lactazole biosynthetic pathway shown in Figure 6. In the first

step, the lazA gene is transcribed and then translated into the

precursor peptide by the ribosome. Although the order of dehy-

droalanine and azole ring formation is still unknown, Ser1, Ser6,

Ser10, and Ser12 in the precursor peptide will be modified

into dehydroalanines by LazB, a lantibiotic-type dehydratase,

whereas Cys5, Cys7, and Cys13 will be converted to thiazoles,

and Ser11 will be converted into an oxazole by the activities of

LazE (a cyclodehydratase) and LazF (a dehydrogenase). It is

likely that N-terminal leader peptide cleavage then occurs,

followed by a hetero-Diels-Alder condensation between Dha1

and Dha12 catalyzed by LazC. However, this model is still

incomplete, as the true function of LazD remains obscure.

The lactazoles we isolated in this study contain unusual

structural features. Even though more than 100 thiopeptides

have been discovered since 1948, only three types of macrocy-

clic rings, with 26, 29, or 35 atoms, have been isolated. Here

we found a 32-membered macrocyclic ring formed by 11 amino

acid residues that bridged the missing link of macrocyclic ring

size. Additionally, the 2-oxazolyl-6-thiazolyl pyridine core, with

the 3-position connected to tryptophan through an amide link-

age, provides a core structure that is different from those of

known thiopeptides. Our report demonstrates that genome

mining is a powerful strategy that can be applied to uncover

biosynthetic pathways for natural drug products. The ability

to engineer these compact biosynthetic clusters to produce

diverse groups of thiopeptide analogs with distinct bioactivities

bodes well for the future of natural product-based drug

discovery.

SIGNIFICANCE

Thiopeptides are one of themajor groups among the natural

antimicrobial products that are biosynthesized by ribo-

somal translational machinery. In this study, we found the

simplest thiopeptide biosynthetic gene cluster from the

S. lactacystinaeus OM-6519 genome. Although the biosyn-

thetic gene cluster was silent in the original strain, transfer

of the cluster toS. lividans TK23 led to the synthesis of struc-

turally unusual thiopeptides, lactazoles. Currently, the ring

size of macrocyclic thiopeptide is limited to 26, 29, or 35

atoms. A 32-membered ring of lactazoles expands the struc-

tural possibility of thiopeptides. The 2-oxazolyl-6-thiazolyl

pyridine core with the 3-position connected to tryptophan

through an amide linkage also provides a unique core

structure in thiopeptides. The formation of this unique 32-

memberedmacrocycle would be catalyzed by LazC. Consis-

tent with its structural uniqueness, the precursor peptide

sequence was classified into a phylogenetically distinct

clade. Unexpectedly, we found that the lactazoles do

not have any antimicrobial activity; instead, they possess

679–688, May 22, 2014 ª2014 Elsevier Ltd All rights reserved 685

NH

HN N

H

HN

O

O

ONH

HN

O

O

O

NH

HN N

H

HN

O

O

ONH

HN

O

ONH

HN

O

ONH

O

O

N

NHO

H2N O

OH2N

NH

HO HS

HO HO

HS

HO

HO

HO

HS

HN N

H

OH2N

O O

HO

O

S

O

OH

NH

HN

NH

HN

O

O

ON

HN

ON

HN

NH

HN

O

O

ON

HN

ON

HN

ONH

O

O

N

O

H2N O

OH2N

NH HO

S S SO

NH

HN N

H

OH2N

O O

HO

O

S

O

OH

NN

S HN

O

NH

O

OH2N

N

O OOH

ON

HNO

NH

OHN

HN

O

O

HN

HN

OH2N

O

NS

NHO

OHN

NS OH

LazA

LP

LP

Dehydration: LazB and LazF

Cyclization: LazEDehydrogenation: LazF

LP cleavage: LazD?Macrocyclization: LazCC-terminal proteolysis: unknown

lazctazole A

Figure 6. Proposed Lactazole Biosynthetic

Pathway

Color coding indicates the biosynthesis steps

of azole ring formation (green), dehydroalanine

production (blue), and pyridine formation (red). LP,

leader peptide.

Chemistry & Biology

A Minimum Gene Set for Thiopeptide Biosynthesis

inhibitory activities for the bone morphogenetic protein

signal cascade in vivo, which could direct thiopeptide family

antibiotics to a new aspect of therapeutic treatment.

The lactazole biosynthetic gene cluster consists of only

six unidirectional genes. It is the smallest cluster among

the known thiopeptide biosynthetic gene clusters. The sub-

stitution of the endogenous promoter with that of godA,

a gene involved in goadsporin biosynthesis, led to a 29.8-

fold increase in the production of lactazole A. Using the

godA promoter to regulate the lactazole biosynthetic

machinery, production of two analogs, S11C and W2S, was

achieved. Thus, this compact biosynthetic machinery has

high potency to lend large diversity to the thiopeptide core

structures.

EXPERIMENTAL PROCEDURES

GenomeMining and Cloning of Lactazole Biosynthetic Gene Cluster

Draft genome sequencing of S. lactacystinaeus OM-6519 was performed

using an Illumina GA-II instrument (Illumina). De novo assembly of sequence

data was performed using Velvet software version 0.7.18. We identified

thiopeptide biosynthetic gene cluster from genome of S. lactacystinaeus

OM-6519, which was found by local BLAST search using godF, godD, and

godG as the queries.

ORFs were deduced from the sequence with the assistance of FramePlot

4.0 beta software (Ishikawa and Hotta, 1999). The corresponding deduced

proteins were compared with other known proteins in the databases using

available BLAST methods (http://www.ncbi.nlm.nih.gov/blast). Amino acid

sequence alignments were performed with CLUSTALW from the DNA Data

Bank of Japan (DDBJ) (http://www.ddbj.nig.ac.jp/index-j.html).

For S. lactacystinaeus OM-6519 cosmid library, the genomic DNA was

digested with Sau3AI and ligated into the BamHI site of pKU465cos (Komatsu

et al., 2010). Packaging and transformation into E. coli DH5a were done

with l-packaging mixtures. pKU465-ltc18-6C including laz gene cluster

was acquired by PCR screening using two sets of primers, LTC11_73_F

and LTC11_73_R, and LTC11_67 godE&G_F and LTC11_67 godE&G_R.

pKU465-ltc18-6C was transformed into S. lividans TK23 to generate

S. lividans ltc18-6C.

686 Chemistry & Biology 21, 679–688, May 22, 2014 ª2014 Elsevier Ltd All rights reserved

Sample Preparation and HPLCDetection of

Lactazoles

Each strain was inoculated into test tubes contain-

ing 10ml of V-22medium. After incubation at 30�Cfor 2 days on a rotary shaker at 200 rpm, 3 ml cul-

tures of the seed culture were transferred into

500 ml K-1 flasks (Preci) containing 100 ml of A-

3M or goadsporin (GS) medium. Fermentation

was carried out at 30�C, 200 rpm, for more than

5 days on the rotary shaker. The 4 ml whole-

fermentation broths were extracted with 5 ml

n-butanol. After evaporation of the n-butanol, the

residue was dissolved in DMSO. HPLC analysis

(Figures 2A–2D) was performed with an HP1100

(Hewlett-Packard) system using a C18 Rainin Mi-

crosorb column (3 mm, 4.6 mm inside diameter 3

100 mm length; Rainin Instrument). MeCN,

0.15% KH2PO4 (pH 3.5), was used as elution

buffer. The column temperature was 40�C, the flow rate was 1.2 ml min�1,

and detection was performed at 254 nm. The MeCN concentration was kept

at 15% for the first 3 min, linearly increased to 85% for the next 22 min, and

kept at 85% for the next 4min. Lactazoles were identified on the basis of reten-

tion times, UV spectra, andmolecular mass. Chromatograms in Figures 2E–2G

were obtained using an ion-trap liquid chromatography (LC)-MS system

equipped with a Cosmosil 5C18 AR-II column (5 mm, 2.0 mm inside

diameter 3 150 mm length; Nacalai Tesque). MeCN and 0.1% HCOOH were

used as eluents. The column temperature was 40�C, the flow rate was

0.3 ml min�1, and detection was performed at 254 nm. The MeCN concentra-

tion was kept at 5% for the first 2 min, linearly increased to 95% over the next

25 min, and kept at 95% for the next 5 min.

Purification of Lactazoles and a DlazC Disruptant Product

After fermentation, the fermentation broth (3 l) was centrifuged at 8,000 rpm for

10 min. The supernatant was discarded, and the mycelial cake was extracted

with 3 l of methanol three times. The mixture was centrifuged at 8,000 rpm

for 10 min, and the methanol layer was separated from the mycelial cake.

After evaporation to remove methanol, the resultant aqueous solution was

partitioned between 60% aqueous methanol and CH2Cl2. The 60% aqueous

methanol-soluble fraction was evaporated and partitioned between H2O and

n-butanol. Evaporation of the n-butanol-soluble fraction gave approximately

1.6 g of extract. A fraction recovered by solving in DMSO was subjected to

Sephadex LH-20 gel filtration chromatography using methanol-CH2Cl2 (1:1)

as eluent, and aliquots of each fraction were analyzed by HPLC described

above. After the fractions containing lactazoles were combined and evapo-

rated (�200 mg), further purified by reverse-phase octadecylsilyl (ODS)

(75 mm, Cosmosil 75C18-PREP; Nacalai Tesque) column chromatography

with a step gradient of acetonitrile/H2O (2:8, 3:7, 4:6, 5:5, 6:4, 7:3, and

8:2 v/v). The fractions containing lactazoles were combined (�49 mg) and

then purified by reverse-phase HPLC (1200 series; Agilent Technologies) sys-

tem using Cosmosil 5PE-MS (5 mm, 10mm inside diameter3 250mm [Nacalai

Tesque], 4 ml min�1, detection at 254 nm) with a gradient elution from 25% to

50% of 0.1% HCOOH buffer with MeCN over 30 min to give lactazole A

(6.7 mg), B (2.7 mg), and C (0.7 mg). For the purification of a DlazC product,

the mycelial cake from 2.4 l fermentation broth was extracted by methanol

as described above. The methanol fraction was washed by hexane and recov-

ered. The recovered fraction was partitioned between 60% aqueous methanol

and CH2Cl2. The 60% aqueousmethanol-soluble fraction was evaporated and

partitioned between H2O and n-butanol. Evaporation of n-butanol-soluble

Chemistry & Biology

A Minimum Gene Set for Thiopeptide Biosynthesis

fraction gave approximately 4.2 g of extract. A fraction recovered by solving in

DMSOwas separated by Sephadex LH-20 gel filtration chromatography using

the method described above. The fractions containing lactazoles were com-

bined and evaporated (�240 mg), further purified by reverse-phase ODS

(75 mm, Cosmosil 75C18-PREP) column chromatography with a step gradient

of acetonitrile/H2O (1:9, 2:8, 3:7, 4:6, and 10:0 v/v). The fractions containing

lactazoles were combined (�48 mg) and then purified by reverse-phase

HPLC (L-2000 system; Hitachi) using Xterra Prep RP18 (5 mm, 10 mm inside

diameter3 150mm [Waters], 3mlmin�1, detection at 254 nm) with an isocratic

elution by 73% of 10 mM NH4HCO3 buffer with MeCN over 30 min to give a

purified compound.

Structure Elucidation of Lactazoles

NMR spectra were obtained with a Bruker AVANCE 500 spectrometer at 500

MHz for 1H and 125 MHz for 13C using residual solvent (DMSO-d6) peaks at

dH 2.49 and at dC 39.8 ppm as chemical shift reference signals. Exact MS

was performed on a time-of-flight instrument (Xevo G2-S Tof; Waters) inter-

faced to an ultraperformance liquid chromatograph (ACQUITY; Waters) oper-

ating in MSE mode with positive electrospray ionization mode. NMR spectra

of a DlazC product were obtained with an ECA-600 (JEOL) spectrometer

at 600 MHz for 1H and 150 MHz for 13C NMR. MS/MS and MS/MS/MS

analyses were performed on an ion-trap instrument (amaZon SL; Bruker Dal-

tonics) interfaced to an HPLC system (Agilent Technologies 1200 series)

operating in MS/MS and MS/MS/MS mode with positive electrospray ioniza-

tion mode.

Subcloning of laz Gene Cluster

(1) pTYM19-ltc6C07: pKU465-ltc18-6C was digested with BamHI, and the re-

sulting 11.9-kb fragment was cloned into the BamHI site of pTYM19, which is

an actinomycetes-E. coli integrating vector (Onaka et al., 2003b), to generate

pTYM19-ltc6C01. pKU465-ltc18-6C was digested with KpnI and PstI, and the

resulting 6.8-kb fragment was cloned into the KpnI and PstI site of pTYM19 to

generate pTYM19-ltc6C02. pTYM19-ltc6C01 was digested with PstI, and the

resulting 9.9-kb fragment was subcloned into the PstI site of pTYM19-ltc6C02

to generate pTYM19-ltc6C03. pTYM19-ltc6C03, which has a 16.8-kb region

including laz cluster from pKU465-ltc18-6C, was digested with EcoRI and

TasI, and the resulting 13.6-kb fragment was subcloned into the EcoRI site

of pTYM19 to generate pTYM19-ltc6C07. (2) pTYM19-lazAF: The fragment

from upstream of lazA to downstream of lazF was amplified as a 10.7-kb

DNA fragment by PCR, using pTYM19-ltc6C03 as the template and two

primers, ltc73+pro-F-EcoRI and ltc68-down-3-EcoRI-2. The PCR-generated

fragment containing the laz cluster was cloned into the EcoRI sites of

pTYM19 to generate pTYM19-lazAF. The new junction regions in the con-

structed plasmids were confirmed by direct sequencing.

Construction of the Strains for Gene Inactivation of lazA, lazB, lazC,

lazD, lazE, lazF, ltc74, and ltc66

For the in-frame deletion of lazA, lazB, lazC, lazD, lazE, lazF, ltc74, and ltc66,

the method of insertional inactivation by a double crossover was performed.

The 2.0-kb upstream and downstream regions of those genes were PCR

amplified. The resulting fragments were digested with EcoRI and XbaI and

ligated into the EcoRI sites of pTYMdis (Onaka et al., 2005) to give

pTYMdis-lazA, pTYMdis-lazB, pTYMdis-lazC, pTYMdis-lazD, pTYMdis-

lazE, pTYMdis-lazF, pTYMdis-ltc74, and pTYMdis-ltc66. pTYM19Xint was

constructed from pTYM19 (Onaka et al., 2003b) by the deletion of

4C31int*. pTYM19 was digested with NcoI and XhoI, and the resulting larger

fragment was converted to a blunt-ended molecule and self-ligated to

generate pTYM19Xint. pTYMdis-lazA, pTYMdis-lazD, pTYMdis-lazE,

pTYMdis-lazF, and pTYMdis-ltc74 were digested with EcoRI, and the result-

ing 4.0-kb fragments were subcloned into the EcoRI site of pTYM19Xint to

generate pTYM19X-lazA, pTYM19X-lazD, pTYM19X-lazE, pTYM19X-lazF,

pTYM19X-ltc74. The generated plasmids were confirmed by digestion with

appropriate restriction enzyme(s) and DNA sequencing using For and

RV primers. The gene disruption procedure was as described previously

(Onaka et al., 2003a). In-frame disruptants, S. lividans ltc18-6CD74, ltc18-

6CDlazA, ltc18-6CDlazB, ltc18-6CDlazC, ltc18-6CDlazD, ltc18-6CDlazE,

ltc18-6CDlazF, and ltc18-6CD66, were generated using pTYMdis derivatives

or pTYM19Xint derivatives.

Chemistry & Biology 21,

lazAwas PCR amplified using LTC73-F-NdeI and LTC73-R-XbaI. The result-

ing 0.4-kb fragment was digested with NdeI and XbaI and cloned into the NdeI

and XbaI sites, downstream of godA promoter, of pTYM1ga to generate

pTYM1ga-lazA.

Bioactivity of Lactazoles

Bioactivity of lactazoles against the eight indicator strains, S. aureus FDA209P

JC-1, B. subtilis ATCC6633, S. lividans, M. luteus ATCC9341, E. coli NIHJ

JC-2, S. cerevisiae S100, C. albicans A9540, and P. chrysogenum at

0.04 mg/5 mm disc on Luria broth glucose agar medium (1% tryptone, 0.5%

yeast extract, 1% NaCl, 0.5% glucose, 1.5% agar) at 30�C for 3 days.

Effect of lactazoles for alkaline phosphatase activity in bone morphogenetic

protein-treated C2C12 (R206H) cell was assayed to determine the inhibition of

osteoblastic differentiation as described previously (Fukuda et al., 2012).

Increase of Lactazole A Production by Promoter Substitution

The fragments of the lactazole biosynthetic gene cluster and pTYM3ga for a

vector were amplified using four sets of primers as follows: TYM3ga+

lazA(Nde) and lazD-up-primer, lazD-down-primer and lazF+TYM3ga(Bam)-r,

lazF+TYM3ga(Bam) and TYM3ga_div_at_apr-oriT_forGA-r, TYM3ga_div_at_

apr-oriT_forGA and TYM3ga+lazA(Nde)-r. The generated four fragments

were assembled by Gibson assembly (Gibson et al., 2009) to generate

pTYM3ga-lazAF. pTYM3ga is an actinomycetes-E. coli integrating vector

containing 4K38-1 integrase, apramycin-resistant gene acc(3)IV, and godA

promoter. S. lividans TK23 was transformed with pTYM3ga-lazAF to

generate the laz cluster-promoter substituted strain, S. lividans gp-lazAF.

The strain was incubated in GS medium for 5 to 10 days, and lactazole A

was detected by HPLC (Agilent Technologies 1200 series, using Cosmosil

5C18-AR-II, 5 mm, 4.6 mm inside diameter 3 250 mm) after extraction as

described above. The amount of lactazole A was quantified from the intensity

of the peak.

Analog Production of Lactazole A by Amino Acid Substitution

For lazA mutagenesis, site-directed mutagenesis was carried out using the

QuikChange site-direct mutagenesis kit (Stratagene) with pTYM1ga-lazA

using QC-lazA-S11C-sense, QC-lazA-S11C-anti, QC-lazA-W2S-sense, and

QC-lazA-W2S-anti to generate pTYM1ga-lazAS11C and pTYM1ga-lazAW2S,

respectively. pTYM19 gt-lazBF in which lazB-F was expressed by godA pro-

moter was constructed and transferred in S. lividans TK23 to generate

S. lividans lazBF. To construct pTYM19 gt-lazBF, pTYM3ga-lazBF was first

constructed by Gibson assembly from four PCR fragments using four sets of

primers as follows: TYM3ga+lazB(Nde) and lazD-up-primer, lazD-down-

primer and lazF+TYM3ga(Bam)-r, lazF+TYM3ga(Bam) and TYM3ga_div_at_

apr-oriT_forGA-r, TYM3ga_div_at_apr-oriT_forGA and TYM3ga+lazB(Nde)-r.

Subsequently, pTYM1ga-lazAS11C and pTYM1ga-lazAW2S were transferred

in S. lividans lazBF to generate S. lividans S11C and S. lividans W2S,

respectively.

ACCESSION NUMBERS

The sequence of laz cluster has been deposited in the DDBJ database (acces-

sion number AB820694).

SUPPLEMENTAL INFORMATION

Supplemental Information includes Supplemental Experimental Procedures,

three figures, and three tables and can be found with this article online at

http://dx.doi.org/10.1016/j.chembiol.2014.03.008.

ACKNOWLEDGMENTS

This work was supported in part by grants-in-aid from the Institution of

Fermentation, Osaka (to S.H., S.A., T.O., and H.O.) and JSPS KAKENHI grant

No. 25108707 (to H.O.). We thank Dr. Ryuji Uchida, Dr. Kiyoko T. Miyamoto at

Kitasato University for experimental assistance with the biological assay and

LC/MS analysis, Dr. YasuoOhnishi, and Dr. Yohei Katsuyama at The University

of Tokyo for helpful suggestions.

679–688, May 22, 2014 ª2014 Elsevier Ltd All rights reserved 687

Chemistry & Biology

A Minimum Gene Set for Thiopeptide Biosynthesis

Received: February 24, 2014

Revised: March 15, 2014

Accepted: March 21, 2014

Published: April 24, 2014

REFERENCES

Arnison, P.G., Bibb, M.J., Bierbaum, G., Bowers, A.A., Bugni, T.S., Bulaj, G.,

Camarero, J.A., Campopiano, D.J., Challis, G.L., Clardy, J., et al. (2013).

Ribosomally synthesized and post-translationally modified peptide natural

products: overview and recommendations for a universal nomenclature. Nat.

Prod. Rep. 30, 108–160.

Bagley, M.C., Dale, J.W., Merritt, E.A., and Xiong, X. (2005). Thiopeptide anti-

biotics. Chem. Rev. 105, 685–714.

Bowers, A.A., Acker, M.G., Koglin, A., and Walsh, C.T. (2010). Manipulation of

thiocillin variants by prepeptide gene replacement: structure, conformation,

and activity of heterocycle substitution mutants. J. Am. Chem. Soc. 132,

7519–7527.

Bycroft, B.W., and Gowland, M.S. (1978). The structures of the highly modi-

fied peptide antibiotics micrococcin P1 and P2. J. Chem. Soc. Chem.

Commun. 256.

Ding, Y., Yu, Y., Pan, H., Guo, H., Li, Y., and Liu, W. (2010). Moving posttrans-

lational modifications forward to biosynthesize the glycosylated thiopeptide

nocathiacin I in Nocardia sp. ATCC202099. Mol. Biosyst. 6, 1180–1185.

Engelhardt, K., Degnes, K.F., and Zotchev, S.B. (2010). Isolation and charac-

terization of the gene cluster for biosynthesis of the thiopeptide antibiotic TP-

1161. Appl. Environ. Microbiol. 76, 7093–7101.

Fukuda, T., Uchida, R., Ohte, S., Inoue, H., Yamazaki, H., Matsuda, D.,

Nonaka, K., Masuma, R., Katagiri, T., and Tomoda, H. (2012). Trichocyalides

A and B, new inhibitors of alkaline phosphatase activity in bone morphoge-

netic protein-stimulated myoblasts, produced by Trichoderma sp. FKI-5513.

J. Antibiot. 65, 565–569.

Gibson, D.G., Young, L., Chuang, R.Y., Venter, J.C., Hutchison, C.A., 3rd, and

Smith, H.O. (2009). Enzymatic assembly of DNA molecules up to several hun-

dred kilobases. Nat. Methods 6, 343–345.

Ishikawa, J., and Hotta, K. (1999). FramePlot: a new implementation of the

frame analysis for predicting protein-coding regions in bacterial DNA with a

high G + C content. FEMS Microbiol. Lett. 174, 251–253.

Komatsu, M., Uchiyama, T., Omura, S., Cane, D.E., and Ikeda, H. (2010).

Genome-minimized Streptomyces host for the heterologous expression of

secondary metabolism. Proc. Natl. Acad. Sci. U S A 107, 2646–2651.

Li, C., and Kelly,W.L. (2010). Recent advances in thiopeptide antibiotic biosyn-

thesis. Nat. Prod. Rep. 27, 153–164.

Li, J., Qu, X., He, X., Duan, L., Wu, G., Bi, D., Deng, Z., Liu, W., and Ou, H.Y.

(2012). ThioFinder: a web-based tool for the identification of thiopeptide

gene clusters in DNA sequences. PLoS ONE 7, e45878.

Liao, R., Duan, L., Lei, C., Pan, H., Ding, Y., Zhang, Q., Chen, D., Shen, B., Yu,

Y., and Liu, W. (2009). Thiopeptide biosynthesis featuring ribosomally synthe-

sized precursor peptides and conserved posttranslational modifications.

Chem. Biol. 16, 141–147.

688 Chemistry & Biology 21, 679–688, May 22, 2014 ª2014 Elsevier

Malcolmson, S.J., Young, T.S., Ruby, J.G., Skewes-Cox, P., and Walsh, C.T.

(2013). The posttranslational modification cascade to the thiopeptide bernina-

mycin generates linear forms and altered macrocyclic scaffolds. Proc. Natl.

Acad. Sci. U S A 110, 8483–8488.

Morris, R.P., Leeds, J.A., Naegeli, H.U., Oberer, L., Memmert, K., Weber, E.,

LaMarche, M.J., Parker, C.N., Burrer, N., Esterow, S., et al. (2009).

Ribosomally synthesized thiopeptide antibiotics targeting elongation factor

Tu. J. Am. Chem. Soc. 131, 5946–5955.

Omura, S., Fujimoto, T., Otoguro, K., Matsuzaki, K., Moriguchi, R., Tanaka, H.,

and Sasaki, Y. (1991). Lactacystin, a novel microbial metabolite, induces neu-

ritogenesis of neuroblastoma cells. J. Antibiot. 44, 113–116.

Onaka, H., Tabata, H., Igarashi, Y., Sato, Y., and Furumai, T. (2001).

Goadsporin, a chemical substance which promotes secondary metabolism

and morphogenesis in streptomycetes. I. Purification and characterization.

J. Antibiot. 54, 1036–1044.

Onaka, H., Taniguchi, S., Igarashi, Y., and Furumai, T. (2003a).

Characterization of the biosynthetic gene cluster of rebeccamycin from

Lechevalieria aerocolonigenes ATCC 39243. Biosci. Biotechnol. Biochem.

67, 127–138.

Onaka, H., Taniguchi, S., Ikeda, H., Igarashi, Y., and Furumai, T. (2003b).

pTOYAMAcos, pTYM18, and pTYM19, actinomycete-Escherichia coli inte-

grating vectors for heterologous gene expression. J. Antibiot. 56, 950–956.

Onaka, H., Nakaho, M., Hayashi, K., Igarashi, Y., and Furumai, T. (2005).

Cloning and characterization of the goadsporin biosynthetic gene cluster

from Streptomyces sp. TP-A0584. Microbiology 151, 3923–3933.

Tocchetti, A., Maffioli, S., Iorio, M., Alt, S., Mazzei, E., Brunati, C., Sosio, M.,

and Donadio, S. (2013). Capturing linear intermediates and C-terminal variants

during maturation of the thiopeptide GE2270. Chem. Biol. 20, 1067–1077.

Walsh, C.T., and Nolan, E.M. (2008). Morphing peptide backbones into hetero-

cycles. Proc. Natl. Acad. Sci. U S A 105, 5655–5656.

Walsh, C.T., Acker, M.G., and Bowers, A.A. (2010). Thiazolyl peptide antibiotic

biosynthesis: a cascade of post-translational modifications on ribosomal

nascent proteins. J. Biol. Chem. 285, 27525–27531.

Wang, J., Yu, Y., Tang, K., Liu, W., He, X., Huang, X., and Deng, Z. (2010).

Identification and analysis of the biosynthetic gene cluster encoding the

thiopeptide antibiotic cyclothiazomycin in Streptomyces hygroscopicus

10-22. Appl. Environ. Microbiol. 76, 2335–2344.

Wieland Brown, L.C., Acker, M.G., Clardy, J., Walsh, C.T., and Fischbach,

M.A. (2009). Thirteen posttranslational modifications convert a 14-residue

peptide into the antibiotic thiocillin. Proc. Natl. Acad. Sci. U S A 106, 2549–

2553.

Young, T.S., and Walsh, C.T. (2011). Identification of the thiazolyl peptide

GE37468 gene cluster from Streptomyces ATCC 55365 and heterologous

expression in Streptomyces lividans. Proc. Natl. Acad. Sci. U S A 108,

13053–13058.

Yu, Y., Duan, L., Zhang, Q., Liao, R., Ding, Y., Pan, H., Wendt-Pienkowski, E.,

Tang, G., Shen, B., and Liu, W. (2009). Nosiheptide biosynthesis featuring a

unique indole side ring formation on the characteristic thiopeptide framework.

ACS Chem. Biol. 4, 855–864.

Ltd All rights reserved