architectures, mechanisms and molecular evolution of natural product methyltransferases

13
Architectures, mechanisms and molecular evolution of natural product methyltransferases†‡ David K. Liscombe,x Gordon V. Louie and Joseph P. Noel * Received 5th March 2012 DOI: 10.1039/c2np20029e Covering: up to January 2012 The addition of a methyl moiety to a small chemical is a common transformation in the biosynthesis of natural products across all three domains of life. These methylation reactions are most often catalysed by S-adenosyl-L-methionine (SAM)-dependent methyltransferases (MTs). MTs are categorized based on the electron-rich, methyl accepting atom, usually O, N, C, or S. SAM-dependent natural product MTs (NPMTs) are responsible for the modification of a wide array of structurally distinct substrates, including signalling and host defense compounds, pigments, prosthetic groups, cofactors, cell membrane and cell wall components, and xenobiotics. Most notably, methylation modulates the bioavailability, bioactivity, and reactivity of acceptor molecules, and thus exerts a central role on the functional output of many metabolic pathways. Our current understanding of the structural enzymology of NPMTs groups these phylogenetically diverse enzymes into two MT-superfamily fold classes (class I and class III). Structural biology has also shed light on the catalytic mechanisms and molecular bases for substrate specificity for over fifty NPMTs. These biophysical-based approaches have contributed to our understanding of NPMT evolution, demonstrating how a widespread protein fold evolved to accommodate chemically diverse methyl acceptors and to catalyse disparate mechanisms suited to the physiochemical properties of the target substrates. This evolutionary diversity suggests that NPMTs may serve as starting points for generating new biocatalysts. 1 Introduction 2 Methyl acceptor diversity 3 Primary structure of NPMTs and identification of NPMT genes 4 Structural biology of NPMTs 5 Architecture of SAM binding 6 Structural basis of NPMT substrate specificity 7 Mechanisms of SAM-dependent methylation 8 Evolution of NPMTs 9 Engineering of NPMTs 10 Conclusions and future perspectives 11 Acknowledgements 12 References 1 Introduction Methylation is a ubiquitous biotransformation in nature, used throughout all branches of metabolism and often key to meta- bolic homeostasis. These transformations modulate diverse biological processes such as cell signaling, and the biosynthesis of complex and sometimes unique specialized metabolites. These reactions are most often catalysed by methyltransferases [MTs; sometimes referred to as (trans)methylases] that rely on the co- substrate{ S-adenosyl-L-methionine (SAM 1; Fig. 1) as an elec- tron-deficient methyl donor. The by-product of methylation is S-adenosyl-L-homocysteine (SAH 2; Fig. 1). Small molecule, or natural product, methyltransferases (NPMTs) participate in the biosynthesis and modification of bioactive molecules derived from several branches of primary and secondary (specialized) metabolism, including membrane components, 1,2 cofactors, 3 prosthetic groups, 4 pigments, 5,6 and signaling 7,8 and defense compounds. 9–11 Howard Hughes Medical Institute, Jack H. Skirball Center for Chemical Biology and Proteomics, Salk Institute for Biological Studies, La Jolla, CA 92037, USA. E-mail: [email protected] † This paper is part of an NPR themed issue on Structural Aspects of Biosynthesis. ‡ In memory of Prof. Dr Joachim Schroder and his contributions to our understanding of plant metabolism. x Current address: Vineland Research and Innovation Centre, Vineland Station, ON, Canada; e-mail: [email protected] { While often classified as a cofactor, herein we use co-substrate to describe SAM, because SAH is catabolized and SAM is not regenerated after methyl transfer. This journal is ª The Royal Society of Chemistry 2012 Nat. Prod. Rep. Dynamic Article Links C < NPR Cite this: DOI: 10.1039/c2np20029e www.rsc.org/npr REVIEW Downloaded by Tokyo Daigaku on 04 September 2012 Published on 01 August 2012 on http://pubs.rsc.org | doi:10.1039/C2NP20029E View Online / Journal Homepage

Upload: yangxiaolong2013

Post on 08-Jul-2015

333 views

Category:

Documents


4 download

TRANSCRIPT

Page 1: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Dynamic Article LinksC<NPR

Cite this: DOI: 10.1039/c2np20029e

www.rsc.org/npr REVIEW

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

EView Online / Journal Homepage

Architectures, mechanisms and molec

ular evolution of natural productmethyltransferases†‡

David K. Liscombe,x Gordon V. Louie and Joseph P. Noel*

Received 5th March 2012

DOI: 10.1039/c2np20029e

Covering: up to January 2012

The addition of a methyl moiety to a small chemical is a common transformation in the biosynthesis of

natural products across all three domains of life. These methylation reactions are most often catalysed

by S-adenosyl-L-methionine (SAM)-dependent methyltransferases (MTs). MTs are categorized based

on the electron-rich, methyl accepting atom, usually O, N, C, or S. SAM-dependent natural product

MTs (NPMTs) are responsible for the modification of a wide array of structurally distinct substrates,

including signalling and host defense compounds, pigments, prosthetic groups, cofactors, cell

membrane and cell wall components, and xenobiotics. Most notably, methylation modulates the

bioavailability, bioactivity, and reactivity of acceptor molecules, and thus exerts a central role on the

functional output of many metabolic pathways. Our current understanding of the structural

enzymology of NPMTs groups these phylogenetically diverse enzymes into two MT-superfamily fold

classes (class I and class III). Structural biology has also shed light on the catalytic mechanisms and

molecular bases for substrate specificity for over fifty NPMTs. These biophysical-based approaches

have contributed to our understanding of NPMT evolution, demonstrating how a widespread protein

fold evolved to accommodate chemically diverse methyl acceptors and to catalyse disparate

mechanisms suited to the physiochemical properties of the target substrates. This evolutionary diversity

suggests that NPMTs may serve as starting points for generating new biocatalysts.

1 Introduction

2 Methyl acceptor diversity

3 Primary structure of NPMTs and identification of

NPMT genes

4 Structural biology of NPMTs

5 Architecture of SAM binding

6 Structural basis of NPMT substrate specificity

7 Mechanisms of SAM-dependent methylation

8 Evolution of NPMTs

9 Engineering of NPMTs

10 Conclusions and future perspectives

11 Acknowledgements

12 References

Howard Hughes Medical Institute, Jack H. Skirball Center for ChemicalBiology and Proteomics, Salk Institute for Biological Studies, La Jolla,CA 92037, USA. E-mail: [email protected]

† This paper is part of an NPR themed issue on Structural Aspects ofBiosynthesis.

‡ In memory of Prof. Dr Joachim Schr€oder and his contributions to ourunderstanding of plant metabolism.

x Current address: Vineland Research and Innovation Centre, VinelandStation, ON, Canada; e-mail: [email protected]

This journal is ª The Royal Society of Chemistry 2012

1 Introduction

Methylation is a ubiquitous biotransformation in nature, used

throughout all branches of metabolism and often key to meta-

bolic homeostasis. These transformations modulate diverse

biological processes such as cell signaling, and the biosynthesis of

complex and sometimes unique specialized metabolites. These

reactions are most often catalysed by methyltransferases [MTs;

sometimes referred to as (trans)methylases] that rely on the co-

substrate{ S-adenosyl-L-methionine (SAM 1; Fig. 1) as an elec-

tron-deficient methyl donor. The by-product of methylation is

S-adenosyl-L-homocysteine (SAH 2; Fig. 1). Small molecule, or

natural product, methyltransferases (NPMTs) participate in the

biosynthesis and modification of bioactive molecules derived

from several branches of primary and secondary (specialized)

metabolism, including membrane components,1,2 cofactors,3

prosthetic groups,4 pigments,5,6 and signaling7,8 and defense

compounds.9–11

{ While often classified as a cofactor, herein we use co-substrate todescribe SAM, because SAH is catabolized and SAM is notregenerated after methyl transfer.

Nat. Prod. Rep.

Page 2: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

Since the first X-ray crystal structure of a NPMT12 reported

early in the 1990’s, dozens of structures have been elucidated for

NPMTs representing a chemically diverse array of methyl

acceptors. Surprisingly, all currently known NPMT structures

belong to either of two (Class I and Class III)13 of 15 currently

recognized protein fold superfamilies of SAM-binding

proteins.14 Even so, the large majority belong to the Class I or

Rossmann-like fold family.15 The structural elucidation of

NPMTs advanced our understanding of both the molecular

determinants for substrate specificity and the varied catalytic

mechanisms of this class of enzymes, while also providing a

foundation for structure-based engineering to generate new

enzymes with altered specificities. Macromolecular (DNA,

Gordon V: Louie

Gordon Louie obtained B.Sc.

Honours (Biochemistry and

Chemistry) and Ph.D.

(Biochemistry) degrees from

the University of British

Columbia. He undertook post-

doctoral research in the labora-

tories of Prof. Thomas Blundell

at Birkbeck College, Univ. of

London (where he determined

the structure of a porphobili-

nogen deaminase, a key enzyme

in tetrapyrrole biosynthesis),

and Prof. Senyon Choe at the

Salk Institute for Biological

Studies (where he characterized

the interaction of diphtheria toxin with its receptor). He subse-

quently worked as a structural biologist at SGX Pharmaceuticals.

Gordon is presently a Research Associate in Prof. Joseph Noel’s

lab at the Salk Institute, focusing on structural analyses of enzymes

of phenylpropanoid metabolism.

David K: Liscombe

David Liscombe completed the

Honours Biology and Pharma-

cology Co-op program at

McMaster University in 2003.

He received his Ph.D. in plant

biochemistry from the Univer-

sity of Calgary in 2008, which

involved the characterization of

methyltransferases from benzy-

lisoquinoline alkaloid biosyn-

thesis. As a postdoc in Sarah

O’Connor’s lab at MIT, David

discovered genes involved in

terpenoid indole alkaloid

biosynthesis. He subsequently

moved to Joseph Noel’s lab at

the Salk Institute to investigate the structural biology of methyl-

transferases and other specialized biosynthetic enzymes. David is

currently a Research Associate in Applied Genomics at Vineland

Research and Innovation Centre in Ontario, Canada.

Nat. Prod. Rep.

RNA, and protein) MTs are essential epigenetic regulators of

gene expression and chromatin structure, and post-translational

modulators of protein function. The structures and functions of

macromolecular MTs are reviewed elsewhere.13,15 In this review,

we discuss the structure, function, and evolution of SAM-

dependent NPMTs, focusing on those with structures reported

in the literature. A compilation of NPMTs with published

structures is summarized in Table 1.

2 Methyl acceptor diversity

All MTs (EC 2.1.1.-) are classified according to the substrate

atom that accepts the methyl group, usually O (54% of EC

subclass), N (23%), or C (18%). S-directed MTs (3% of EC

subclass) and NPMTs that accommodate other acceptors (such

as halides; 2%) are rare but notable,16–18 and some NPMTs

transfer a methyl moiety to more than one type of acceptor

atom.19

Considering all domains of life, the most abundant NPMTs

are O-directed MTs (OMTs). The OMT subfamily in certain

plants and bacteria has undergone tremendous genetic and

functional expansion. For example, poplar trees (Populus sp.)

encode 26 small molecule OMTs,20 whereas only two OMTs are

found in humans [catechol OMT (COMT) and N-acetyl-sero-

tonin OMT] and yeast [Saccharomyces cerevisiae; trans-aconitate

methyltransferase (TMT1)21 and cantharidin resistance gene

(Crg1)22]. Humans and yeast are not remarkable sources of

complex natural products, however, so their lack of encoded

OMT genes is not surprising as small molecule OMTs participate

almost exclusively in specialized metabolic pathways. Hydroxyl

moieties of phenolics, such as catechol, and hexoses, along with

carboxylic acids or CoA esters are the most common substrates

for NPMTs (Fig. 2). OMTs tend to be regiospecific, but some,

such as those involved in phenylpropanoid and flavonoid

biosynthesis in photosynthetic organisms, are less selective,

Joseph P: Noel

Joseph Noel obtained a Bach-

elor of Science degree in Chem-

istry from the University of

Pittsburgh at Johnstown in

1985. He received his Ph.D. in

chemistry from the Ohio State

University in 1990, working on

the enzymology of phospholi-

pases with Professor Ming-Daw

Tsai. As a postdoctoral fellow

with the late Paul B. Sigler in

the Department of Molecular

Biophysics and Biochemistry at

Yale, Joe elucidated the struc-

ture of heterotrimeric G-

proteins. Joe is currently

director of the Jack H. Skirball Center for Chemical Biology and

Proteomics at the Salk Institute, professor at the Salk Institute and

investigator with the Howard Hughes Medical Institute.

This journal is ª The Royal Society of Chemistry 2012

Page 3: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Fig. 1 S-Adenosyl-L-methionine (SAM) the methyl donor for SAM-

dependent natural product methyltransferases (NPMTs). (A) NPMTs

use SAM co-substrate as a reactive electron deficient methyl group

(green) donor for transfer to a electron-rich methyl acceptor (Nu:). (B) In

addition to the methylated product, S-adenosyl-L-homocysteine (SAH)

forms and is a potent inhibitor of SAM-dependent MTs. (C) Sinefungin,

a fungal-derived SAM-analog possessing an amine group (red) in place

of the methylsulfonium moiety serves as a competitive inhibitor of

SAM-dependent MTs.

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

capable of sequential methylations of the same or similar

substrates.23–25

Natural product N-directed MTs (NMTs), though not as

numerous as OMTs, are broadly represented across all domains

of life. NMTs are commonplace in signal transduction pathways

in animals, where they modulate the activity of signaling mole-

cules [COMT, phenylethanolamine NMT (PNMT), histamine

NMT (HNMT), indolethylamine NMT (INMT)]. There are also

a few examples of iterative NMTs. Plasmodium falciparum

phosphoethanolamine NMT (PfPEANMT) catalyzes the

This journal is ª The Royal Society of Chemistry 2012

trimethylation of phosphoethanolamine in phosphocholine

biosynthesis.2 Caffeine biosynthesis in certain plants employs a

bifunctional NMT, dimethylxanthine NMT (DXNMT), which

performs sequential methylations at two N-containing sites.26

Similarly, bacterial NMTs TylM1 and DesVI catalyze analogous

dimethylations of hexosamine moities.27,28 Characterized natural

product NMTs are collectively responsible for the methylation of

a wide variety of substrates, such as non-ribosomal peptides

(both peptide bonds and side chains), hexosamines, primary

amines, secondary amines (i.e. indoles, imidazoles, more complex

alkaloids), and tertiary amines (Table 1, Fig. 3).

Small molecule C-methyltransferases are most often found in

bacterial and plant systems, and are relatively scarce in other

branches of the eukaryotic lineage. Yeast encodes only two,21

and no small molecule CMTs have been detected or predicted to

occur in humans. Like OMTs, the known CMTs participate

primarily in specialized metabolism (Fig. 3), methylating

substrates such as tetrapyrroles, phenolics, aliphatics, and

hexos(amin)es. We also consider cyclopropane synthases as

SAM-dependent MTs, although they technically transfer a

methylene group.1

The few S-directed MTs identified and characterized to date

exist in plants and mammals. In plants, they produce

volatile halogen and sulfur compounds or biosynthetically tailor

thiocyanates.16,29,19,30,31 Human thiopurine S-MTs (TPMT)

participate in the detoxification of xenobiotics.18

3 Primary structure of NPMTs and identification ofNPMT genes

The NPMTs vary in length, typically spanning 200–500 amino-

acid residues corresponding to monomeric molecular masses of

ca. 25–55 kDa. Almost all NPMTs isolated to date possess an a/b

structure: alternating a-helices and b-strands along the length of

the polypeptide chain. Comparative sequence analyses of SAM-

dependent MTs have identified a series of conserved motifs

shared among these proteins.14,32–35 Generally arranged in

sequential order across the core MT domain, Motifs I–VI reside

in regions associated with SAM co-substrate binding (Fig. 4).

These motifs are widely conserved across NPMTs, albeit to

varying degrees, and are considered to be defining features of

SAM-dependent MTs.14,32,33,36,37 Kozbial and Mushegian (2005)

provide the most recent survey of conserved motifs in MTs.14

Motif I is present in the majority of MTs (69 of 84 MT proteins

surveyed32) and is often used for the initial bioinformatic iden-

tification of putative MTs. This motif spans the loop preceding

the first b-strand (b1) of the core Rossmann fold leading into

the following a-helix (aA). It includes a nine-residue amino

acid block with the consensus sequence (V/I/L)(L/V)(D/E)(V/I)-

G(G/C)G(T/P)G.32 This nine-residue structure incorporates the

glycine-rich ‘‘GxGxG’’ signature sequence, a SAM-binding motif

found in almost all SAM-dependent MTs.14,32,35,38 Although

none of the three glycines of the GxGxG motif is universally

conserved, substitutions typically encompass small

sidechains.14,38

Motif II spans b2 and the adjoining turn. Here, the consensus

sequence in plant OMTs (aka Motif B35) differs from that

observed in a larger survey of functionally-diverse MTs.32 Only

two residues, DA, are common to both consensus sequences,35

Nat. Prod. Rep.

Page 4: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Table 1 Published tertiary structures of natural product methyltransferases

Name Organism Acceptora PathwaySubstrateclass Foldb

Structuralsimilarityc Mechanismd

Metal-dependent? PDBe Ref.

BACTERIALDnrK Streptomyces peucetius O, 5 daunorubicin phenolic I, D RdmB PD No 1TW2 80RebM Lechevalieria

aerocolonigenesO, 6 rebeccamycin hexose I, M CPFASs,

DnrKAB No 3BUS 86

SynOMT Synechocystis sp. strainPCC 6803

O, 12 hydroxycinnamicacids

phenolic I, D PFOMT,COMT,

M Yes 3CBG 24

BcOMT2 Bacillus cereus O, 14 ? phenolic I, D CCoAOMT,COMT

M Yes 3DUL 79

NcsB1 Streptomycescarzinostaticus

O, 11 neocarzinostatin phenolic I, D DnrK, RdmB, AB No 3I53 85

DhpI Streptomyces luridus O, 19 dehydrophos phosphoryl I, D NNMT ? No 3OU2 87NovP Streptomyces spheroides O, 21 novobiocin hexose I, D COMT,

SynOMTAB(M) Yes 2WK1 78

CalO1 Micromonosporaechinospora LL6600

O, 16 calicheamicin phenolic I, D ChOMT,MmcR

AB No 3LST 88

MmcR Streptomyces lavendulae O, 18 mitomycin quinone/mitosane

I, D DnrK, NcsB1 AB No 3GWZ 89

MycE Micromonosporagriseorubida

O, 20 mycinamicin hexose I, T COMT AB(M) Yes 3SSM 50

PhzM Pseudomonas aeruginosa N, 22 procyanin phenazine I, D CaOMT,IOMT

PD? No 2IP2 90

DesVI Streptomyces venezuelae N, 23 erythromycin hexosamine(primary)

I, D GNMT,TylM1

PD No 3BXO 28

MtfA Amycolatopsis orientalis N, 25 chloroeremomycin primaryamine

I, D PhzM, DesVI AB No 3G2M 84

NodS Bradyrhizobium japonicum N, 28 nodulation factor hexosamine(primary)

I, M DhpI AB No 3OFJ 91

TylM1 Streptomyces fradiae N, 24 tylosin hexosamine(primary)

I, D DesVI,GNMT

AB? No 3PFG 27

CbiF Bacillus megaterium C, 43 vitamin B12 tetrapyrrole III,D CobA PD No 1CBF 3CPFASs Mycobacterium

tuberculosisC, 38 cyclopropyl lipids aliphatic I, D RebM AB No 1KP9 1

CbiT Methanothermobacterthermautotrophicus

C, 41 vitamin B12 n/a I, T COMT n/a n/a 1KXZ 54

CysG Salmonella enterica C, 41 siroheme tetrapyrrole III,D

CobA AB No 1PJQ 4

CobA Pseudomonas denitrificans C, 41 tetrapyrroles tetrapyrrole III,D

CbiF AB No 1S4D 56

BchU Chlorobium tepidum C, 41 bacteriochlorophyllc

tetrapyrrole I, D PhzM,LpCaOMT

AB No 1X19 53

MtCbiL Methanothermobacterthermoautotrophicus

C, 41 tetrapyrroles tetrapyrrole III,D

CbiF, CobA AB No 2QBU 57

CbiL Chlorobium tepidum C, 41 vitamin B12 tetrapyrrole III,D

CbiF, CobA AB No 2E0K 55

TcaB9 Micromonospora chalcea C, 39 D-tetronitrose hexosamine I, M 3DLI AB No 3NDI 52GPPMT Streptomyces lasaliensis C, 40 2-methylisoborneol aliphatic (I) RebM ? ? n/a 92NirE Pseudomonas aeruginosa C, 41 heme d1 cofactor tetrapyrrole III,

DCobA AB No 2YBO 58

LiOMT Leptospira interrogans ? ? ? I, D CCoAOMT M (Yes) 2HNK 93RdmB Streptomyces purpurascens n/a anthracycline tertiary

carbonI, D IOMT, DnrK n/a No 1QZZ 37,

75

PLANTChOMT Medicago truncatula O, 10 chalcone phenolic I, D CaOMT,

IOMTAB No 1FPQ 5

I7OMT Medicago truncatula O, 8 isoflavone phenolic I, D CaOMT,ChOMT

AB No 1FPX 5

MtCaOMT Medicago sativa O, 13 phenylpropanoid phenolic I, D I7OMT,CalO1

AB No 1KYW 77

SAMT Clarkia breweri O, 7 salicylic acidsignalling

phenolic I, D DXMT,XMT, HNMT

PD No 1M6E 7

CCoAOMT Medicago sativa O, 18 hydroxycinnamyl-CoAs

phenolic I, D COMT M Yes 1SUI 51

HIOMT Medicago truncatula O, 8 isoflavone phenolic I, D I7OMT? AB No 1ZHF 9PFOMT Mesembryanthemum

crystallinumO, 15 phenylpropanoid phenolic I, D CCoAOMT M Yes 3C3Y 25

IAMT Arabidopsis thaliana O, 9 indole acetic acid carboxy(indole)

I, D SAMT PD No 3B5I 74

LpCaOMT Lolium perenne O, 13 phenylpropanoid phenolic I, D AB No 3P9C 23

Nat. Prod. Rep. This journal is ª The Royal Society of Chemistry 2012

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

Page 5: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Table 1 (Contd. )

Name Organism Acceptora PathwaySubstrateclass Foldb

Structuralsimilarityc Mechanismd

Metal-dependent? PDBe Ref.

I7OMT,CalO1

AtHOL1 Arabidopsis thaliana NCS- > I >Br > Cl

thiocynates, halides(not F)

thiocyanate/halide

I, D TPMT ? No 3LCC 19

DXMT Coffea canephora N, 27, 29 caffeinebiosynthesis

purine I, D SAMT PD No 2EFJ 26

XMT Coffea canephora N, 26 caffeinebiosynthesis

purine I, D SAMT PD No 2EG5 26

ANIMALTPMT Mus musculus S, 42 thiopurine

xenobioticsthiopurine I, M PNMT,

HNMTAB No 3BGD 18

COMT Rattus norvegicus O, 4 catechol phenolic I, M PFOMT AB/M?14 Yes 1VID 12GNMT R. norvegicus N, 30 glycine primary

amineI, T TylM1 AB No 1XVA 72

HNMT Homo sapiens N, 31 histamine imidazole I, M TPMT PD No 1JQD 94PNMT H. sapiens N, 32 adrenaline primary

amineI, M GNMT PD No 1HNN 8

GANMT R. norvegicus N, 34 creatine secondaryamine

I, D COMT,GNMT

AB No 1KHH 95,96

PfPENMT Plasmodium falciparum N, 35 phosphocholine primary/s/tertamine

I, D CPFASs AB No 3UJ6 2

NNMT H. sapiens N, 33 nicotinamide pyridine I, M PNMT,INMT

PD No 3ROD 73

a Acceptor atom, compound number in Fig. 2–3. b Class, oligomerization state (M, monomer; D, dimer; T, tetramer). c Structural homologs in italicswere determined using the DALI server (http://ekhidna.biocenter.helsinki.fi/dali_server/), or from PDB records at http://www.rcsb.org. d PD, proximityand desolvation effects; AB, acid–base; AB(M) acid–base with possible metal participation; M, metal-dependent. e Representative PDB code provided.

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

with the aspartate residue reflecting the conservation of an acidic

residue near the C-terminus of b2.

Motif III spans b3, followed by Motif IV spanning b4 and the

adjoining loops. Both motifs include a partially-conserved acidic

residue at the C-terminus.14 Motif V, occurring in the helix (aC)

following Motif IV, sometimes harbors hydrophobic residues

that sandwich the adenine moiety of SAM.14 A well-conserved

glycine residue is characteristic of Motif VI, which corresponds

to b5 and the loop between aC and b5.14 Some NPMTs contain

multiple MT domains,39 or contain heterologous domains that

catalyze reactions other than methylation.4

Many NPMT genes have been isolated based on these

conserved motifs. PCR-based screening with degenerate primers

restrained to the Motif I consensus sequence have been used to

amplify fragments of NPMT genes.40–42 Transcriptome and

genome sequencing projects often rely on these sequence signa-

tures and homology to known MTs as a means of isolating

previously undiscovered NPMT genes.10,42,43 Computational

methods based on motif scanning and/or structural homology

have been developed for the identification and characterization

of MT-encoding genes, to elucidate the ‘methyltransferome’ in

whole genomes20,21,44–47 and to define MT domains within

biosynthetic gene clusters.38

4 Structural biology of NPMTs

Protein X-ray crystallography has been used extensively over the

last two decades to determine the tertiary and quaternary

structures of numerous NPMTs (Table 1). The first MT structure

solved was the cytosine C5-specific DNA MT, M.HhaI.48 The

This journal is ª The Royal Society of Chemistry 2012

SAM binding domain of M.HhaI incorporates the Rossmann

fold,48 an a/b domain well known for binding nucleotide-con-

taining cofactors, such as NAD.49 Soon after, the first structure

of an NPMT was reported,12 and surprisingly, the fold of cate-

chol OMT (COMT; Table I1) was strikingly similar to the DNA

MT. Indeed, both enzymes include a core seven b-strand Ross-

mann fold, which serves as a SAM-binding domain. Further-

more, structural comparison of M.HhaI, COMT, and NAD-

dependent alcohol dehydrogenases demonstrates that the

respective Rossmann folds and bound nucleotide-based cofactor/

co-substrate superimpose remarkably well (Fig. 5).

The similarity between SAM binding domains of M.HhaI and

COMT suggested initially that all MTs might share a common

structure.12 The so-called Class I, Rossmann-like MT fold is

shaped by the alternating a-helices and b-strands of the poly-

peptide (Fig. 4, 6a), which form a relatively planar b-sheet

sandwiched by a-helices (Fig. 6b). The N-terminal b-strand

inserts into the middle of the b-sheet, such that the strand

topology is 3214576, with the seventh strand antiparallel to all

other strands (Fig. 5a). The functionally important, conserved

residues of the substrate-binding and catalytic sites are typically

located in the C-terminal regions of b-strands or in the adjoining

loops. This core fold is often elaborated by additional helices or

b-hairpins.

Since the structure of COMT was reported, more than 50

NPMT structures have been solved and published (Table 1), and

even more structures have been deposited in the Protein Data

Bank (PDB) but not yet described in the literature. These anal-

yses further establish the structural conservation of the core of

most NPMTs, and additionally delineate the structural diversity

Nat. Prod. Rep.

Page 6: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Fig. 2 Chemical diversity of natural product OMT substrates. Compounds are numbered, and named as appropriate. Methylation target sites are

highlighted in green. Structurally-characterized NPMTs and their substrate numbers are indicated in Table 1.

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

of the embellishing domains appended to or inserted into the

primary structure of the core Rossman-like fold. Most

commonly, the additional domains represent N-terminal exten-

sions that mediate oligomerization (quaternary structure) and/or

modulate substrate specificity.14

For a number of NPMTs, catalytic activity is dependent on a

coordinated divalent cation (Table 1), such as Mg2+ (magne-

sium)51 or Ca2+ (calcium),52 and structural studies have eluci-

dated the architecture and functional role of the metal binding

site. Mg2+- and Ca2+-dependent NPMTs typically use an acidic

triad [DD(D/N)] for metal coordination. A zinc-containing

NPMT uses four cysteine residues,53 although metal binding

occurs distal to the active site and may serve only in structure

stabilization. As best exemplified by the metal-dependent OMTs,

divalent cations typically participate directly in substrate binding

and/or catalysis.

The initial presumption that all NPMTs share the same

structural core, a Rossmann-like fold, was disproved with the

structural characterization of cobalt-precorrin-4-MT (CbiF) and

the discovery of the Class III MT fold.3 CbiF and five other

closely related MTs act sequentially in the biosynthesis of the

corrin ring of vitamin B12 (a tetrapyrrole) in bacteria. To date,

the Class III fold associates with tetrapyrrole MTs only.

Conversely, other NPMTs known to methylate tetrapyrrole-

Nat. Prod. Rep.

containing substrates, namely BchU53 and CbiT,54 belong to the

Class I fold.

CbiF and its Class III relatives are more closely related to the

GTPase fold typified by a kidney-shaped arrangement of two a/b

domains linked by a single coil (Fig. 6).3,4,55–58 While there is no

topological similarity between the two a/b domains, both contain

a five-stranded b-sheet sandwiched by four a-helices (Fig. 6). A

‘‘GxGxG’’ motif is located in the C-terminal end of b1 and the

loop leading to aA of Class III NPMTs, a region that is also

ascribed to SAM binding3 in this family of tertiary structures.

5 Architecture of SAM binding

In all Class I NPMTs structurally characterized to date, SAM

occupies the spatially equivalent position along the C-terminal

end of the core b-sheet of the Rossmann-like domain (Fig. 7A,B),

despite only weak conservation of the SAM-binding residues.

The Class I NPMTs bind SAM (or SAH) in an extended

conformation maintained by a network of hydrogen bonds and

van der Waals interactions5 involving residues alongMotifs I–III

(Fig. 7A,B). One or more residues in the GxGxG motif are in

contact with the carboxypropyl portion of SAM, while the

conserved acidic residue in Motif II forms hydrogen bonds with

the ribosyl moiety (Fig. 7B). Variable residues in the C-terminal

This journal is ª The Royal Society of Chemistry 2012

Page 7: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Fig. 3 Chemical diversity of natural product N-, C-, and S-directed MT substrates. Compounds are numbered and named. Methylation sites are

highlighted in green. Structurally-characterized NPMTs and their substrate numbers are indicated in Table 1. Due to the size and complexity of

tetrapyrrole substrates, the regiospecificities of uroporphyrinogen-like tetrapyrrole MTs (CysG, NirE, CobA, BchU, and CbiL) are indicated using a

tetrapyrrole scaffold for illustrative purposes (compound 41).

Fig. 4 A schematic diagram of the primary and secondary structure of a

typical NPMT, emphasizing conserved motifs used to identify a putative

MT. N- and C- termini are shown in black circles. a-Helices are shown in

red, b-strands in yellow, and adjoining loops are green. Conserved resi-

dues are stacked for each motif (I–VI), and the highly-conserved

‘‘GxGxG’’ motif is in bold fonts.

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

region of b2 (motif II) and the conserved acidic residue in b3

interact with the adenosyl moiety, while variable residues

C-terminal to b4 (Motif IV) appear to contact the amino and

sulfonium groups in the methionine fragment of SAM.14,32

Structure comparison of a number of metal-independent,

This journal is ª The Royal Society of Chemistry 2012

homodimeric plant OMTs reveals substantial flexibility in the

connection between the SAM binding domain and the central

core of the homodimer, which is formed by oligomerization and

substrate-binding domains.23 Thus, the MTs of this lineage

apparently utilize an open conformation for facilitating entry of

SAM and the phenolic substrate (and exit of the reaction prod-

ucts), and a closed conformation for establishing the catalytically

appropriate positioning of substrate and co-substrate prompting

transmethylation.23,85

All Class III MTs bind SAM in a jack-knifed conformation

such that, as Schubert and colleagues describe, SAM/SAH fits

into its binding site ‘‘similar to a two-pronged plug in a socket’’

(Fig. 7C,D).3 It is thought that this conformation might promote

presentation of the methyl group to the bulky precorrin

substrate.3 The SAM-binding site of Class III MTs involves side-

chain and main-chain interactions with residues located on the

b1-aA segment, the polypeptide loop linking b4 and aD, and aE

residues. These three sections line the large trough separating the

N- and C-terminal domains.3

Nat. Prod. Rep.

Page 8: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Fig. 5 The core fold of Class I NPMT incorporates a Rossmann fold.49

Similarly oriented structures of catechol OMT (A) with its bound

cosubstrate SAM (grey spheres), and cofactor-binding domain of alcohol

dehydrogenase (B; PDB: 6ADH, chain A), with bound NAD (grey

spheres), displayed side-by-side. Helices are shown in red, beta-strands in

yellow, and loops in green. Structures were initially superimposed using

the SSM superpose function in Coot.97 Image was generated with

MacPyMol.

Fig. 6 Topologies and folds of NPMTs. (A) Typical topology of a Class

I, Rossmann-like NPMT. (B) Tertiary structure of COMT with bound

SAH, representative of the Class I fold. (C) Topology and (D) tertiary

structure of CbiF with bound SAH illustrate the typical Class III

NPMTs. The diverse N-terminal region of Class I NPMTs is shown in

grey. The SAM/SAH binding site is highlighted in orange, bound SAH

ligands are shown in light blue, and conserved motifs are indicated with

Roman numerals (I, II, III). a-Helices are shown in red, b-strands in

yellow, and adjoining loops are green.

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

The by-product of methylation, SAH 2 (Fig. 1), is a potent

inhibitor of all SAM-dependent methyltransferases. In fact,

SAM binding sites in MT crystal structures are often occupied by

SAH, either intentionally through co-crystallization or crystal

soaks with SAH, or as a result of SAM degradation or enzyme-

catalyzed transmethylation. The binding affinity of SAH has

been exploited for MT isolation.59–61 Sinefungin 3, a fungal-

derived SAM analog and a competitive inhibitor of SAM-

dependent MTs,62–66 is also regularly employed in biochemical

and structural investigations of NPMTs. In vitro assays of MTs

can be adversely affected by the accumulation of the inhibitory

SAH by-product. Recently developed coupled assays of MT

activity utilize SAH-catabolizing enzymes, which act both to

prevent SAH accumulation and to generate a free thiol (homo-

cysteine) that can be quantified spectrophotometrically.67,68

Because the interaction of SAM with MTs typically does not

involve the donor methyl group, MTs can also readily accept

chemically generated S-substituted SAH analogs. Thus, as first

demonstrated with DNA MTs,69 NPMTs can utilize SAM

analogs and catalyze the transfer of a non-natural functional

group to a suitable acceptor atom.70,71

6 Structural basis of NPMT substrate specificity

NPMTs are capable of methylating an expansive repertoire of

substrates representing a diversity of chemical scaffolds found in

nature (Fig. 2–3). In contrast to DNA MTs, which possess some

conserved motifs for recognition of common features of the

macromolecular substrates,15 small molecule MTs do not appear

to possess widely conserved structural determinants for methyl

acceptor recognition. Instead, the core Rossmann-like fold of

Class I methyltransferases often bear structural elaborations,

including N-terminal extensions, discrete domains and active-site

caps. These structural appendages are critical determinants of the

functional evolution of MTs as they typically include amino acid

residues that contribute substantially to substrate binding and

the positioning of the methyl accepting atom.7,72,73 For example,

with most metal-independent, homodimeric plant OMTs, the

Nat. Prod. Rep.

N-terminal domains of the polypeptide chain are responsible

for substrate binding and dimerization.5,23,78 In these OMTs, the

C-terminal SAM-binding domain nevertheless plays a key role in

fully sequestering the substrate upon formation of the catalyti-

cally primed and closed conformational state.5,23,78 Indeed, non-

productive ligand complexes observed with alfalfa CaOMT77

may be representative of exploratory ‘‘pre-binding’’ modes of the

incoming substrate molecule with CaOMT that is in an open

conformational state and lacking a fully formed phenolic-

substrate binding pocket.23 Such considerations highlight the

importance of conformational dynamics for the substrate-

binding and catalytic activities of the MTs. Interestingly, heter-

odimerization among four distinct but closely related OMTs

involved in berberine biosynthesis can yield a substrate-speci-

ficity profile different from that of any of the four homodimeric

enzymes.41

Structural biology has contributed to our understanding of

substrate discrimination within NPMT subfamilies that are

known to accept similar chemical scaffolds, structure–function

relationships that can be understood only in the context of

appropriately configured active site architectures.7,9,23,74 Detailed

structural analyses of NPMTs bound to substrates, products, or

analogs have shed light on the enzymatic determinants of

substrate preferences.5,9,23For example, CaOMT possesses broad

This journal is ª The Royal Society of Chemistry 2012

Page 9: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Fig. 7 Binding modes of SAM NPMT active sites. (A) Superimposed

structures of a representative set of Class I NPMTs, CCoAOMT (grey)

with bound SAH (yellow), and SAMT (green) with bound SAH

(magenta) illustrating the canonical SAM/SAH binding site in Class I,

Rossmann-like NPMTs. (B) Zoomed-in view of SAM/SAH binding sites

using the same color scheme as (A). Only b-strands (b1-4) and adjoining

loops involved in co-substrate binding are depicted. Conserved motifs are

labeled, and the conserved acidic residues (in this case, both are Asp

residues) in Motif II are shown as sticks and labeled with an asterisk. (C)

Superimposed structures of CbiF (orange) with bound SAH (yellow), and

CobA (blue) with bound SAH (cyan) illustrating SAM/SAH binding sites

typical of Class III NPMTs. (D) Close-up view of Class III SAM/SAH

binding sites using same color scheme as (C). Residues within 4 �A of SAH

are shown, illustrating the conservation of the co-substrate binding site in

the Class III MT folds. Structures were superimposed using the SSM

superpose function in Coot.97 Images were generated with MacPyMol.

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

substrate selection (in vitro), with bi-functional activity in meta-

O-methylation of both the 3- and 5-hydroxyl groups of phenolic

monolignol-precursors bearing alcohol, aldehyde, or acid forms

of the propenyl side group. The interactions dictated by the

substrate’s C3 and C9(g) substituents likely govern the substrate-

preference patterns of the CaOMTs.23 In particular, for angio-

sperm CaOMTs, occupancy of a predominantly hydrophobic

cavity by a C3 methoxy-substituent on the substrate imposes a

preference for 5-O-methylation activity, while the presence of

only one hydrogen-bonding residue near the substrate’s C9

functionality underlies the preference for the singly oxygenated

alcohol and aldehyde forms of the monolignol precursor.

The initial characterization of putative NPMTs is challenging

due to the unknown identity of the true in vivo substrate(s) and

the unavailability of synthetic substrates suitable for assays and

further functional studies. Homology, or even substantial

sequence similarity, does not necessarily reveal useful informa-

tion as to the nature of the methyl acceptor or substrate

This journal is ª The Royal Society of Chemistry 2012

specificity of an orphan NPMT. For example, an NMT involved

in terpenoid indole alkaloid biosynthesis in Madagascar peri-

winkle (Catharanthus roseus) is unexpectedly most similar to

g-tocopherol CMTs involved in vitamin E biosynthesis. Indeed,

these enzymes methylate structurally and chemically disparate

substrates (36 versus 37 in Fig. 3) originating in very distinct

realms of metabolism.10

Phylogenetic analyses can assist in predicting substrate speci-

ficity, but even close phylogenetic relationships can be

misleading. For example, RdmB is phylogenetically and struc-

turally related to the OMT DnrK, but RdmB functions as a

hydroxylase.37,75 Furthermore, there are no obvious trends in

mechanistic strategies of NPMTs with respect to acceptor spec-

ificity (Table 1), except that NMTs and CMTs are rarely, if ever,

metal-dependent. This presence or absence of metal cation

binding further illustrates how structural enzymology can

provide important clues to facilitate the rapid elucidation of

substrate selection. Thus, even the most readily discernable

structural information, primary structure, can also provide

some hints, based on the presence of metal-coordinating motifs

(i.e. DDD/N).

Finally, the development of a substrate tagging approach

using SAM analogues and crosslinking offers additional exper-

imental methods for elucidating substrate specificity.70 Targeted,

activity-based metabolite profiling has been employed to identify

in vivo substrates for NPMTs,10 and a systems biology approach

has efficiently elucidated the biological functions of an orphan

NPMT.22

7 Mechanisms of SAM-dependent methylation

NPMT-catalyzed transmethylation occurs via SN2-like nucleo-

philic substitution,76 and detailed structural analyses of NPMTs

reveal several requisite features of this reaction. Firstly, the

active-site architecture of the NPMTmust ensure that the methyl

acceptor is positioned reasonably close to the donor electron

deficient methyl group of SAM, usually within about 3 �A (to

methyl carbon, or ca. 4 �A between acceptor and sulfonium

moiety; Fig. 8). In addition, the acceptor must serve as the most

chemically reactive and spatially proximal nucleophile in the

vicinity of the electron deficient methyl moiety, a consideration

that may demand the enzymic activation of the acceptor through

proton abstraction or in some cases the extrusion of water. As

such, NPMTs have evolved three currently recognized chemical

mechanisms for catalyzing the transmethylation reaction: (i)

proximity and desolvation, (ii) general acid/base-mediated

catalysis, and (iii) metal-dependent mechanisms (Fig. 8).

A primary mechanistic role for proximity and desolvation

(PD, Fig. 8A) in an NPMT-catalyzed transmethylation reaction

was first posited with the structural characterization of SAMT, a

defining member of the SABATH (Salicylic Acid, Benzoic Acid,

THeobromine synthase) family of plant MTs.7 The PD mecha-

nism does not require the direct participation of a catalytic

group(s) from the enzyme, but rather the architecture and

chemical environment of the enzyme active-site ensure that the

acceptor is in close proximity to the donor methyl group and

suitably oriented for nucleophilic substitution, and that water

(solvent) molecules are excluded from the donor–acceptor

interface (desolvation).7 Many other MTs also appear to utilize

Nat. Prod. Rep.

Page 10: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Fig. 8 Catalytic strategies for methylation used by NPMTs. (A) The

‘‘proximity and desolvation effects’’ mechanism, exemplified by DnrK-

mediated anthracycline O-methylation.80 The product, 4-methoxy-3-

rhodomycin T (magenta), and SAH (yellow) from a ternary complex with

DnrK are shown. Mutagenesis of the closest possible general base, Y142

(grey), did not have a substantial effect on catalytic rate. The methylated

oxygen, which is in proximity (distance in green) to the sulfonium group

of SAM is indicated with an arrow. (B) The ‘‘acid-base’’ mechanism of

methyl-transfer demonstrated by PfPEANMT.2 His and Tyr residues

(green) work in-concert as a ‘general base’ to deprotonate the substrate,

in this case phosphoethanolamine (magenta) driving the SN2-transfer of

the methyl group from SAM (yellow), which is proximal to the acceptor

nitrogen (distance shown in green). (C) Metal-dependent methylation

catalysis illustrated by CCoAOMT. Metal-mediated deprotonation of

the acceptor hydroxyl group generates an oxyanion adjacent (distance in

green) to the reactive and electron deficient methyl group of SAM,

Nat. Prod. Rep.

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

this strategy (i.e.DnrK, Fig. 8A), irrespective of methyl acceptor

or substrate specificity (Table I).

General acid/base-mediated transmethylation (Fig. 8B) by an

NPMT typically involves an essential catalytic residue that acts

as a general base to deprotonate and thereby activate the methyl

acceptor for nucleophilic attack on the reactive methyl group of

SAM.5,77 Adjacent active-site residues often ensure optimal

orientation of the catalytic base, or work in concert to form a

proton shuttle system (Fig. 8B).2,5 The cyclopropane synthases

(CPFASs)1 employ a carbonate ion as a general base, which may

deprotonate the methyl group of the carbocation intermediate

and thus enable formation of the cyclopropane product.

Conserved residues involved in carbonate-ion binding seem to

distinguish methylene transferases from other NPMTs.14

There appear to be two types of metal-dependent NPMTs.

First are those that rely on a divalent cation solely for substrate

coordination and require a nearby residue as a general base for

deprotonation of the methyl acceptor (i.e. hexose MTs like

NovP,78 and MycE50). Metal-dependent NPMTs of the second

type (almost exclusively phenolic OMTs from plants) use a

catalytic mechanism (Fig. 8C) in which the metal ion perturbs the

pKa of the substrate’s phenolic hydroxyl group and thus

promotes loss of the hydroxyl proton (as a hydronium ion) and

formation of a nucleophilic phenolate anion.24,25,51,79 Such a

reaction mechanism was originally proposed for COMT, but

subsequent studies suggest that an active-site lysine residue is

responsible for deprotonation of the acceptor, indicative of a

general acid/base mechanism.15

8 Evolution of NPMTs

Kozbial andMushegian suggested that the last universal common

ancestor of cellular life possessed asmany as twenty SAM-binding

proteins from at least five distinct fold classes.14 Class I MTs are

predicted to have been well represented in this primordial reper-

toire of enzymes,which probably included activities for producing

SAM, synthesizing polyamines, and methylating several

substrates.14 In general, the subsequent evolutionary trajectories

of macromolecule MTs and small-molecule MTs diverged

significantly, as no single small-molecule MT appears to be

conserved across all domains of life. Therefore, most NPMTs

have apparently been ‘‘tailored to function’’ in a clade- and

sometimes species-specific manner. Abundant evidence supports

the rapid expansion of NPMT gene families in several lineages,

due to genomic duplication events.10,20,74 Concomitant or subse-

quent gene fusions probably added additional levels of substrate

recognition, at its most extreme in the emergence of entirely new

domains. Neofunctionalization following duplication ultimately

led to large families of MTs involved in many different biological

processes.20,34As such, we now find that most NPMTs participate

in so-called specializedmetabolic pathways, i.e. biosynthetic grids

restricted to certain taxa or even to individual species.

thereby promoting methyl transfer. The divalent metal ion, in this case

Ca2+, is shown as a grey sphere. The substrate caffeoyl-CoA (with its 3-O-

methyl group removed) is colored magenta, and SAH is yellow. The blue

circle labeled with a ‘W’ represents a vicinal water molecule that abstracts

the proton from the methyl acceptor atom.

This journal is ª The Royal Society of Chemistry 2012

Page 11: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

Aclacimycin 10-hydroxylase (RdmB) represents an intriguing

example of neofunctionalization of the Class I MT fold.37,75

While indications from overall structure, SAM dependence, and

similarity to DnrK80 point to an MT function for RdmB, SAM

co-substrate binding in RdmB is atypical, such that methyl

transfer cannot occur. RdmB has instead evolved an unexpected

function as a hydroxylase.37 This example emphasizes the

importance of empirical functional characterization of enzymes,

and suggests that there are likely other mis-annotations of genes

based on homology and overall structural similarity. It will be

interesting to learn of additional instances where the Class I MT

scaffold has acquired or extended its functionality to catalyze

reactions other than methylation.

Moreover, in termsof abroader viewof themolecular evolution

of the NPMTs, one can ask, does the wide diversity of sequences,

appendages to their core SAM-binding fold, and their associated

substrate specificities/promiscuities indicate that these enzymes

are more structurally and catalytically malleable and therefore

readily evolvable? The exploration of this notion will require that

we progress from single enzyme stamp collecting to more

comprehensive studies of NPMT structure-function relationships

using an exponentially expanding database of sequences, folds

and activities, tied together by consideration of the evolutionary

mechanisms and selective pressures accompanying the amazing

adaptability of the SAM-binding systems in biocatalysis.

9 Engineering of NPMTs

The diverse substrate repertoires of the NPMTs, which never-

theless share a conserved protein fold, suggests that the NPMTs

might serve as useful starting points for protein engineering to

generate new biocatalysts. However, although the potential

malleability of NPMTs is often discussed in the literature, there

are relatively few reports of the successful rational engineering of

NPMTs or the use of these enzymes in the context of metabolic

engineering. This is partly due to the difficulty associated with

SAM regeneration in a metabolic engineering context, but also,

because we have yet to go beyond a narrow focus on individual

steps in specific metabolic pathways.

The NPMTs of plant phenylpropanoid metabolism and of the

SABATH family provide a number of examples where relatively

few amino-acid substitutions in the active site cause a shift in the

substrate specificity and/or regiospecificity of methylation,81 or

an expansion of the accepted substrate range.7,77,82,83 The relaxed

substrate selectivity (or ‘promiscuity’) of some NPMTs also

makes them attractive targets for metabolic engineering.50,84,85

Regiospecific methylation in organic synthesis can be a difficult

task. Such reactions often require elaborate schemes for the

addition of protecting groups and can suffer from low yields. The

design of engineered MTs to carry out these reactions could

improve the efficiency of bioactive molecule production through

semi-synthesis or recombinant production platforms albeit

done in concert with the development of cost effective SAM

regenerating systems.

10 Conclusions and future perspectives

Unquestionably, the structural elucidation of NPMT enzymes

has significantly advanced our understanding of how these

This journal is ª The Royal Society of Chemistry 2012

enzymes function in a biological context and has provided

considerable insight into their molecular evolution. The dozens

of structure determinations of enzymes derived from all three

domains of life provide what is ostensibly a comprehensive

representation of the structural diversity of NPMTs. However, a

number of NPMT lineages have yet to be explored on the

structural level, including sterol MTs, g-tocopherol MT-related

enzymes, and the ubiE/COQ5 family. Filling in the remaining

gaps in our structure-function landscape of the NPMT super-

family will further our understanding of the structural determi-

nants governing substrate recognition and catalysis, and foster

the synthesis of a universal theory of the evolution of NPMT

structure and function.

11 Acknowledgements

J.P.N. is a Howard HughesMedical Investigator. Research in the

Noel Laboratory is supported by the Howard Hughes Medical

Institute and the National Science Foundation (MCB-0645794,

MCB-0718064 and EEC-0813570). D.K.L. is a Natural Sciences

and Engineering Research Council of Canada (NSERC) Post-

doctoral Fellow.

12 References

1 C.-C. Huang, C. V. Smith, M. S. Glickman, W. R. Jacobs andJ. C. Sacchettini, J. Biol. Chem., 2002, 277, 11559–11569.

2 S. G. Lee, Y. Kim, T. D. Alpert, A. Nagata and J. M. Jez, J. Biol.Chem., 2012, 287, 1426–1434.

3 H. L. Schubert, K. S. Wilson, E. Raux, S. C. Woodcock andM. J. Warren, Nat. Struct. Biol., 1998, 5, 585–592.

4 M. E. Stroupe, H. K. Leech, D. S. Daniels, M. J. Warren andE. D. Getzoff, Nat. Struct. Biol., 2003, 10, 1064–1073.

5 C. Zubieta, X. Z. He, R. A. Dixon and J. P. Noel, Nat. Struct. Biol.,2001, 8, 271–279.

6 Z. Cheng, S. Sattler, H. Maeda, Y. Sakuragi, D. A. Bryant andD. DellaPenna, Plant Cell, 2003, 15, 2343–2356.

7 C. Zubieta, J. R. Ross, P. Koscheski, Y. Yang, E. Pichersky andJ. P. Noel, Plant Cell, 2003, 15, 1704–1716.

8 J. L. Martin, J. Begun, M. J. McLeish, J. M. Caine andG. L. Grunewald, Structure, 2001, 9, 977–985.

9 C.-J. Liu, B. E. Deavours, S. B. Richard, J.-L. Ferrer, J. W. Blount,D. Huhman, R. A. Dixon and J. P. Noel, Plant Cell, 2006, 18,3656–3669.

10 D. K. Liscombe, A. R. Usera and S. E. O’Connor, Proc. Natl. Acad.Sci. U. S. A., 2010, 107, 18793–18798.

11 D. K. Liscombe and P. J. Facchini, Curr. Opin. Biotechnol., 2008, 19,173–180.

12 J. Vidgren, L. A. Svensson and A. Liljas, Nature, 1994, 368, 354–358.13 J. L. Martin and F. M. McMillan, Curr. Opin. Struct. Biol., 2002, 12,

783–793.14 P. Z. Kozbial and A. R. Mushegian, BMC Struct. Biol., 2005, 5, 19.15 H. L. Schubert, R. M. Blumenthal and X. Cheng, Trends Biochem.

Sci., 2003, 28, 329–335.16 H. Coiner, G. Schr€oder, E. Wehinger, C.-J. Liu, J. P. Noel,

W. Schwab and J. Schr€oder, Plant J., 2006, 46, 193–205.17 C. Lomax, W.-J. Liu, L. Wu, K. Xue, J. Xiong, J. Zhou,

S. P. McGrath, A. A. Meharg, A. J. Miller and F.-J. Zhao, NewPhytol., 2012, 193, 665–672.

18 Y. Peng, Q. Feng, D. Wilk, A. A. Adjei, O. E. Salavaggione,R. M. Weinshilboum and V. C. Yee, Biochemistry, 2008, 47, 6216–6225.

19 J. W. Schmidberger, A. B. James, R. Edwards, J. H. Naismith andD. O’Hagan, Angew. Chem., Int. Ed., 2010, 49, 3646–3648.

20 A. Barakat, A. Choi, N. B. M. Yassin, J. S. Park, Z. Sun andJ. E. Carlson, Gene, 2011, 479, 37–46.

21 T. Wlodarski, J. Kutner, J. Towpik, L. Knizewski, L. Rychlewski,A. Kudlicki, M. Rowicka, A. Dziembowski and K. Ginalski, PLoSOne, 2011, 6, e23168.

Nat. Prod. Rep.

Page 12: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

22 E. Lissina, B. Young, M. L. Urbanus, X. L. Guan, J. Lowenson,S. Hoon, A. Baryshnikova, I. Riezman, M. Michaut, H. Riezman,L. E. Cowen, M. R. Wenk, S. G. Clarke, G. Giaever andC. Nislow, PLoS Genet., 2011, 7, e1002332.

23 G. V. Louie, M. E. Bowman, Y. Tu, A. Mouradov, G. Spangenbergand J. P. Noel, Plant Cell, 2010, 22, 4114–4127.

24 J. G. Kopycki, M. T. Stubbs, W. Brandt, M. Hagemann, A. Porzel,J. Schmidt, W. Schliemann, M. H. Zenk and T. Vogt, J. Biol.Chem., 2008, 283, 20888–20896.

25 J. G. Kopycki, D. Rauh, A. A. Chumanevich, P. Neumann, T. Vogtand M. T. Stubbs, J. Mol. Biol., 2008, 378, 154–164.

26 A. A. McCarthy and J. G. McCarthy, Plant Physiol., 2007, 144, 879–889.

27 A. E. Carney and H. M. Holden, Biochemistry, 2011, 50, 780–787.28 E. S. Burgie and H. M. Holden, Biochemistry, 2008, 47, 3982–

3988.29 Y. Nagatoshi and T. Nakamura, J. Biol. Chem., 2009, 284, 19301–

19309.30 H. Toda and N. Itoh, Phytochemistry, 2011, 72, 337–343.31 J. M. Attieh, A. D. Hanson and H. S. Saini, J. Biol. Chem., 1995, 270,

9250–9257.32 R. M. Kagan and S. Clarke, Arch. Biochem. Biophys., 1994, 310, 417–

427.33 D. Ingrosso, A. V. Fowler, J. Bleibaum and S. Clarke, J. Biol. Chem.,

1989, 264, 20131–20139.34 G. Schluckebier, M. O’Gara, W. Saenger and X. Cheng, J. Mol. Biol.,

1995, 247, 16–20.35 C. P. Joshi and V. L. Chiang, Plant Mol. Biol., 1998, 37, 663–674.36 J. Posfai, A. S. Bhagwat, G. P�osfai and R. J. Roberts, Nucleic Acids

Res., 1989, 17, 2421–2435.37 A. Jansson, H. Koskiniemi, A. Erola, J. Wang, P. M€ants€al€a,

G. Schneider and J. Niemi, J. Biol. Chem., 2005, 280, 3636–3644.38 M. Ansari, J. Sharma, R. S. Gokhale and D. Mohanty, BMC

Bioinformatics, 2008, 9, 454.39 M. L. Nuccio, M. J. Ziemak, S. A. Henry, E. A. Weretilnyk and

A. D. Hanson, J. Biol. Chem., 2000, 275, 14095–14101.40 B. Dumas, J. Van Doorsselaere, J. Gielen, M. Legrand, B. Fritig,

M. Van Montagu and D. Inz�e, Plant Physiol., 1992, 98, 796–797.41 S. Frick and T. M. Kutchan, Plant J., 1999, 17, 329–339.42 G. Schr€oder, E. Wehinger and J. Schr€oder, Phytochemistry, 2002, 59,

1–8.43 D. K. Liscombe and P. J. Facchini, J. Biol. Chem., 2007, 282, 14741–

14751.44 D. K. Liscombe, J. Ziegler, J. Schmidt, C. Ammer and P. J. Facchini,

Plant J., 2009, 60, 729–743.45 J. E. Katz, M. Dlaki�c and S. Clarke, Mol. Cell. Proteomics, 2003, 2,

525–540.46 T. C. Petrossian and S. G. Clarke, Mol. Cell. Proteomics, 2009, 8,

1516–1526.47 T. C. Petrossian and S. G. Clarke, Mol. Cell. Proteomics, 2011, 10,

M110.000976.48 X. Cheng, S. Kumar, J. Posfai, J. W. Pflugrath and R. J. Roberts,

Cell, 1993, 74, 299–307.49 M. G. Rossmann, D.Moras andK.W. Olsen,Nature, 1974, 250, 194–

199.50 D. L. Akey, S. Li, J. R. Konwerski, L. A. Confer, S. M. Bernard,

Y. Anzai, F. Kato, D. H. Sherman and J. L. Smith, J. Mol. Biol.,2011, 413, 438–450.

51 J.-L. Ferrer, C. Zubieta, R. A. Dixon and J. P. Noel, Plant Physiol.,2005, 137, 1009–1017.

52 N. A. Bruender, J. B. Thoden, M. Kaur, M. K. Avey andH. M. Holden, Biochemistry, 2010, 49, 5891–5898.

53 K. Wada, H. Yamaguchi, J. Harada, K. Niimi, S. Osumi, Y. Saga,H. Oh-Oka, H. Tamiaki and K. Fukuyama, J. Mol. Biol., 2006,360, 839–849.

54 J. P. Keller, P. M. Smith, J. Benach, D. Christendat, G. T. deTitta andJ. F. Hunt, Structure, 2002, 10, 1475–1487.

55 K. Wada, J. Harada, Y. Yaeda, H. Tamiaki, H. Oh-Oka andK. Fukuyama, FEBS J., 2007, 274, 563–573.

56 J. V�evodov�a, R. M. Graham, E. Raux, H. L. Schubert, D. I. Roper,A. A. Brindley, A. Ian Scott, C. A. Roessner, N. P. J. Stamford,M. Elizabeth Stroupe, E. D. Getzoff, M. J. Warren andK. S. Wilson, J. Mol. Biol., 2004, 344, 419–433.

57 S. Frank, E. Deery, A. A. Brindley, H. K. Leech, A. Lawrence,P. Heathcote, H. L. Schubert, K. Brocklehurst, S. E. J. Rigby,

Nat. Prod. Rep.

M. J. Warren and R. W. Pickersgill, J. Biol. Chem., 2007, 282,23957–23969.

58 S. Storbeck, S. Saha, J. Krausze, B. U. Klink, D. W. Heinz andG. Layer, J. Biol. Chem., 2011, 286, 26754–26767.

59 T. Lenz, P. Poot, E. Weinhold and M. Dreger, Methods Mol. Biol.,2012, 803, 97–125.

60 C. Dalhoff, M. H€uben, T. Lenz, P. Poot, E. Nordhoff, H. K€oster andE. Weinhold, ChemBioChem, 2010, 11, 256–265.

61 L. Wirsing, K. Naumann and T. Vogt, Anal. Biochem., 2011, 408,220–225.

62 C. S. Pugh, R. T. Borchardt and H. O. Stone, J. Biol. Chem., 1978,253, 4075–4077.

63 R. T. Borchardt, L. E. Eiden, B. Wu and C. O. Rutledge, Biochem.Biophys. Res. Commun., 1979, 89, 919–924.

64 D. D. Smith and S. J. Norton,Biochem. Biophys. Res. Commun., 1980,94, 1458–1462.

65 M. Vedel, F. Lawrence, M. Robert-Gero and E. Lederer, Biochem.Biophys. Res. Commun., 1978, 85, 371–376.

66 M. T. McCammon and L. W. Parks, J. Bacteriol., 1981, 145, 106–112.67 C. L. Hendricks, J. R. Ross, E. Pichersky, J. P. Noel and Z. S. Zhou,

Anal. Biochem., 2004, 326, 100–105.68 S. Biastoff, M. Teuber, Z. Zhou and B. Dr€ager,PlantaMed., 2006, 72,

1136–1141.69 C. Dalhoff, G. Lukinavicius, S. Klimas�auskas and E. Weinhold, Nat.

Chem. Biol., 2006, 2, 31–32.70 B. W. K. Lee, H. G. Sun, T. Zang, B. J. Kim, J. F. Alfaro and

Z. S. Zhou, J. Am. Chem. Soc., 2010, 132, 3642–3643.71 Y. Luo, S. Lin, J. Zhang, H. A. Cooke, S. D. Bruner and B. Shen, J.

Biol. Chem., 2008, 283, 14694–14702.72 Z. Fu, Y. Hu, K. Konishi, Y. Takata, H. Ogawa, T. Gomi,

M. Fujioka and F. Takusagawa, Biochemistry, 1996, 35, 11985–11993.

73 Y. Peng, D. Sartini, V. Pozzi, D. Wilk, M. Emanuelli and V. C. Yee,Biochemistry, 2011, 50, 7800–7808.

74 N. Zhao, J.-L. Ferrer, J. Ross, J. Guan, Y. Yang, E. Pichersky,J. P. Noel and F. Chen, Plant Physiol., 2008, 146, 455–467.

75 A. Jansson, J. Niemi, Y. Lindqvist, P. M€ants€al€a and G. Schneider, J.Mol. Biol., 2003, 334, 269–280.

76 D. O’Hagan and J. W. Schmidberger, Nat. Prod. Rep., 2010, 27, 900.77 C. Zubieta, P. Kota, J.-L. Ferrer, R. A. Dixon and J. P. Noel, Plant

Cell, 2002, 14, 1265–1277.78 I. G�omez Garc�ıa, C. E. M. Stevenson, I. Us�on, C. L. Freel Meyers,

C. T. Walsh and D. M. Lawson, J. Mol. Biol., 2010, 395, 390–407.79 J.-H. Cho, Y. Park, J.-H. Ahn, Y. Lim and S. Rhee, J. Mol. Biol.,

2008, 382, 987–997.80 A. Jansson, H. Koskiniemi, P. M€ants€al€a, J. Niemi and G. Schneider,

J. Biol. Chem., 2004, 279, 41149–41156.81 J. Wang and E. Pichersky, Arch. Biochem. Biophys., 1999, 368, 172–

180.82 M.-W. Bhuiya and C.-J. Liu, J. Biol. Chem., 2010, 285, 277–285.83 S. Lee, S. Y. Shin, Y. Lee, Y. Park, B. G. Kim, J.-H. Ahn, Y. Chong,

Y. H. Lee and Y. Lim,Bioorg.Med. Chem. Lett., 2011, 21, 3866–3870.84 R. Shi, S. S. Lamb, B. Zakeri, A. Proteau, Q. Cui, T. Sulea, A. Matte,

G. D. Wright and M. Cygler, Chem. Biol., 2009, 16, 401–410.85 H. A. Cooke, E. L. Guenther, Y. Luo, B. Shen and S. D. Bruner,

Biochemistry, 2009, 48, 9590–9598.86 S. Singh, J. G. McCoy, C. Zhang, C. A. Bingman, G. N. Phillips and

J. S. Thorson, J. Biol. Chem., 2008, 283, 22628–22636.87 J.-H. Lee, B. Bae, M. Kuemin, B. T. Circello, W. W. Metcalf,

S. K. Nair and W. A. van der Donk, Proc. Natl. Acad. Sci.U. S. A., 2010, 107, 17557–17562.

88 A. Chang, S. Singh, C. A. Bingman, J. S. Thorson and G. N. Phillips,Acta Crystallogr., Sect. D: Biol. Crystallogr., 2011, 67, 197–203.

89 S. Singh, A. Chang, R. D. Goff, C. A. Bingman, S. Gr€uschow,D. H. Sherman, G. N. Phillips Jr. and J. S. Thorson, Proteins:Struct., Funct., Bioinf., 2011, 79, 2181–2188.

90 J. F. Parsons, B. T. Greenhagen, K. Shi, K. Calabrese, H. Robinsonand J. E. Ladner, Biochemistry, 2007, 46, 1821–1828.

91 O. Cakici, M. Sikorski, T. Stepkowski, G. Bujacz andM. Jaskolski, J.Mol. Biol., 2010, 404, 874–889.

92 O. Ariyawutthiphan, T. Ose, M. Tsuda, Y. Gao, M. Yao, A. Minami,H. Oikawa and I. Tanaka, Acta Crystallogr., Sect. F: Struct. Biol.Cryst. Commun., 2011, 67, 417–420.

93 X. Hou, Y. Wang, Z. Zhou, S. Bao, Y. Lin and W. Gong, J. Struct.Biol., 2007, 159, 523–528.

This journal is ª The Royal Society of Chemistry 2012

Page 13: Architectures, mechanisms and molecular evolution of natural product methyltransferases

Dow

nloa

ded

by T

okyo

Dai

gaku

on

04 S

epte

mbe

r 20

12Pu

blis

hed

on 0

1 A

ugus

t 201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2NP2

0029

E

View Online

94 J. R. Horton, K. Sawada, M. Nishibori, X. Zhang and X. Cheng,Structure, 2001, 9, 837–849.

95 J. Komoto, Y. Huang, Y. Takata, T. Yamada, K. Konishi,H. Ogawa, T. Gomi, M. Fujioka and F. Takusagawa, J. Mol. Biol.,2002, 320, 223–235.

This journal is ª The Royal Society of Chemistry 2012

96 J. Komoto, T. Yamada, Y. Takata, K. Konishi, H. Ogawa, T. Gomi,M. Fujioka and F. Takusagawa, Biochemistry, 2004, 43, 14385–14394.

97 P. Emsley, B. Lohkamp, W. G. Scott and K. Cowtan, ActaCrystallogr., Sect. D: Biol. Crystallogr., 2010, 66, 486–501.

Nat. Prod. Rep.