Exploring Fungal Polyketide C-Methylation through Combinatorial

Oct 23, 2018 - Exploring Fungal Polyketide C-Methylation through Combinatorial Domain Swaps. Philip A. Storm , Paramita Pal , Callie R. Huitt-Roehl , ...
0 downloads 0 Views 554KB Size
Subscriber access provided by Kaohsiung Medical University

Letter

Exploring fungal polyketide C-methylation through combinatorial domain swaps Philip A. Storm, Paramita Pal, Callie R. Huitt-Roehl, and Craig A. Townsend ACS Chem. Biol., Just Accepted Manuscript • DOI: 10.1021/acschembio.8b00429 • Publication Date (Web): 23 Oct 2018 Downloaded from http://pubs.acs.org on October 24, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

39x29mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 18

1 1

Title:

2

Exploring fungal polyketide C-methylation through combinatorial

3

domain swaps

4 5

Authors:

6

Philip A. Storm, Paramita Pal, Callie R. Huitt-Roehl, and Craig A. Townsend*

7

Department of Chemistry, Johns Hopkins University, Baltimore, Maryland, 21218, United States

8

*Correspondence: [email protected]

9

Supporting Information

10 11 12

Abstract: Polyketide C-methylation occurs during a programmed sequence of dozens of

13

reactions carried out by multidomain polyketide synthases (PKSs). Fungal PKSs

14

perform these reactions iteratively, where a domain may be exposed to and act upon

15

multiple enzyme-tethered intermediates during biosynthesis. We surveyed a collection

16

of C-methyltransferase (CMeT) domains from non-reducing fungal PKSs to gain insight

17

into how different methylation patterns are installed. Our in vitro results show that

18

control of methylation resides primarily with the CMeT, and CMeTs can intercept and

19

methylate intermediates from non-cognate non-reducing PKS domains. Furthermore,

20

the methylation pattern is likely imposed by a competition between methylation or

21

ketosynthase-catalyzed extension for each intermediate. Understanding site-specific

22

polyketide C-methylation may facilitate targeted C-C bond formation in engineered

23

biosynthetic pathways.

ACS Paragon Plus Environment

Page 3 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

2 24 25

Text: Polyketide synthases (PKSs) have been long-standing targets for engineering

26

biosynthetic pathways because of the inherent structural and stereochemical complexity

27

built into their products.1 PKSs tether growing intermediates to the enzyme by

28

thioesterification to the phosphopantetheine arm of a carrier protein that delivers the

29

intermediates to client domains for successive rounds of elongation and modification.

30

The core reaction—ketosynthase-catalyzed C-C bond formation by decarboxylative

31

Claisen condensation—extends the intermediate by two carbons each cycle. However,

32

PKSs occasionally implement a secondary C-C bond forming reaction through

33

regiospecific -methylation by an S-adenosylmethionine (SAM)-dependent C-

34

methyltransferase (CMeT) embedded in the PKS.

35

Type I highly-reducing PKSs (HR-PKSs) found in both marine cyanobacteria and

36

fungi, modular and iterative systems respectively, can methylate polyene -ketoacyl

37

substrates as in curacin and lovastatin biosynthesis, through use of a PKS-embedded

38

CMeT.2, 3 In both examples, multiple -ketoacyl species of different chain length could

39

be presented to the CMeT, but only a single regiospecific methylation is observed.

40

Some fungal non-reducing PKSs (NR-PKSs) also contain embedded CMeT domains

41

and are known to methylate linear substrates ranging from C4 to C14 prior to cyclization

42

by the product template domain, with many of these CMeTs acting multiple times during

43

iterative catalysis.4, 5 Tolerance for such different acceptor substrates across all PKS

44

CMeTs may reflect an inherently adaptable active site allowing for substrate flexibility.

45

Although structures of PKS CMeTs remain few in number, features common to

46

these CMeTs may be advantageous to future efforts to engineer site-specific C-C bond

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 18

3 47

formation. In Type I PKS CMeTs, there is a two-subdomain architecture comprised of

48

an N-terminal cap and C-terminal Rossmann fold, with the active site positioned in a

49

cleft at their interface, a diagnostic indicator that the CMeT is active.3, 4 The CurJ and

50

PksCT CMeT active sites are large and hydrophobic, even though some methylated

51

substrates are as small as acetoacetyl-holo-ACP. Although no acceptor substrates have

52

been observed crystallographically in CMeT active sites, only the ketoacyl moiety

53

appears to make specific contacts with binding pocket residues, the catalytic His-Glu

54

dyad.4 Therefore, the N-terminal subdomain and large acceptor binding pockets of PKS

55

CMeTs could modulate selectivity or activity towards non-native substrates.

56

A collection of differently regiospecific CMeTs could be added to an engineered

57

polyketide biosynthetic pathway to elicit a programmable pattern of methylation.

58

Previous work has leveraged the ability to deconstruct multienzymes into their

59

constituent domains as a way to explore NR-PKS activity. Chain length, cyclization, and

60

product release have been surveyed by domain deconstruction and combinatorial

61

domain swap reactions.6-8 We sought to expand this approach to study fungal

62

polyketide C-methylation.

63

In order to build a domain library for combinatorial swap experiments, we

64

surveyed the literature for fungal Group VI and Group VII NR-PKSs.5 Only those PKSs

65

that had a known product associated with them were included, to allow for comparison

66

of the native methylation pattern to the pattern seen with in vitro domain swap reactions.

67

We omitted PKSs that utilized complex starter units derived from upstream HR-PKSs or

68

FASs such as the asperfuranone PKSs AfoG and AfoE and the alkylisoquinoline

69

synthase PkiA.5, 9 The candidates selected were PksCT (Monascus purpureus)10, MpaC

ACS Paragon Plus Environment

Page 5 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

4 70

(Penicillium brevicompactum)11, DtbA (Aspergillus niger)12, as well as PkeA, AusA, and

71

PkbA (A. nidulans).5 Each of these PKSs utilizes an acetyl starter unit, produces tetra-

72

or pentaketides, and installs one to three C-methyl groups having differing

73

regiochemistry in the native product, and we reasoned that their similarity would

74

maximize the chances for successful non-cognate interactions (Fig. 1).

75 76 77 78 79 80 81 82

Figure 1: Fungal NR-PKSs selected for domain deconstruction and combinatorial swaps. Group VI and VII NR-PKSs share a common domain architecture (top), with Group VI terminating in a TE domain to release the polyketide as the carboxylic acid and Group VII terminating in a R domain to give an aldehyde. DtbA and PkbA contain tandem ACP domains. (SAT: starter unit-ACP transacylase, KS: ketosynthase, MAT: malonyl-CoA:ACP transacylase, PT: product template, ACP: acyl carrier protein, CMeT: C-methyltransferase, R: reductase, TE: thioesterase).

Using the Udwary-Merski algorithm in combination with the early experience of

83

dissecting PksCT domains4, 13, we generated expression constructs for the SAT-KS-

84

MAT tridomain, and the CMeT and ACP mono-domains (Supplementary Information

85

Table 2). These three proteins would constitute the “minimal methylating PKS,” with all

86

necessary domains to carry out multiple rounds of extension and methylation. Based on

87

previously published PksCT results, we decided to omit PT and R or TE domains and

88

relied on spontaneous intramolecular C-O cleavage of the holo-ACP thioester to release

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 18

5 89

intermediates as pyrones, a commonly observed PKS derailment product.4 This

90

experimental design focused on the known interplay between extension and methylation

91

prior to PT-catalyzed cyclization or R/TE-catalyzed release. Non-methylating fungal NR-

92

PKSs quickly elongate substrates to the full chain-length8, but NR-PKSs with CMeT

93

domains have inherently slower extension activity to allow transit to the ACP-bound

94

intermediates to dock with the CMeT4. Not all desired fragments could be expressed or

95

were soluble, especially for the SAT-KS-MAT fragments; this outcome is consistent with

96

previous attempts to express SAT-KS-MATs across multiple fungal NR-PKS groups and

97

did not improve when alternate cut sites were used.8 Additionally, no fragments from

98

MpaC were soluble. Overall, ACP and CMeT monodomain fragments were generally

99

well behaved for the remaining PKSs in our library.

100

From the soluble, expressed fragments we were able to conduct exploratory

101

reactions to gauge the cross-compatibility of CMeTs with components from other NR-

102

PKSs. All apo-ACPs were activated with Sfp from Bacillus subtilis to give holo-ACP.14

103

PKS domain fragments were incubated with acetyl-SNAC, malonyl-SNAC, and SAM,

104

and reactions were extracted and analyzed by HPLC and UPLC-ESI-MS. Reconstitution

105

of the minimal PksCT (Figure 2A) was discussed in detail previously4, and the minimal

106

PkeA reaction demonstrated similar behavior (Figure 2B). The major product was the

107

unmethylated triketide 1, however the tetraketide 2 was also detected as a minor

108

product. Thus, the spontaneous release of unmethylated intermediates appears to be a

109

general trend for CMeT-containing NR-PKSs. We speculate that tetraketide 2 may be

110

favored in PkeA because the native substrate is only methylated once following the third

ACS Paragon Plus Environment

Page 7 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

6 111

extension, so longer unmethylated substrates must be better tolerated for productive

112

biosynthesis.

113 114 115 116 117 118

Figure 2: Reconstitution of the “minimal methylating PKS” for. A) PksCT minimal reconstitution with non-cognate ACPs. B) PkeA minimal reconstitution with non-cognate ACPs. C) Addition of non-cognate CMeTs to minimal PksCT D) Addition of non-cognate CMeTs to minimal PkeA E) PksCT and F) PkeA SAT-KS-MAT were combined with paired ACP and CMeT. Traces are vertically offset for clarity. See SI Tables 3 and 4 for UV/Vis and mass data.

119

Subsequently, we swapped ACPs to these minimal PKS reactions to determine

120

the compatability of non-cognate ACPs to participate in catalysis. As observed for non-

121

methylating NR-PKSs, ACPs are generally interchangeable, though the yield of pyrone

122

products varied.8 In non-methylating NR-PKSs, dissection generally reduces yield, but

123

the differences in product profiles seen in the above ACP swaps are consistent with

124

previous work.8, 15

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 18

7 125

We then added CMeT monodomains to PksCT or PkeA minimal PKSs. From the

126

observed pyrones, several aspects of C-methylation were evident. Non-cognate CMeTs

127

generally are capable of intercepting acyl-holo-ACP species and reduce the yield of

128

unmethylated polyketides. Because these pyrones have very similar UV-Vis absorption

129

spectra, we assume peak areas can be roughly correlated. In PksCT-based reactions

130

(Figure 2C), the triketides 3, 4, and 5 accumulate when non-cognate CMeTs are added

131

and are significantly less abundant than 1 in the absence of a CMeT. The position of the

132

methyl group for the singly methylated triketides 3 and 4 has been determined by

133

comparison to an authentic standard of 3.

134

In PkeA-based reactions (Figure 2D), non-cognate CMeTs resulted in the same

135

triketides and additional tetraketides. Notably, the addition of PkeA CMeT to the

136

cognate PkeA SAT-KS-MAT + ACP produced the singly methylated tetraketide 6,

137

consistent with the native PkeA methylation pattern. Because mass fragmentation does

138

not unambiguously reveal the methylation pattern in all cases, we synthesized a series

139

of singly and doubly methylated tetraketides to identify the positions of methyl groups on

140

the detected pyrones. By comparison to these synthetic standards, the position of

141

methyl groups on tetraketides 6 and 7 could be definitively assigned (see Fig. 3). AusA,

142

DtbA, PkbA, and PksCT CMeTs again produced the singly and doubly methylated

143

triketides 3, 4, and 5, although the reduction of 1 seen in Fig. 2C was not as

144

pronounced. This finding could reflect a higher rate of extension by the PkeA KS during

145

the first and second rounds, limiting the chance for methylation of those intermediates.

146

The presence of singly methylated triketide 3 and doubly methylated tetraketide 7

147

indicate that PksCT CMeT could not always intercept acetoacetyl-holo-ACP following

ACS Paragon Plus Environment

Page 9 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

8 148

the first extension cycle. The KS domain of PksCT was previously shown to be sensitive

149

to the absence of methylation on the growing intermediate8. Because the native product

150

of PkeA is not methylated following the first two extensions, the KS may be sensitive to

151

the presence of methylation on early stage intermediates, resulting in low overall yield.

152

A second set of reactions was also carried out, analogous to those above, but

153

the ACP and CMeT were cognate and paired with a non-cognate SAT-KS-MAT (Figure

154

2E and 2F). In most reactions, the observed products were consistent whether the ACP

155

was cognate to the SAT-KS-MAT or CMeT, with some exceptions. For PksCT SAT-KS-

156

MAT with paired PkbA ACP and CMeT, triketides 4 and 5 were observed in addition to

157

triketides 1 and 3 seen in Figure 2C. These products could reflect preferential binding to

158

the CMeT resulting in methylation. Paired PkeA ACP and CMeT were less effective at

159

producing methylated pyrones with PksCT SAT-KS-MAT than when the ACP was

160

paired with the tridomain, perhaps because PkeA CMeT does not natively methylate

161

early acyl intermediates, and spontaneous release at the triketide stage is faster than

162

the methylation by PkeA CMeT. PksCT ACP and CMeT generated the late-stage tetra-

163

and pentaketide pyrones 8 and 9, as seen previously.4 O

O R1

O

R2

164 165 166 167 168 169

7,

O

10 HO

HO 1: 3: 4: 5:

6,

O O

O HO

12

O

O O

O

O

HO

HO 13

O

O

14

15

R1 =H, R 2=H R1 =CH 3, R 2=H R1 =H, R 2=CH 3 R1 =CH 3, R 2=CH 3

Figure 4: Synthetic standards of potential methylation products. Triacetic acid lactone (1) and variously methylated derivatives 3–5 were obtained by published procedures16, 17 or from commercial sources. These were further elaborated to tetraketides 6, 7, 10, 12–15 as detailed in the Supplementary Information.

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 18

9 170

In total, this series of observations reflects the in trans nature of acyl-ACP:client

171

domain interactions in domain-dissected reconstitution reactions. The growing acyl-ACP

172

intermediates must leave the KS active site, travel through solvent, and dock within the

173

CMeT active site. That the CMeT lowers the rate of extension so significantly, suggests

174

that it strongly competes with the KS for binding of acyl-ACP intermediates. In non-

175

methylating NR-PKSs, extension is very rapid and the growing poly -ketone

176

intermediate likely remains shielded in the KS, as almost no products of intermediate

177

chain length are detected.18 The addition of a CMeT obviously slows extension as the

178

population of intermediate acyl-ACP species is now split among one additional

179

competing binding partner.

180

The methylation program for each CMeT does appear to be intrinsic, even in the

181

context of domain dissection. The non-cognate CMeTs methylate substrates with less

182

fidelity than cognate systems, with a bias towards the native methylation pattern of the

183

CMeT. For example, most of the non-cognate CMeTs added to minimal PksCT did not

184

universally eliminate production of the unmethylated triketide 1, even though the

185

cognate PksCT CMeT completely eliminates this product.4 Without more domain

186

fragments, especially SAT-KS-MATs, it is unclear how sensitive KS domains might be

187

to alternate methylation patterns. The absence of improperly methylated intermediates

188

in PksCT reconstitution suggests high fidelity to the “programmed” methylation, but not

189

all of the above reaction outcomes are equally informative. The differing processivity of

190

extension between PksCT and PkeA SAT-KS-MAT domains in the absence of CMeTs

191

supports KS-mediated stabilization of intermediates in accord with the observed

192

methylation programming. For example, if methylating acetoacetyl-ACP after the first

ACS Paragon Plus Environment

Page 11 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

10 193

Claisen condensation renders the intermediate an unfavorable substrate for continued

194

cycles in PkeA, available free holo-ACP will decrease as the diketide is less susceptible

195

to release as no pyrone formation is possible before the triketide is formed.

196

These experiments also lack the PT and R or TE domains. Little is known about

197

how these domains are affected by the substrate methylation pattern. Sensitivity of the

198

KS and PT to the stereochemistry of methyl groups, is also unknown. Domain

199

dissection certainly increases the likelihood of racemization of a given -methyl position

200

as the acyl-ACP docks to other domains in trans, a solvent-exposed process. The

201

interdependence of catalytic activity and substrate affinities (including ACP:client

202

domain binding) in iterative PKS domains is linked by co-evolution and will likely remain

203

an unresolved problem in the field until multidomain structures are available.

204

The molecular basis of programmed methylation patterns is uncertain, but

205

attempts have been made to functionally account for how iterative catalysis operates.

206

Studies on the fungal HR-PKS LovB, from lovastatin biosynthesis, recently probed how

207

acyl-ACP intermediates are appropriately directed towards the next catalytic domain in

208

the programmed sequence.19 A diene triketide is -methylated by a CMeT prior to

209

ketoreductase (KR)-catalyzed reduction of the -ketone to the corresponding alcohol.

210

Because the -position is most acidic before -reduction, therefore most available for

211

methylation, the CMeT must act first. Like the NR-PKSs discussed above, LovB

212

appears to be sensitive to substrates that are not methylated; in the absence of SAM,

213

intermediates are spontaneously released prematurely as pyrones.2 By providing

214

synthetic substrates, the authors demonstrated that the CMeT has high catalytic

215

efficiency towards only the on-path diene triketide, whereas the KR is less selective and

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 18

11 216

less efficient. The proper order of methylation and reduction is functionally enforced by

217

C-methylation being very fast for only a single intermediate while ketoreduction is

218

slower but acts on a wider range of substrates, consistent with only a single methylation

219

but multiple ketoreductions during synthesis.

220

We wanted to probe whether a similar process might be followed in methylating

221

NR-PKSs by balancing rates of extension and methylation. A panel of specific synthetic

222

intermediates could not be screened due to the reactivity of poly--ketone

223

intermediates, so therefore the relative amount of CMeT was used to modulate the rate

224

of methylation (Figure 4).

225 226 227 228 229 230

Figure 4: Increasing relative amounts of CMeT results in hypermethylation of pentaketide

231

When even very small amounts of CMeT were added to the reaction, the

232

intermediates. A) PksCT SAT-KS-MAT and ACP were incubated with increasing concentrations of PksCT CMeT. Traces are vertically stacked and the peak for 1 is truncated for clarity. B) Control of Cmethylation programming is proposed to rely on the rates of extension and methylation for each acyl-ACP intermediate produced.

production of triketide lactone 1 was dramatically reduced, and as the relative

ACS Paragon Plus Environment

Page 13 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

12 233

concentration of CMeT increased the total number of singly methylated pyrones 3, 4,

234

and 10 increased. The methyl position of 10 was determined by comparison to our

235

synthesized standards (see SI for details). At approximately equimolar CMeT, multiply

236

methylated pyrones 5, 8, and 9 dominated the reaction profile. And with excess CMeT,

237

a new pair of product peaks appeared, proposed to be the hypermethylated pentaketide

238

11. Because 11 elutes as two peaks with identical mass, fragmentation, and

239

chromophores, just like on-path pentaketide 9, it is likely to retain the stereocenters that

240

give rise to diastereomers—demonstrating non-native methylation following the final

241

extension.

242

We propose that the kinetic balance of extension and methylation determines the

243

observed methylation pattern. As with LovB, different domains may have catalytic

244

efficiency towards different substrates such that only the programmed outcome is

245

kinetically favorable. At low concentrations of CMeT, not all programmed methylations

246

can occur before the KS performs another extension generating an off-path

247

intermediate. At high concentration, the selectivity against positions that are not part of

248

the “methylation program” can be overcome by increasing the overall rate of

249

methylation. These experiments cannot replicate the kinetic advantage found in an

250

intact NR-PKS, where the effective concentrations of acyl-ACP and client domains are

251

high. It is also possible that other conformational changes transmitted within an intact

252

NR-PKS could influence control of methylation. The primary determinant of the

253

methylation pattern, however, appears to be the CMeT domain itself, in addition to

254

surveillance by the KS.

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 18

13 255

In summary, we assembled a small library of CMeTs and expanded an in vitro

256

domain swap approach to explore the factors controlling polyketide C-methylation. We

257

found that NR-PKS CMeTs were capable of methylating acyl-ACP intermediates in

258

trans, and that the methylation program is primarily intrinsic to the CMeT. The product

259

profiles could be modulated based on the relative concentration of CMeT, even allowing

260

for non-native methylation. Together the data suggest that programmed methylation is a

261

kinetic balance between the rates of extension and methylation, although a molecular or

262

structural basis for differential rates remains to be established. A more comprehensive

263

collection of CMeT domains, development of chimeric intact NR-PKSs, and structures of

264

multidomain PKS fragments with bound substrates will offer further insight into this

265

complex multidomain choreography.

266 267 268

Methods All reagents were purchased from Sigma Aldrich except 4-hydroxy-3,6-dimethyl-

269

2H-pyran-2-one (3) which was purchased from Acros, AcSNAC, which was synthesized,

270

and MalSNAC, which was prepared enzymatically by MatB and purified.7, 8 Standard

271

molecular biology procedures were used for isolation of DNA and the creation of

272

expression plasmids. Sequencing was performed at the Johns Hopkins University

273

Synthesis and Sequencing Facility. M. purpureus NRRL 1596, A. niger NRRL 328, A.

274

nidulans NRRL 194, and P. brevicompactum NRRL 2011 were received from USDA

275

ARS and cultured on PDA plates. Collected mycelia were flash frozen, ground under

276

N2(l) and gDNA was purified using the DNeasy Plant Mini kit (Qiagen). Overlap

277

extension polymerase chain reaction was used to amplify exons and assemble intron-

ACS Paragon Plus Environment

Page 15 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

14 278

free expression inserts, which were ligated into pET-24a or pET-28a (Novagen) with N-

279

and/or C-terminal His6 tags.

280

Details of expression constructs and domain boundaries can be found in the

281

Supplemental Information. Plasmids were transformed to E. coli BL21(DE3) by

282

electroporation and maintained on LB agar plates. Expression cultures were grown in

283

LB media at 37 C to OD600 of 0.6. Cultures were cold-shocked in an ice bath for about

284

1 h, and induced with 1 mM IPTG (GoldBio) overnight at 18 C, shaking at 225 rpm. Cell

285

pellets were harvested by centrifugation for 15 min, at 4,000 x g, 4 C and either flash

286

frozen in N2(l) and stored at -80 C until use, or resuspended immediately in 5 mL g-1

287

cell pellet in lysis buffer (50 mM potassium phosphate, 300 mM NaCl, 10 % (v/v)

288

glycerol, pH 7.6). Resuspended cells were lysed by sonication on ice for 10 x 10 s at

289

40% amplitude (Vibra-Cell Ultrasonic Processor, Sonics & Materials, Inc.). The lysate

290

was cleared by centrifugation for 25 min at 27,000 x g, 4 C and batch bound to Co2+-

291

TALON (Clontech) at 4 C for 1 h. Expressed proteins were purified by gravity column

292

chromatography as follows: the protein-bound resin was washed with lysis buffer (10

293

CV), followed by lysis buffer + 2 mM imidazole (5 CV) and elution with lysis buffer + 100

294

mM imidazole (5 CV). Domain fragments were dialyzed into reaction buffer (100 mM

295

potassium phosphate, 5 % (v/v) glycerol, pH 7.0), concentrated by diafiltration, and

296

used immediately or flash frozen in N2(l) and stored at -80 °C.

297

Protein concentration was determined by Bradford assay (BioRad, Hercules, CA)

298

in duplicate using bovine serum albumin as a standard. Prior to in vitro reactions, ACPs

299

(50 μM) were activated by Sfp (1 μM) with CoASH (100 μM) and MgCl2 (10 mM) in

300

reaction buffer for 1 h at room temperature as previously reported.8 Reconstitution

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 18

15 301

reactions contained 10 μM of each included domain with 0.5 mM AcSNAC, 2 mM

302

MalSNAC, 2 mM SAM, and 1 mM TCEP in reaction buffer totaling 250 μL. After 4 h at

303

room temperature, the reactions were quenched with 5 μL concentrated HCl and

304

extracted into ethyl acetate 3 x 250 μL. The combined organic extracts were dried to a

305

residue and resuspended in 250 μL of 20 % (v/v) aqueous acetonitrile. Extracts were

306

analyzed on an Agilent 1200 HPLC with autosampler by injecting 100 μL onto a Prodigy

307

ODS3 column (4.5 x 250 mm, 5μ, Phenomenex, Torrence, CA) with 5-85 % (v/v)

308

MeCN/H2O, with 0.1 % (v/v) formic acid, over 40 min at 1 mL min-1 and monitored at

309

280 nm. Mass spectrometric analysis was done using a Waters Acquity Xevo G-2

310

UPLC-ESI-MS in positive ion mode, with 5 μL injected on a BEHC C18 column and 10-

311

90 % (v/v) MeCN/H2O, with 0.1 % (v/v) formic acid, over 10 min.

312 313

• ASSOCIATED CONTENT

314

Supporting Information

315

Details of cloning, characterization of enzymatic products, syntheses of standards to

316

identify tri- and tetraketides, associated structural characterizations, comparisons and

317

1H-

318

available free of charge via the internet at http://pubs.acs.org.

319

• AUTHOR INFORMATION

320

Corresponding Author

321

*Email: [email protected]

322

Notes

323

The authors declare no competing financial interest.

and 13C-NMR spectra are provided in the Supporting Information. This information is

ACS Paragon Plus Environment

Page 17 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

16 324

• ACKNOWLEDGEMENTS

325

This work was supported by National Institutes of Health grants RO1 ES001670 and

326

T32 GM080189. We also thank I. P. Mortimer (JHU) for UPLC-MS support.

327

• REFERENCES

328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363

1. Staunton, J.; Weissman, K. J., Polyketide biosynthesis: a millennium review. Nat. Prod. Rep. 2001, 18, 380-416. 2. Ma, S. M.; Li, J. W.; Choi, J. W.; Zhou, H.; Lee, K. K.; Moorthie, V. A.; Xie, X.; Kealey, J. T.; Da Silva, N. A.; Vederas, J. C.; Tang, Y., Complete reconstitution of a highly reducing iterative polyketide synthase. Science 2009, 326, 589-592. 3. Skiba, M. A.; Sikkema, A. P.; Fiers, W. D.; Gerwick, W. H.; Sherman, D. H.; Aldrich, C. C.; Smith, J. L., Domain Organization and Active Site Architecture of a Polyketide Synthase C-methyltransferase. ACS Chem. Biol. 2016, 11, 3319-3327. 4. Storm, P. A.; Herbst, D. A.; Maier, T.; Townsend, C. A., Functional and Structural Analysis of Programmed C-Methylation in the Biosynthesis of the Fungal Polyketide Citrinin. Cell Chem. Biol. 2017, 24, 316-325. 5. Ahuja, M.; Chiang, Y. M.; Chang, S. L.; Praseuth, M. B.; Entwistle, R.; Sanchez, J. F.; Lo, H. C.; Yeh, H. H.; Oakley, B. R.; Wang, C. C., Illuminating the diversity of aromatic polyketide synthases in Aspergillus nidulans. J. Am. Chem. Soc. 2012, 134, 8212-8221. 6. Crawford, J. M.; Thomas, P. M.; Scheerer, J. R.; Vagstad, A. L.; Kelleher, N. L.; Townsend, C. A., Deconstruction of iterative multidomain polyketide synthase function. Science 2008, 320, 243-246. 7. Vagstad, A. L.; Newman, A. G.; Storm, P. A.; Belecki, K.; Crawford, J. M.; Townsend, C. A., Combinatorial Domain Swaps Provide Insights into the Rules of Fungal Polyketide Synthase Programming and the Rational Synthesis of Non-Native Aromatic Products. Angew. Chem. Int. Ed. Engl. 2013, 52, 1718-1721. 8. Newman, A. G.; Vagstad, A. L.; Storm, P. A.; Townsend, C. A., Systematic domain swaps of iterative, nonreducing polyketide synthases provide a mechanistic understanding and rationale for catalytic reprogramming. J. Am. Chem. Soc. 2014, 136, 7348-7362. 9. Chiang, Y. M.; Szewczyk, E.; Davidson, A. D.; Keller, N.; Oakley, B. R.; Wang, C. C., A gene cluster containing two fungal polyketide synthases encodes the biosynthetic pathway for a polyketide, asperfuranone, in Aspergillus nidulans. J. Am. Chem. Soc. 2009, 131, 2965-2970. 10. Shimizu, T.; Kinoshita, H.; Ishihara, S.; Sakai, K.; Nagai, S.; Nihira, T., Polyketide synthase gene responsible for citrinin biosynthesis in Monascus purpureus. Appl. Environ. Microbiol. 2005, 71, 3453-3457. 11. Regueira, T. B.; Kildegaard, K. R.; Hansen, B. G.; Mortensen, U. H.; Hertweck, C.; Nielsen, J., Molecular basis for mycophenolic acid biosynthesis in Penicillium brevicompactum. Appl. Environ. Microbiol. 2011, 77, 3035-3043.

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 18

17 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390

12. Yeh, H.; Chang, S. L.; Chiang, Y. M.; Bruno, K. S.; Oakley, B. R.; Wu, T. K.; Wang, C. C., Engineering Fungal Nonreducing Polyketide Synthase by Heterologous Expression and Domain Swapping. Org. Lett. 2013, 15, 756-759. 13. Udwary, D. W.; Merski, M.; Townsend, C. A., A method for prediction of the locations of linker regions within large multifunctional proteins, and application to a type I polyketide synthase. J. Mol. Biol. 2002, 323, 585-598. 14. Quadri, L. E.; Weinreb, P. H.; Lei, M.; Nakano, M. M.; Zuber, P.; Walsh, C. T., Characterization of Sfp, a Bacillus subtilis phosphopantetheinyl transferase for peptidyl carrier protein domains in peptide synthetases. Biochemistry 1998, 37, 1585-1595. 15. Huitt-Roehl, C. R.; Hill, E. A.; Adams, M. M.; Vagstad, A. L.; Li, J. W.; Townsend, C. A., Starter unit flexibility for engineered product synthesis by the nonreducing polyketide synthase PksA. ACS Chem. Biol. 2015, 10, 1443-1449. 16. Yokoe, H.; Mitsuhashi, C.; Matsuoka, Y.; Yoshimura, T.; Yoshida, M.; Shishido, K., Enantiocontrolled total syntheses of breviones A, B, and C. J. Am. Chem. Soc. 2011, 133, 8854-8857. 17. Leiris, S. J.; Khdour, O. M.; Segerman, Z. J.; Tsosie, K. S.; Chapuis, J. C.; Hecht, S. M., Synthesis and evaluation of verticipyrone analogues as mitochondrial complex I inhibitors. Bioorg. Med. Chem. 2010, 18, 3481-3493. 18. Vagstad, A. L.; Bumpus, S. B.; Belecki, K.; Kelleher, N. L.; Townsend, C. A., Interrogation of global active site occupancy of a fungal iterative polyketide synthase reveals strategies for maintaining biosynthetic fidelity. J. Am. Chem. Soc. 2012, 134, 6865-6877. 19. Cacho, R. A.; Thuss, J.; Xu, W.; Sanichar, R.; Gao, Z.; Nguyen, A.; Vederas, J. C.; Tang, Y., Understanding Programming of Fungal Iterative Polyketide Synthases: The Biochemical Basis for Regioselectivity by the Methyltransferase Domain in the Lovastatin Megasynthase. J. Am. Chem. Soc. 2015, 137, 15688-15691.

ACS Paragon Plus Environment