Class I methyltransferase VioH catalyzes unusual SAM cyclization

12 hours ago - Class I methyltransferase VioH catalyzes unusual SAM cyclization leading to 4-methylazetidinecarboxylic acid formation in vioprolide ...
0 downloads 0 Views 619KB Size
Subscriber access provided by AUSTRALIAN NATIONAL UNIV

Article

Class I methyltransferase VioH catalyzes unusual SAM cyclization leading to 4-methylazetidinecarboxylic acid formation in vioprolide biosynthesis Fu Yan, and Rolf Müller ACS Chem. Biol., Just Accepted Manuscript • DOI: 10.1021/acschembio.8b00958 • Publication Date (Web): 12 Dec 2018 Downloaded from http://pubs.acs.org on December 13, 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 15 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

3

Class I methyltransferase VioH catalyzes unusual SAM cyclization leading to 4-methylazetidinecarboxylic acid formation in vioprolide biosynthesis

4

Fu Yan, Rolf Müller*

5 6 7

Helmholtz Institute for Pharmaceutical Research Saarland (HIPS), Helmholtz Centre for Infection Research and Department of Pharmacy at Saarland University, Saarland University Campus, Building E8.1, 66123 Saarbrücken, Germany.

8

*E-mail: [email protected]

1 2

9 10

ABSTRACT

11 12 13 14 15 16 17 18 19 20

SAM-dependent methyltransferases are intensely studied since they play important roles in the methylation of biomolecules in all domains of life. In this study we describe that the methyltransferase VioH from Cysotobacter violaceus catalyzes a so far unknown cyclization of SAM to azetidine-2-carboxylic acid (AZE), which is proposed to be the precursor of the unusual 4-methylazetidinecarboxylic acid (MAZ) moiety of vioprolides. In vitro biochemical investigations reveal that SAM is converted to AZE in the presence of VioH, while MAZ is generated by coexpression of VioH and the radical SAM enzyme VioG in Myxococcus xanthus or by combination of VioH and the cell lysate of M. xanthus expressing VioG. Thus, our findings unveil a novel function of SAM-dependent methyltransferases and shed light on the biosynthetic mechanism of MAZ formation.

21

INTRODUCTION

22 23 24 25 26 27 28 29 30 31 32 33 34 35 36

S-adenosyl-L-methionine (SAM) dependent methyltransferases (SAM-MTs, EC 2.1.1) are omnipresent in all domains of life, performing critical roles in biosynthetic modification and functional regulation of biomolecules like nucleotides, proteins and natural products.(1–3) They usually transfer a methyl group from the ubiquitous cofactor SAM to accepting atoms such as O, C, N, S, P or even halides.(2) The versatility of SAM-MTs is also reflected by their broad variety of substrates. Up to date, more than 300 members of SAM-MTs have been classified in the Enzyme Classification system according to their substrate specificities, and over 1.8 million methyltransferase (MT) protein sequences have been deposited in the UniProt database. Based on the structural fold, MTs are currently divided into five classes, among which class I MTs represent the largest group.(3) MTs are also frequently involved in natural product biosynthetic pathways and methylation of secondary metabolites can significantly affect their bioactivities.(3) Elucidation of reaction mechanisms and rational engineering of MTs thus can contribute to the generation of novel natural product derivatives with modified bioactivities. Here, we report a novel function for a class I MT from the 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

37 38

vioprolide biosynthetic pathway, which features a previously undescribed SAM cyclization activity to yield azetidine-2-carboxylic acid (AZE).

39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62

Vioprolides are a class of anti-fungal and cytotoxic peptolides produced by the myxobacterium Cysotobacter violaceus Cb vi35,(4) the biosynthetic pathway of which has been elucidated recently.(5) Vioprolides A and C contain an unusual 4methylazetidinecarboxylic acid (MAZ) moiety (Figure 1). In addition to vioprolides, the presence of MAZ in a natural product has only been reported for bonnevillamide A, which is a linear heptapeptide isolated from Streptomyces sp. GSL-6B recently (Figure 1).(6) The origin of the MAZ moiety in bonnevillamide is unknown, whereas feeding studies in vioprolide biosynthesis revealed that MAZ originates from methionine.(5) In the biosynthesis pathway of vioprolides the class I SAM-MT-like protein VioH and the radical SAM protein VioG were identified by gene deletion as essential for MAZ formation. VioH was proposed to cyclize the α-aminobutyric carboxylic acid moiety of SAM to form AZE, while VioG was hypothesized to methylate the resulting AZE via a radical methylation mechanism.(5) Currently, only few natural compounds are known that contain an azetidine moiety (Figure 1).(6, 7–15, 16) In plants, methionine was speculated to be the source of the AZE.(17–19) The formation of AZE in nicotianamine, a metal chelator ubiquitously produced in plants, was found to depend on SAM cyclization catalyzed by nicotianamine synthase (NAS).(20) Sequence analysis revealed a MT-like domain in the C-terminus of NAS.(21) However, the role of the MT-like domain in NAS is unclear and the catalytic mechanism for AZE formation remains elusive. The azetidine ring in okamarine B and D, indole alkaloids from Penicillium and Aspergillus species, is formed by intramolecular cyclization catalyzed by an α-ketoglutaratedependent dioxygenase OkaE,(22) while the azetidine moiety in polyoxins, nucleoside antibiotics from streptomycetes, originates from isoleucine and seems to be formed by a different mechanism.(23,24)

63 ACS Paragon Plus Environment

Page 2 of 15

Page 3 of 15 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

64

ACS Chemical Biology

Figure 1 Structures of azetidine-containing natural products.

65

RESULTS AND DISCUSSION

66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91

To investigate the mechanism of MAZ formation in vioprolide biosynthesis, we aimed to characterize the enzymatic activities of VioG and VioH in vitro. VioH was heterologously produced in Escherichia coli and purified with a yield of 4.4 mg mL-1 (Supplementary Figure 1), while no soluble VioG could be obtained from E. coli under various conditions. Purification of His6-VioG from Myxococcus xanthus failed (data not shown). A homogeneous fraction of VioH was incubated with SAM, and the resulting products were modified by Marfey’s reagent Nα-(5-Fluoro-2,4-dinitrophenyl)-D-leucinamide (D-FDLA) and analyzed by UPLC-MS. As shown in Figure 2, a compound featuring the same mass and retention time as the AZE-DLA reference compound (preparation described in Exp. Procedures) was generated by incubation of VioH and SAM. No AZE was generated in presence of VioH which was denatured at 100 oC. Detailed analysis revealed that the catalytic efficiency of VioH was remarkably low. Incubation of 50 μM VioH with 2 mM SAM at 30 oC for 16 h showed continuous production of AZE (Supplementary Figure 2). Similar results were obtained by incubation of 50 μM VioH and 50 μM SAM at 30 oC for 24 h. However, incubation of VioH and SAM in equivalent concentrations at 40 oC reavealed the decrease of reaction speed after 8 hours (Supplementary Figure 3). It seems that SAM is more stable at 30 oC and was largely degraded at 40 oC after 8 hours. We then analyzed the effects of pH, temperature and buffer on the catalytic efficiency of VioH. The activity assay was performed in a pH range from 6.5 to 12.5 and showed that the formation of AZE reached its maximum at around pH 9 and significantly decreased at pH > 11.5 (Figure 2). However, the lower production of AZE at high pH may also result from the lack of stability of SAM. Temperature curve assays revealed a steady increase of AZE production from 25 oC to 45 oC and a decline at higher temperature (Figure 2). Even at temperature as high as 50 oC, VioH still retained activity. Unexpectedly, AZE-DLA was detected in the samples incubated at high temperatures. To clarify this finding, aliquots containing only buffer and SAM were incubated at 25–100 oC for 2 h. UPLC-MS analysis detected the formation of a compound exhibiting the same mass and rentention time with AZE-DLA (Supplementary Figure 4), indicating a chemical formation of AZE from SAM which, to the best of our knowledge, has not been described in the literature. The chemically formed AZE could be detected at very low levels after reactions below 40 oC, with an increase in the production at higher temperatures. UPLC-MS/MS analysis demonstrated the VioHindependent formation of AZE at increased temperatures, especially in the range of 70– 100 oC (Supplementary Figure 5). Considering the stability of SAM, the optimal catalytic temperature for VioH is between 40 and 45 oC. To avoid possibly interfering effects from buffer components, the storage and reaction buffer of VioH was changed from Tris-NaCl to PBS, in which VioH showed a slightly higher activity (Supplementary Figure 6). The formation of AZE at concentrations below those of VioH present in the assays indicated

92 93 94 95 96 97 98 99 100 101 102 103

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

104 105 106 107 108 109 110 111

a very low in vitro kcat. Due to the observed low catalytic efficiency of VioH and the issue with stability of SAM, we were not able to determine catalytic constants in vitro. Notably, AZE has been reported as a toxic non-proteinogenic amino acid, the incorporation of which can cause misfolding of proteins and inhibition of cell growth.(25–27) Hence, a low level of intracellular AZE production may protect C. violaceus from such side-effects during vioprolide biosynthesis. In addition, since VioG and VioH are proposed to work synergistically in the formation of MAZ,(5) we cannot exclude that VioG has a crucial impact on the catalytic activity of VioH, which we can currently not test in vitro.

112 113 114 115 116 117 118

Figure 2 Production analysis of AZE generated in vitro. (a) scheme of Marfey’s derivatization. (b) UPLC-MS analysis AZE-DLA. The reactions were performed at 30 oC for 1 h. Similar results were obtained by using SAM from a different vendor (Supplementary Figure 7). Extracted ion chromatograms of AZE-DLA (m/z 396.14 ± 0.05 [M+H]+) is shown. (c) pH dependence of VioH activity. (d) temperature dependence of VioH activity.

119 120 121 122 123 124 125 126

In addition to the detection of AZE, we also monitored the reaction products of SAM. As expected, the methylthioadenosine split from SAM was detected (Supplementary Figure 8). As previously reported, SAM is not stable in aqueous solution. At different pH conditions, SAM decomposes into 5’-deoxy-5’-(methylthio)adenine (MTA), adenine, S(5’-deoxy-ribosyl)-L-methionine, methionine, homoserine or homoserine lactone.(28–30) In our reactions, we detected all the hydrolytic products of SAM except for S-(5’-deoxyribosyl)-L-methionine (Supplementary Figure 8). To test if the formation of hydrolysis products from SAM mignt be catalyzed by VioH to form AZE, VioH was incubated with ACS Paragon Plus Environment

Page 4 of 15

Page 5 of 15 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

127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145

ACS Chemical Biology

MTA, methionine, homoserine or homoserine lactone. Subsequent UPLC-MS analysis revealed that none of these substrates could be cyclized by VioH (Supplementary Figure 9). Considering the structural similarities between SAM and its hydrolytic fragments, these degradation products from SAM may bind to VioH and inhibit the catalysis. Since SAM is degraded to methylthioadenosine naturally in aqueous solution, an SN2 reaction is proposed to take place at the C4 position of the resulting aminobutyric acid moiety, yielding homoserine, homoserine lactone or AZE (Figure 3). It is reasonable to assume that VioH deprotonates the amino group of SAM and facilitates the intramolecular cyclization between C4 and the amino group (Figure 3). Protein BLAST search against the protein databank uncovered few homologous proteins exhibiting conserved SAMbinding regions with less than 45% sequence identity to VioH, while very low sequence similarity was found among VioH and other structurally-known methyltransferases (Supplementary Figure 10). Although residues in the SAM-binding region are conserved, SAM-binding residues could vary tremendously.(1) Additionally, homology modelling of VioH is not accurate because of a lack of a good template protein structure. Thus it is difficult to predict the crucial catalytic residues in VioH based on in silico analysis. Initial attempts to (co-)crystallize VioH without and with substrate or product have not been successful so far and will be continued using a broader set of conditions in order to obtain detailed molecular insights into the catalytic mechanism.

146 147

Figure 3 Proposed mechanism of AZE formation.

148 149 150 151 152 153 154 155 156 157

Since soluble expression of VioG in E. coli was not achievable, we attempted to investigate the function of VioG and to produce MAZ in vivo. Three heterologous expression strains, Myxococcus xanthus::Ptet-vioG, M. xanthus::Ptet-vioH and M. xanthus::Ptet-vioGH, were generated producing either VioG, VioH, or both enzymes, respectively. After cultivation, the crude extracts of the cultures were modified by Marfey’s reagent and were analyzed by UPLC-MS. As is shown in Figure 4, MAZ could be detected in the crude extracts from M. xanthus::Ptet-vioGH in the presence of vitamin B12. Meanwhile, small amounts of AZE could also be detected. In addition, the mutant M. xanthus::Ptet-vioH produced some AZE, while neither AZE nor MAZ could be found from the mutant M. xanthus::Ptet-vioG. In our previous gene inactivation studies,

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

158 159 160

deletion of vioG in the vio gene cluster abolished the production of vioprolides A and C without generation of AZE-containing vioprolides,(5) which may result from the low catalytic efficiency of VioH in the absence of VioG.

161 162 163 164 165 166 167 168 169 170 171

Figure 4 UPLC-MS analysis of AZE-DLA and MAZ-DLA from M. xanthus mutants. Extracted ion chromatograms of AZE-DLA (m/z 396.14 ± 0.05 [M+H]+, 13.22 min), MAZDLA (m/z 410.20 ± 0.05 [M+H]+, 14.40 min) and D-Pro-DLA (m/z 410.20 ± 0.05 [M+H]+, 14.15 min) are shown. A, M. xanthus negative control; B, M. xanthus::Ptet-vioG; C, M. xanthus::Ptet-vioH; D, M. xanthus::Ptet-vioGH without feeding of vitamin B12; E, M. xanthus::Ptet-vioGH fed with vitamin B12; F, MAZ-DLA reference; G, AZE-DLA reference; H, D-Pro-DLA reference. The AZE-DLA peaks in D and E were magnified by 10 fold above the line of the chromatogram. The peaks of D-Pro-DLA are shown because of their isobaric mass to MAZ-DLA. The retention time of L-Pro-DLA is out of the time frame shown (retention time 15.84 min under these conditions).

172 173 174 175 176 177 178 179 180 181 182 183 184 185

After showing the formation of MAZ in the heterologous M. xanthus strains, we tested an alternative in vivo – in vitro combined approach to reconstitute MAZ production. The M. xanthus::Ptet-vioG cell lysates were incubated with purified recombinant VioH, and the products were modified by D-FDLA and analyzed by UPLC-MS. As shown in Figure 5, the formation of MAZ could only be detected in the presence of VioH in the cell lysate. Notably, in the absence of VioH, MAZ could not be generated by incubation of M. xanthus::Ptet-vioG cell lysates with AZE. It seems that free AZE cannot be methylated by VioG. It is likely that VioG can only accept modified or protein-bound AZE, hence an interaction of VioG and VioH might be essential for the generation of MAZ. In previous in vivo experiments, vitamin B12 was found essential for the formation of MAZ, and thus it was added to the cultivation medium of M. xanthus::Ptet-vioG. Supplementing vitamin B12 to cell lysates did not show an effect on the production of MAZ (Figure 5). Interestingly, the variation of SAM concentrations also showed no significant impact on the formation of MAZ. Even if no further SAM was added to the reaction, MAZ was still

ACS Paragon Plus Environment

Page 6 of 15

Page 7 of 15 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

186 187 188 189 190 191 192 193 194 195

ACS Chemical Biology

formed (Figure 5, D and F), indicating that the SAM present in the cell lysate seems to be sufficient for the observed product formation. The overall low production yield of MAZ in different reaction systems indicated limited catalytic activity of VioG, which may result from the residual oxygen in the reactions. Since the [4Fe–4S] cluster in radical SAM enzymes is sensitive to oxygen,(31) incubation of M. xanthus::Ptet-vioG cell lysates and VioH under strict anaerobic conditions may improve the production of MAZ. In addition, it cannot be excluded that SAM is methylated by VioG before being cyclized by VioH. However, no products related to methyl-SAM could be detected in the incubation of the cell lysate with SAM. Due to the difficulties in obtaining soluble VioG and methyl-SAM, the mechanism of MAZ formation cannot be elucidated in detail at present.

196 197 198 199 200 201 202 203 204 205 206 207

Figure 5 UPLC-MS analysis of MAZ-DLA generated by whole cell lysate and VioH. Extracted ion chromatograms of MAZ-DLA and D-proline (m/z 410.20 ± 0.05 [M+H]+) are shown. A, M.xanthus DK1622 (∆mchA) + VioH + SAM + VB12; B, M.xanthus::Ptet-vioG + AZE + SAM without VioH; C, M.xanthus::Ptet-vioG + VioH + SAM + VB12; D, M.xanthus::Ptet-vioG + VioH + VB12 without SAM; E, M.xanthus::Ptet-vioG + VioH + SAM without VB12; F, M.xanthus::Ptet-vioG + VioH + AZE without SAM; G, M.xanthus::Ptet-vioG + VioH + SAM +VB12 without ferric ion; H, M.xanthus::Ptet-vioG + VioH + SAM without VB12 and ferric ion; I, M.xanthus::Ptet-vioG + VioH + SAM + VB12 + (NH4)2Fe(SO4)2; J, M.xanthus::Ptet-vioG + VioH + SAM + VB12 + FeCl2; K, MAZ-DLA reference; L, D-Pro-DLA reference. FeCl3 was added to A – F. Detailed reaction conditions are provided in Supplementary Table S1.

208

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

209 210 211 212 213 214 215 216 217 218

In summary, we reconstituted azetidine-2-carboxylic acid formation by an unusual class I SAM-dependent methyltransferase VioH in vitro. This study thus provides the first example for a new function of class I SAM-MT: the intramolecular cyclization of SAM. The reconstitution of AZE and MAZ via in vivo genetic engineering and in vitro biochemical reactions sets the stage for the detailed elucidation of the biosynthetic mechanism of these unusual amino acids and provides opportunities to generate novel natural products by applying them to combinatorial biosynthesis. Considering that nicotianamine synthase in plants also contains a MT-like domain, a SAM-cyclizing function of MTs could be more widespread among organisms from different kingdoms than currently expected.

219 220

METHODS

221

Strains and cultivation conditions

222 223 224 225

Strains used in this study are listed in Supplementary Table S2. E. coli strains were cultivated in LB broth for routine propagation. M. xanthus DK1622 was cultivated in CTT medium. Kanamycin (50 μg mL-1), chloramphenicol (20 μg mL-1) or oxytetracycline (5 μg mL-1) was added to culture when needed.

226

Construction of expression plasmids

227 228 229 230 231 232 233

Primer sequences are listed in Supplementary Table S2. The genes vioG and vioH were amplified from p15A-Ptet-vio using the primer pairs NheI-orf5-5 / EcoRI-orf5-3 and NheIorf6-5 / EcoRI-orf6-3, and was inserted between NheI and EcoRI sites downstream of the TEV-protease cleavage site on the expression vector pET28b-sumo-tev, resulting His6-Sumo-TEV-VioG and His6-Sumo-TEV-VioH expression vectors pET-STVioG and pET-STVioH, respectively. The open reading frames were validated by Sangersequencing.

234

Generation of M. xanthus mutants

235 236 237 238 239 240 241 242 243 244

The VioG, VioH and VioGH expression constructs p15A-Ptet-vioG, p15A-Ptet-vioH and p15A-Ptet-vioGH were generated by engineering p15A-Ptet-vio. The cm-ccdB cassette was amplified from p15A-ccdB-Cm using primers tetR-cmR-XmaJI-5 and Orf3-ccdB-3, and was transferred into recombinase induced E. coli GBred-gyrA462::p15A-Ptet-vio. The vio gene cluster of p15A-Ptet-vio was then replaced by cmR-ccdB via Red/ET recombination,(32) the resulting plasmid p15A-Ptet-vioorfs-cmccdB was further hydrolyzed by XhoI and cyclized by T4 DNA ligase, yielding p15A-Ptet-cmccdB. The genes vioGH were amplified from p15A-Ptet-vio using the primers Ptet-orf56-5 and Ptetorf56-3 and, the resulting fragment was constructed to p15A-Ptet-cmccdB by Red/ET recombination, yielding p15A-Ptet-vioGH. The fragment cmR-delorf6 was amplified from ACS Paragon Plus Environment

Page 8 of 15

Page 9 of 15 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

245 246 247 248 249 250 251 252 253 254 255 256 257

p15A-Ptet-cmccdB using primers orf5-cmR-XmaJI-5 and cmR-XmaJI-MluI-3, and was transferred into recombinase proficient E. coli GBred::p15A-Ptet-vioGH. The vioH of p15A-Ptet-vioGH was then replaced by cmR gene via Red/ET recombination. The resulting plasmid p15A-Ptet-vioG-cmR was hydrolyzed by XmaJI and cyclized by T4 DNA ligase, yielding p15A-Ptet-vioG. The fragment tetR-Ptet was amplified from p15APtet-vioGH using primers tetR-Ptet-XmaJI-5 and Ptet-orf6-I, and vioH was amplified using primers Ptet-orf6-II and orf6-MluI-3. The two fragments were ligated by overlap extension PCR to generate tetR-Ptet-vioH cassette. The tetR-Ptet-vioG cassette of p15A-Ptet-vioG was replaced by tetR-Ptet-vioH cassette at XmaJI and MluI sites, yielding p15A-Ptet-vioH. The resulting plasmids p15A-Ptet-vioG, p15A-Ptet-vioH and p15A-Ptet-vioGH were transferred into Myxococcus xanthus DK1622 (∆mchA), and the Ptet-vioG-, Ptet-vioH- or Ptet-vioGH-containing cassette was integrated to the chromosome via transposition.

258

Protein expression and purification

259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277

The VioH expression vector pET-STVioH was transferred into E. coli Rosetta (DE3). The resulting strains were cultivated in LB broth at 37 oC overnight. Proportion of 1% (v/v) overnight cultures was inoculated into ZYM-5052 auto-induction medium. After cultivation at 37 oC for 30 min, the cultures were moved to 16 oC and cultivated for 2 days. The cells were collected and stored at -80 oC before protein purification. The cells were suspended in 200 mL wash buffer (20 mM Tris, 150 mM NaCl, 20 mM imidazole, pH 8.0) and were lysed by cell disruptor (Microfluidics Corp.). A tablet of cOmplete™, EDTA-free Protease Inhibitor Cocktail (Roche) was added during cell lysis. Cell debris was removed by centrifugation at 19000 rpm and 4 oC for 40 min. The supernatant was loaded to HisTrap HP 5 mL column connected to AKTA avant and His6-Sumo-TEV-VioH was eluted by a gradient increase of imidazole concentration. The protein was desalted on HiPrep Desalting column (20 mM Tris, pH 8.0), and was then purified with HiTrap Q HP 5 mL column connected to AKTA purifier (gradient concentration of NaCl from 0– 100%). The His6-Sumo-tag was cleaved by incubation with TEV-protease at 4 oC overnight, and was removed by incubation with Ni-NTA agarose resin (Qiagen) and loading to gravity column. The eluents containing tag-free VioH were collected and buffer exchanged using HiPrep Desalting column (20 mM Tris, 200 mM NaCl, 10% glycerol, pH 8.0). The homogenous VioH were concentrated using centrifugal filter (Millipore, 10 kDa cut off), snap frozen in liquid nitrogen and stored at -80 oC.

278

In vitro biochemical assays

279 280 281 282 283

To measure the activity of VioH, 50 μL reaction containing 300 μM protein, 8 mM MgCl2, 2 mM SAM (Sigma or MedChem Express) and buffer (20 mM Tris, 200 mM NaCl, pH 8.0) was incubated at 30 oC for 1 hour. The VioH heated at 100 oC for 5 min was set as negative control. Large molecules were removed by centrifugal filter (Millipore, 10 kDa cut off), and the flow through was adjusted to 50 μL with buffer and modified by Marfey’s 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

284 285 286 287 288 289 290 291 292 293 294 295 296

derivatization (see below). To measure the impact of temperature, 50 μL reaction containing 10 μM protein and 2 mM SAM was incubated at 25–100 oC for 2 hours. The test of pH effects was performed in 50 μL reaction containing 10 μM protein, 2 mM SAM and buffer (20 mM Tris, 200 mM NaCl, pH 6.5–13), and was incubated at 40 oC for 2 hours. The reaction buffer was adjusted to pH 6.5–13 using HCl or NaOH. Since the pH of VioH storage buffer has an effect on the pH of reaction buffer, the pH in the measurements was initially determined by measuring mixtures of storage buffer and reaction buffers at larger scale using a pH meter, and the pH of actual reaction systems was monitored by pH strips. To test effects of buffer component to the activity of VioH, the storage buffer was changed to phosphate buffered saline (PBS, pH 8.0) using Hiprep Desalting column. The 50 μL reactions containing 10 μM protein, 2 mM SAM and PBS buffer (pH 8.0) were performed at 30–90 oC for 2 hours. The reactions were directly modified by Mafey’s derivatization (see below).

297

Derivatization of AZE and LC-MS measurement

298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322

The AZE or MAZ was analyzed by Marfey’s assays.(33) The volume of 20 μL 1M NaHCO3 and 20 μL Marfey’s reagent Nα-(5-Fluoro-2.4-dinitrophenyl)-D-leucinamide (DFDLA, 1% in acetone) were added to 50 μL samples. After incubation at 40 oC with shaking (700 rpm) for 1h, the reaction was stopped by adding 10 μL 2 M HCl and 300 μL acetonitrile. UPLC-MS measurements of the modified AZE (AZE-DLA) were performed on an Dionex Ultimate 3000 RSLC system (Thermo Scientific) using a BEH C18, 100 x 2.1 mm, 1.7 μm dp column (Waters, Germany) coupled to a Bruker Maxis 4G mass spectrometer. Separation of 1 μL sample was achieved by a linear gradient from (A) H2O to (B) methanol at a flow rate of 600 μL min-1 and 45 oC: 0–1 min, 5–10% B; 1–15 min, 10–35% B; 15–22 min, 35–55% B; 22–25 min, 55–80% B; 25–26 min, 80% B, 26– 26.5 min, 80–5% B, 26.5–31 min, 5% B. Full scan mass spectra were acquired in a range from 100–1000 m/z. Quantification of AZE-DLA was carried out on Dionex Ultimate 3000 RSLC system (Thermo Scientific) coupled to a Bruker amaZon speed mass spectrometer. Sulfamethizole (MW 270.333 g mol-1) was added to the samples and served as internal reference. AZE referential compound (Sigma Aldrich) was modified with D-FDLA and diluted to 0.01–5 μM. Calibration curve was obtained by measurements of AZE-DLA references. Separation of 1 μL sample was achieved on a ACQUITY UPLC CSHTM Fluoro-Phenyl column (100 x 2.1 mm, 1.7 μm dp, Waters, Germany) by a linear gradient from (A) H2O to (B) acetonitrile at a flow rate of 600 μL min-1 and 45 oC: 0–0.5 min, 5% B; 0.5–4 min, 5–25% B; 4–18 min, 25–33% B; 18–19.5 min, 33–95% B; 19.5–20.5 min, 95% B, 20.5–20.8 min, 95–5% B, 20.8–21.5 min, 5% B. Full scan mass spectra were acquired in a range from 50–800 m/z. The areas of MS/MS fragments of sulfamethizole (155.85 ± 0.3, [M+H]+) and AZE-DLA (351.08 ± 0.3, [M+H]+) (Supplementary Figure 11) were integrated and analyzed using the software Bruker Compass QuantAnalysis Version 2.2.

323 ACS Paragon Plus Environment

Page 10 of 15

Page 11 of 15 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

324

In vivo production of MAZ

325 326 327 328 329 330 331 332 333

The mutants M. xanthus::Ptet-vioG, M. xanthus::Ptet-vioH and M. xanthus::Ptet-vioGH were cultivated in 50 mL CTT medium (casitone 10 g L-1; 10 mM Tris∙HCl, pH 7.6; 1 mM K2HPO4, pH 7.6; 8 mM MgSO4) at 30 oC for 4.5 days. The portion of 1% Amberlite XAD16 resin was added to the cultures. The compounds were extracted with methanol and evaporated in a vacuum system. The dryness was suspended in 50 μL ddH2O and modified with Marfey’s reagent as mentioned above. Authentic L-MAZ (synthesized previously),(5) L-AZE (Sigma), and D-proline were modified with D-FDLA and served as references. The production of MAZ was measured on Bruker Maxis 4G mass spectrometer as mentioned above.

334

Production of MAZ from whole cell extracts

335 336 337 338 339 340 341 342 343 344 345

M. xanthus DK1622 (∆mchA) and M. xanthus::Ptet-vioG were cultivated in CTT medium containing 2 mg L-1 vitamin B12 at 30 oC for 2 days. The production of MAZ using whole cell extracts of M. xanthus::Ptet-vioG and the purified VioH was achieved in 250 μL PCR tube. A total of 250 μL reaction system contains ~30 mg M. xanthus::Ptet-vioG cells, 50 μM VioH, 1 mM vitamin B12, 8 mM MgCl2, 0.5 mM FeCl3 / (NH4)2Fe(SO4)2 / FeCl2, 0.5 mM Na2S and 2 mM sodiumdithionate. The tubes were flushed under nitrogen immediately after adding 0.2 mg mL-1 lysozyme and CelLytic B cell lysis reagent (Sigma). The reaction was performed at room temperature for 5 hours. The particles and large molecules were removed by centrifugal filter (Millipore, 10 kDa cut off), and 50 μL eluent was modified by Marfey’s reagent. The production of MAZ was measured by UPLC-MS as mentioned above.

346 347

ACKNOWLEDGEMENT

348 349

We thank M. Miethke for helpful suggestions regarding the manuscript and D. Sauer for LC-MS measurements.

350

FUNDING SOURCES

351 352

This work was partially funded by the Deutsche Forschungsgemeinschaft (MU 1254/142).

353

SUPPORTING INFORMATION

354 355

Supporting Information Available: This material is available free of charge via the Internet.

356

SDS-PAGE and mass spectra of VioH, LC-MS analysis, strains and primers.

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

357

REFERENCES

358 359 360

1. Martin, J. L., and McMillan, F. M. (2002) SAM (dependent) I AM. The Sadenosylmethionine-dependent methyltransferase fold, Curr. Opin. Struct. Biol. 12, 783– 793.

361 362

2. Schubert, H. L., Blumenthal, R. M., and Cheng, X. (2003) Many paths to methyltransfer. A chronicle of convergence, Trends Biochem. Sci. 28, 329–335.

363 364 365

3. Struck, A.-W., Thompson, M. L., Wong, L. S., and Micklefield, J. (2012) S-adenosylmethionine-dependent methyltransferases. Highly versatile enzymes in biocatalysis, biosynthesis and other biotechnological applications, Chembiochem 13, 2642–2655.

366 367 368

4. Schummer, D., Höfle, G., Forche, E., Reichenbach, H., Wray, V., and Domke, T. (1996) Antibiotics from gliding bacteria, LXXVI. Vioprolides: new antifungal and cytotoxic peptolides from Cystobacter violaceus, Liebigs Ann./Recl. 1996, 971–978.

369 370 371 372 373

5. Yan, F., Auerbach, D., Chai, Y., Keller, L., Tu, Q., Hüttel, S., Glemser, A., Grab, H. A., Bach, T., Zhang, Y., and Müller, R. (2018) Biosynthesis and heterologous production of vioprolides: rational biosynthetic engineering and unprecedented 4methylazetidinecarboxylic acid formation, Angew. Chem. Int. Ed. Engl. Epub 2018. DOI: 10.1002/anie.201802479.

374 375 376

6. Wu, G., Nielson, J. R., Peterson, R. T., and Winter, J. M. (2017) Bonnevillamides, Linear Heptapeptides Isolated from a Great Salt Lake-Derived Streptomyces sp, Mar. Drugs Epub 2017. DOI: 10.3390/md15070195.

377 378

7. Isono, K., Asahi, K., and Suzuki, S. (1969) Polyoxins, antifungal antibiotics. XIII. Structure of polyoxins, J. Am. Chem. Soc. 91, 7490–7505.

379 380

8. Noma, M., Noguchi, M., and Tamaki, E. (1971) A new amino acid, nicotianamine, from tobacco leaves, Tetrahedron Lett. 12, 2017–2020.

381 382 383

9. Kristensen, I., and Larsen, P. O. (1974) Azetidine-2-carboxylic acid derivatives from seeds of Fagus silvatica L. and a revised structure for nicotianamine, Phytochemistry 13, 2791–2798.

384 385 386 387

10. Kobayashi, J.'i., Cheng, J.-F., Ishibashi, M., Wälchli, M. R., Yamamura, S., and Ohizumi, Y. (1991) Penaresidin A and B, two novel azetidine alkaloids with potent actomyosin ATPase-activating activity from the Okinawan marine sponge Penares sp, J. Chem. Soc., Perkin Trans. 1 55, 1135–1137.

388 389 390

11. Kitajima, M., Kogure, N., Yamaguchi, K., Takayama, H., and Aimi, N. (2003) Structure reinvestigation of gelsemoxonine, a constituent of Gelsemium elegans, reveals a novel, azetidine-containing indole alkaloid, Org. Lett. 5, 2075–2078.

391 392 393 394

12. Akihisa, T., Mafune, S., Ukiya, M., Kimura, Y., Yasukawa, K., Suzuki, T., Tokuda, H., Tanabe, N., and Fukuoka, T. (2004) (+)- and (-)-syn-2-isobutyl-4-methylazetidine-2,4dicarboxylic acids from the extract of Monascus pilosus-fermented rice (red-mold rice), J. Nat. Prod. 67, 479–480.

ACS Paragon Plus Environment

Page 12 of 15

Page 13 of 15 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

395 396 397

13. HAYASHI, H., TAKIUCHI, K., MURAO, S., and ARAI, M. (1988) Okaramine B, an insecticidal indole alkaloid, produced by Penicillium simplicissimum AK-40, Agric. Biol. Chem. 52, 2131–2133.

398 399 400 401

14. Di, Y.-T., He, H.-P., Wang, Y.-S., Li, L.-B., Lu, Y., Gong, J.-B., Fang, X., Kong, N.-C., Li, S.-L., Zhu, H.-J., and Hao, X.-J. (2007) Isolation, X-ray Crystallography, and Computational Studies of Calydaphninone, a New Alkaloid from Daphniphyllum calycillum, Org. Lett. 9, 1355–1358.

402 403 404

15. Saito, S., Kubota, T., Fukushi, E., Kawabata, J., Zhang, H., and Kobayashi, J.’i. (2007) Calyciphylline C, a novel Daphniphyllum alkaloid from Daphniphyllum calycinum, Tetrahedron Lett. 48, 1587–1589.

405 406 407

16. Kawai, S., Itoh, K., Takagi, S.-i., Iwashita, T., and Nomoto, K. (1988) Studies on phytosiderophores. Biosynthesis of mugineic acid and 2′-deoxymugineic acid in Hordeum vulgare L. var. Minorimugi, Tetrahedron Lett. 29, 1053–1056.

408 409 410

17. Leete, E., Davis, G. E., Hutchinson, C. R., Woo, K. W., and Chedekel, M. R. (1974) Biosynthesis of azetidine-2-carboxylic acid in Convallaria majalis, Phytochemistry 13, 427–433.

411 412

18. Leete, E. (1975) Biosynthesis of azetidine-2-carboxylic acid from methionine in Nicotiana tabacum, Phytochemistry 14, 1983–1984.

413 414 415

19. Leete, E., Louters, L. L., and Prakash Rao, H. S. (1986) Biosynthesis of azetidine-2carboxylic acid in Convallaria majalis: Studies with N-15 labelled precursors, Phytochemistry 25, 2753–2758.

416 417 418

20. Higuchi, K., Suzuki, K., Nakanishi, H., Yamaguchi, H., Nishizawa, N. K., and Mori, S. (1999) Cloning of nicotianamine synthase genes, novel genes involved in the biosynthesis of phytosiderophores, Plant Physiol. 119, 471–480.

419 420

21. Kozbial, P. Z., and Mushegian, A. R. (2005) Natural adenosylmethionine-binding proteins, BMC Struct. Biol. 5, 19.

421 422 423

22. Lai, C.-Y., Lo, I.-W., Hewage, R. T., Chen, Y.-C., Chen, C.-T., Lee, C.-F., Lin, S., Tang, M.-C., and Lin, H.-C. (2017) Biosynthesis of Complex Indole Alkaloids. Elucidation of the Concise Pathway of Okaramines, Angew. Chem. Int. Ed. 56, 9478–9482.

424 425 426

23. Isono, K., Funayama, S., and Suhadolnik, R. J. (1975) Biosynthesis of the polyoxins, nucleoside peptide antibiotics. New metabolic role for L-isoleucine as a precursor for 3ethylidene-L-azetidine-2-carboxylic acid (polyoximic acid), Biochemistry 14, 2992–2996.

427 428 429 430

24. Chen, W., Huang, T., He, X., Meng, Q., You, D., Bai, L., Li, J., Wu, M., Li, R., Xie, Z., Zhou, H., Zhou, X., Tan, H., and Deng, Z. (2009) Characterization of the polyoxin biosynthetic gene cluster from Streptomyces cacaoi and engineered production of polyoxin H, J. Biol. Chem. 284, 10627–10638.

431 432

25. Fowden, L., and Richmond, M. H. (1963) Replacement of proline by azetidine-2carboxylic acid during biosynthesis of protein, Biochim. Biophys. Acta 71, 459–461.

ACS Paragon Plus Environment

history

of

S-

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

433 434

26. Troxler, R. F., and Brown, A. S. (1974) Metabolism of l-azetidine-2-carboxylic acid in the alga Cyanidium caldarium, Biochim. Biophys. Acta 366, 341–349.

435 436

27. Bach, T. M. H., and Takagi, H. (2013) Properties, metabolisms, and applications of (L)-proline analogues, Appl. Microbiol. Biotechnol. 97, 6623–6634.

437 438 439

28. Desiderio, C., Cavallaro, R. A., Rossi, A. de, D'Anselmi, F., Fuso, A., and Scarpa, S. (2005) Evaluation of chemical and diastereoisomeric stability of S-adenosylmethionine in aqueous solution by capillary electrophoresis, J. Pharm. Biomed. Anal. 38, 449–456.

440 441

29. PARKS, L. W., and SCHLENK, F. (1958) The stability and hydrolysis of Sadenosylmethionine; isolation of S-ribosylmethionine, J. Biol. Chem. 230, 295–305.

442 443

30. Borchardt, R. T. (1979) Mechanism of alkaline hydrolysis of S-adenosyl-Lmethionine and related sulfonium nucleosides, J. Am. Chem. Soc. 101, 458–463.

444 445

31. Broderick, J. B., Duffus, B. R., Duschene, K. S., and Shepard, E. M. (2014) Radical S-adenosylmethionine enzymes, Chemical reviews 114, 4229–4317.

446 447 448

32. Wang, H., Bian, X., Xia, L., Ding, X., Müller, R., Zhang, Y., Fu, J., and Stewart, A. F. (2014) Improved seamless mutagenesis by recombineering using ccdB for counterselection, Nucleic Acids Res. 42, e37.

449 450

33. Bhushan, R., and Brückner, H. (2004) Marfey's reagent for chiral amino acid analysis. A review, Amino acids 27, 231–247.

451

ACS Paragon Plus Environment

Page 14 of 15

Page 15 of 15 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

618x303mm (96 x 96 DPI)

ACS Paragon Plus Environment