structural and mechanistical studies on hydroperoxide conversions

32
1 Structural and mechanistic studies of hydroperoxide conversions catalyzed by a CYP74 clan epoxy alcohol synthase from amphioxus (Branchiostoma floridae) Mats Hamberg 1 , Bulat I. Khairutdinov 3 , Julia Scholz 2 , Florian Brodhun 2 , Ellen Hornung 2 , Ivo Feussner 2 , and Alexander N. Grechkin 3 1 Division of Physiological Chemistry II, Department of Medical Biochemistry and Biophysics, Karolinska Institutet, SE-17177 Stockholm, Sweden; 2 Georg-August-University, Albrecht von-Haller Institute for Plant Sciences, Department of Plant Biochemistry, Justus-von-Liebig-Weg 11, D-37075 Göttingen, Germany; 3 Kazan Institute of Biochemistry and Biophysics, Russian Academy of Sciences, P.O. Box 30, 420111 Kazan, Russia Running title: Studies on amphioxus epoxy alcohol synthase Corresponding author: Mats Hamberg Division of Physiological Chemistry II, Department of Medical Biochemistry and Biophysics, Karolinska Institutet, SE-17177 Stockholm, Sweden Tel. 46 8 52587640; E-mail [email protected] Footnote: Abbreviations: AOS, allene oxide synthase; DES, divinyl ether synthase; EAS, epoxy alcohol synthase; HAS, hemiacetal synthase; GC-MS, gas-liquid chromatography-mass spectrometry; HODE, hydroxyoctadecadienoic acid; HPODE, hydroperoxyoctadecadienoic acid; HPOTrE, hydroperoxyoctadecatrienoic acid; MC, (-)-menthoxycarbonyl; TMS, trimethylsilyl; NMR, nuclear magnetic resonance; SP-HPLC, straight-phase high-performance liquid chromatography; UV, ultraviolet. by guest, on April 4, 2018 www.jlr.org Downloaded from

Upload: doandieu

Post on 03-Feb-2017

217 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Structural and mechanistical studies on hydroperoxide conversions

1

Structural and mechanistic studies of hydroperoxide conversions catalyzed by a CYP74 clan epoxy alcohol synthase from amphioxus (Branchiostoma floridae)

Mats Hamberg1, Bulat I. Khairutdinov3, Julia Scholz2, Florian Brodhun2, Ellen Hornung2, Ivo

Feussner2, and Alexander N. Grechkin3

1 Division of Physiological Chemistry II, Department of Medical Biochemistry and Biophysics,

Karolinska Institutet, SE-17177 Stockholm, Sweden; 2 Georg-August-University, Albrecht von-Haller

Institute for Plant Sciences, Department of Plant Biochemistry, Justus-von-Liebig-Weg 11, D-37075

Göttingen, Germany; 3 Kazan Institute of Biochemistry and Biophysics, Russian Academy of

Sciences, P.O. Box 30, 420111 Kazan, Russia

Running title: Studies on amphioxus epoxy alcohol synthase

Corresponding author:

Mats Hamberg

Division of Physiological Chemistry II, Department of Medical Biochemistry and Biophysics,

Karolinska Institutet, SE-17177 Stockholm, Sweden

Tel. 46 8 52587640; E-mail [email protected]

Footnote: Abbreviations: AOS, allene oxide synthase; DES, divinyl ether synthase; EAS, epoxy

alcohol synthase; HAS, hemiacetal synthase; GC-MS, gas-liquid chromatography-mass spectrometry;

HODE, hydroxyoctadecadienoic acid; HPODE, hydroperoxyoctadecadienoic acid; HPOTrE,

hydroperoxyoctadecatrienoic acid; MC, (-)-menthoxycarbonyl; TMS, trimethylsilyl; NMR, nuclear

magnetic resonance; SP-HPLC, straight-phase high-performance liquid chromatography; UV,

ultraviolet.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 2: Structural and mechanistical studies on hydroperoxide conversions

2

Abstract

The CYP74 family of cytochrome P450 enzymes consisting of allene oxide synthase, divinyl ether

synthase and hemiacetal synthase has important roles in plant oxylipin biosynthesis, transforming

lipoxygenase-generated fatty acid hydroperoxides into a range of biologically active compounds.

Biochemical and phylogenetic studies have shown that CYP74-related enzymes are found also in

marine invertebrates, such as coral, sea anemone and amphioxus (Branchiostoma floridae) (Lee, D.-S,

Nioche, P., Hamberg, M. and Raman, C.S., Nature 455:363-368 (2008)). Here, a CYP74 clan enzyme

from amphioxus, BfEAS, displaying epoxy alcohol synthase activity was incubated with

hydroperoxides of linoleic and linolenic acids and the products identified by chemical transformations,

mass spectrometry and NMR. Linoleic acid 13(S)-hydroperoxide was transformed into 12(R),13(S)-

epoxy-11(S)-hydroxy-9(Z)-octadecenoic acid, and an analogous conversion was observed with

linolenic acid 13(S)-hydroperoxide. Incubation of linoleic acid 13(S)-[18O2]-hydroperoxide led to the

formation of epoxy alcohol retaining both atoms of 18O, thus indicating a mechanism involving

homolysis of the hydroperoxide, formation of the epoxide by alkoxy radical attack at the α-

unsaturation, and rebound of hydroperoxide oxygen to form the 11(S) alcohol group. The 9(S)-

hydroperoxide of linoleic acid produced 9(S),10(R)-epoxy-11(S)-hydroxy-12(Z)-octadecenoic acid

accompanied by smaller amounts of 9(S),14(R,S)-dihydroxy-10(E),12(E)-octadecadienoic acid and the

new macrolactone 9(S),10(R)-epoxy-11(E)-octadecen-13(S)-olide. In conclusion, three relevant

lipoxygenase hydroperoxides were converted to epoxy alcohols by BfEAS, supporting the discovery

that EAS constitutes a fourth category in the CYP74 clan of enzymes.

Supplementary Key Words Fatty acid hydroperoxide / CYP74 / Epoxy alcohol synthase /

amphioxus (Branchiostoma floridae)

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 3: Structural and mechanistical studies on hydroperoxide conversions

3

Introduction

Cytochrome P-450 enzymes belonging to the CYP74 family, i.e. allene oxide synthase (AOS),

divinyl ether synthase (DES) and hemiacetal synthase (HAS; also referred to as hydroperoxide lyase),

are responsible for most of the conversions which take place with lipoxygenase-generated

hydroperoxides in plants. Products formed in CYP74-initiated pathways include the jasmonates as

well as short-chain aldehydes and alcohols, compounds which have important roles to fulfill in plant

development and defense. Studies of the mechanism of CYP74-catalyzed reactions indicate that the

first step of hydroperoxide conversion results in an epoxyallylic radical which serves as a common

intermediate in the formation of products by AOS, DES and HAS (1). As seen in Fig. 1, the native

Fe(III) form of the various CYP74 enzymes is converted to Fe(IV)-OH during the initial step, and this

oxidized form is utilized to accomplish the subsequent transformations into allene oxides, divinyl

ethers and hemiacetals.

Epoxy alcohols constitute another group of biologically important oxylipins generated from

lipoxygenase-derived hydroperoxides. Two distinct mechanisms exist for the formation of epoxy

alcohols (2), i.e. either reduction of the hydroperoxide functional group into an alcohol and

epoxidation of one of the conjugated double bonds, or homolytic cleavage of the hydroperoxide O-O

bond, cyclization of the resulting alkoxy radical by attack at the neighboring unsaturated carbon atom

and recapture of OH. The former mechanism has been observed in the vanadium-promoted

degradation of hydroperoxides (3), for peroxygenase (4,5) and for a number of unidentified epoxy

alcohol synthases present in certain plants and fungi (ref. 6, and references cited therein). The latter

mechanism most often results in the formation of trans-configured epoxy alcohols and has been

observed for various nonenzymatic (7,8) and lipoxygenase-promoted conversions (9,10), as well as for

specific enzymatic hydroperoxide isomerizations catalyzed by lipoxygenase-type enzymes (11,12).

Recently structural, biochemical, and phylogenetic studies showed that amphioxus (Branchiostoma

floridae) contains several copies of CYP74-related genes, at least one of which encoding an epoxy

alcohol synthase (BfEAS) (13). The aim of the present work was to further characterize this enzyme

with respect to substrate specificity, products and mechanism.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 4: Structural and mechanistical studies on hydroperoxide conversions

4

MATERIALS AND METHODS

Fatty acid hydroperoxides of >98% chemical purity were purchased from Lipidox, Stockholm,

Sweden. Erythro- and threo-10,11- and 11,12-dihydroxyoctadecanoates were prepared by cis-

hydroxylation (osmium tetroxide) or trans-hydroxylation (performic acid followed by saponification)

of the corresponding (Z)-monounsaturated fatty acid methyl esters (4). Standards of racemic and

optically active short chain 2-hydroxy acids used as references during steric analysis were available

since earlier work (cf. ref. 14,15). Agarose was purchased from Biozym Scientitic GmbH, and

restriction enzymes were from MBI Fermentas. Frozen adult amphioxus were a generous gift from Dr.

Linda Holland (Scripps Institution of Oceanography, University of California, San Diego, U.S.A.).

Cloning and expression

A codon-optimized gene encoding for BfEAS (CYP440A1, GenBank: ACD88492.1) (13) was

synthesized by the GENEWIZ Company (South Plainfield, USA) with a C-terminal hexahistidine

sequence essential for immobilized Ni-affinity chromatography. In order to improve expression of

soluble protein, the MAKKTSS-sequence was added at the N-terminus. The resulting construct was

cloned into the pET-vector and transformed into E. coli Bl21star. For heterologous expression the

recombinant cells were pre-cultivated in LB- or 2xYT-medium until an OD600 of approx. 0.6-0.8 was

reached. Protein expression was induced by the addition of 0.1 mM IPTG. 80 µg/mL δ-aminolevulinic

acid and 150 µM ammonium iron citrate were added at the time point of induction in order to ensure

that the co-factor availability is not limiting expression of BfEAS. After cultivation at 16 °C for 3 d

cells were harvested by centrifugation (8000 x g, 20 min, 4 °C), shock frozen in liquid nitrogen and

stored at -20 °C.

Cell lysis and affinity purification

Cell disruption was performed according to a protocol adapted from that established by

Richardson et al. (16). Briefly, cells from approx. 1L of culture were resuspended in 150 mL 100 mM

Tris/HCl (pH 7.8) containing 20 % glycerol. After addition of 0.2 mg/mL lysozyme the cell

suspension was stirringly incubated at 4 °C for 30 min and centrifuged for 10 min at 3000 x g.

Sedimented spheroblasts were dissolved in 50 mL of 10 mM phosphate buffer (pH 8.0) containing 14

mM MgAc2, 60 mM KAc, 0.1 mM DTT and were frozen at -80 °C (>12 h). PMSF was added to a

final concentration of 0.5 mM and cell lysis was increased by applying a sonifier cell disruptor B15

from Branson (5 x 1 min at 40% power and 40% pulse). Cell debris was removed by centrifugation for

20 min at 50 000 x g. The resulting supernatant was applied on a HisTrapTM HP column from GE

Healthcare that was pre-equilibrated with 50 mM phosphate buffer (pH 8.0) containing 1M NaCl and

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 5: Structural and mechanistical studies on hydroperoxide conversions

5

0.5 M urea. After washing with 10-20 column volumes of this buffer unspecifically bound proteins

were eluted by increasing the imidazol concentration to 15 mM. BfEAS was eluted by constantly

increasing the imidazol concentration to 0.3 M imidazol within 20 min at a flow rate of 1 mL/min. The

purity of BfEAS was assessed by means of SDS-PAGE analysis and Coomassie-staining.

Spectroscopic analysis

UV/vis spectra of the affinity purified BfEAS were performed at room temperature in phosphate

buffer (pH 8.0) using a Cary 100 Bio spectrophotometer from Varian. CD-spectra of BfEAS were

obtained under similar conditions using a Chirsacan CD Spetrometer from Applied Photophysics.

Incubations

The standard incubation of BfEAS, which was scaled up as needed, involved addition of

recombinant enzyme to the fatty acid hydroperoxide (300 µM) in 0.1 M potassium phosphate buffer

pH 7.4 (4 mL). The mixture was stirred at 23oC for 15 min and then extracted with diethyl ether. For

profiling of products using GC-MS, aliquots were treated with diazomethane followed by

trimethylchlorosilane-hexamethyldisilazane-pyridine (2:1:2, v/v/v) to generate methyl ester /

trimethylsilyl (TMS) ether derivatives.

For incubations with amphioxus homogenate, 150 mg of tissue in 10 mL of 0.1 M potassium

phosphate buffer pH 7.4 containing 100 µM linoleic acid was homogenized at 0o for 2 x 30 sec using

an Ultraturrax. The homogenate was stirred at 23o for 30 min, then extracted with diethyl ether.

Material was derivatized as methyl esters - TMS ether derivatives and analyzed by GC-MS using a

library of oxylipin standard mass spectra for product identification.

Preparative HPLC

Material generated in large scale incubations using 10-20 mg of hydroperoxide was treated with

diazomethane and subjected to straight-phase (SP) HPLC using a column of Nucleosil 50-7 (250 x 10

mm, Marcherey-Nagel, Düren, Germany). The column was eluted at a flow rate of 4 mL/min with

solvent systems of 0.3 - 5% 2-propanol in hexane as indicated. Serially connected detectors for

measurement of the absorption at 210 nm and refractive index were used.

Steric analysis

Analytes were derivatized to methyl ester / (-)-menthoxycarbonyl (MC) derivatives, purified by

TLC and subjected to oxidative ozonolysis as described (15). The methyl-esterified products

containing MC derivatives of short chain 2-hydroxyesters were analyzed by GC-MS using the

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 6: Structural and mechanistical studies on hydroperoxide conversions

6

corresponding derivatives of 2(S)- and 2(R,S)-hydroxyesters as references (14,15). For determination

of erythro-/threo configurations of vicinal methyl dihydroxyoctadecanoates, samples were derivatized

to bis-TMS ether derivatives and the retention times on GC-MS were recorded.

GC-MS analysis

A Hewlett-Packard model 5970B mass selective detector connected to a Hewlett-Packard model

5890 gas chromatograph equipped with a 12 m phenylmethylsilicone capillary column was used for

GC-MS. Helium was used as the carrier gas. The oven temperature was raised from 120oC to 300oC at

a rate of 10oC/min. Alternatively, GC-MS analyses were performed using a Shimadzu QP5050A mass

spectrometer connected to Shimadzu GC-17A gas chromatograph equipped with an MDN-5S (5%

phenyl 95% methylpolysiloxane) fused capillary column (length, 30 m; ID 0.25 mm; film thickness,

0.25 µm). Helium at a flow rate of 30 cm/s was used as the carrier gas. Injections were made in the

split mode using an initial column temperature of 120 °C. The temperature was raised at 10 °C/min

until 240 °C. Full scan or selected ion monitoring (SIM) analyses were both performed using electron

impact ionization at 70 eV.

NMR spectroscopy

The 1H-NMR, COSY, NOESY, HSQC, HSQC-TOCSY and HMBC spectra (600 MHz,

[2H6]benzene, 296 K) were recorded with Bruker Avance III 600 instrument.

RESULTS

Expression and purification

The BfEAS-gene used in the present study has previously been identified and partly

characterized by Lee and co-workers (13). We adapted the reported protocol by using the pET-vector

system for recombinant expression in analogy to the method recently established for the analysis of

CYP74-enzymes from the moss Physcomitrella patens (17). Briefly, we used a cloning strategy that

led to the expression of BfEAS with a C-terminal hexahistidine-tag and an N-terminal MAKKTSS-

sequence extension. Expression of soluble protein was improved by the addition of iron ammonium

citrate and δ-amino-levulinic acid at the time point of induction. Using a single affinity purification

step we were able to obtain BfEAS with a purity of >90% (Fig. 2).

Spectral properties

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 7: Structural and mechanistical studies on hydroperoxide conversions

7

We observed a reddish/brownish color of the protein fraction eluted from the immobilized Ni2+-

column indicating the presence of heme as a co-factor. Indeed, when we analyzed the UV/VIS

spectrum of BfEAS we found an absorption profile characteristic for heme proteins. Four distinct

absorption maxima were observed at 565 nm (α), 539 nm (β), 420 nm (γ, Soret-band), and 368 nm (δ)

(Fig. 3). The relatively low absorption of the Soret-band indicated that the heme-occupancy was

significantly reduced as it has previously been observed for recombinant CYP74-enzymes from P.

patens (17). We calculated the approximate heme content of recombinant BfEAS as 37% based on a

theoretical molar extinction coefficient at 280nm (ε280nm = 89 500 1/(M.cm)) and an expected molar

extinction coefficient of the Soret-band of εSoret = 100 000 1/(M.cm). In order to gather information

about the secondary structure of BfEAS we recorded CD-spectra of the purified enzyme in the far UV.

The respective spectrum is shown in Fig. 3. It shows to main negative peaks at 212 nm and 222 nm

and positive peak at 194 nm. These signals are indicative for proteins with a mainly alpha-helical fold.

Structures of reaction products formed from fatty acid hydroperoxides incubated with BfEAS

Compound 1 – A GC-MS chromatogram of products (Me ester/TMS derivatives) formed upon

incubation of 13(S)-HPODE with BfEAS is shown in Fig. 4A. Compound 1 constituted >95% of the

products (the structural formulae of Compound 1 and other products are shown in Fig. 5). The mass

spectrum (Me/TMS, Fig. 4B) showed [M – Me]+ at m/z 383 and a prominent fragment TMSO+=CH-

CH=CH-(CH2)7-COOCH3 at m/z 285. These results suggested the structure 12,13-epoxy-11-hydroxy-

9-octadecenoic acid as shown in the fragmentation scheme (Fig. 4B, insert).

Larger amounts of the methyl ester of Compound 1 required for structural analysis was obtained

following purification of methyl-esterified material from pooled standard incubations by SP-HPLC

using a solvent system of 2.5% 2-propanol in hexane.

The NMR data for Compound 1 (Me ester) are presented in Table 1. The double bond has the

"Z" configuration as seen from the J9,10 = 11 Hz, and the epoxide is cis-configured as shown by the

J12,13 = 4.3 Hz. The spin-spin interaction between H11 and H12 (J11,12 = 7.7 Hz) does not allow

assignment of the erythro/threo configuration because of the small differences between the relevant

coupling constants in the NMR spectra of erythro and threo cis-epoxyalcohols (12). However, the 2D-

NOESY spectrum (data not shown) revealed an intense cross-peak between H11 and H12 as well as a

less prominent cross-peak between H11 and H13, and these nuclear Overhauser effects indicate the

spatial proximity between these protons (especially H11 and H12) in agreement with the threo

configuration. In addition, the 2D-NOESY data showing a pronounced cross-peak between H11 and

H14a,b gave further evidence for the presence of a cis-epoxide group.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 8: Structural and mechanistical studies on hydroperoxide conversions

8

An aliquot of 100 µg of the methyl ester of Compound 1 was hydrogenated with 5 mg of Pt

catalyst in 1 mL of methanol affording the dihydro derivative. The epoxide function was

deoxygenated by treatment with triphenylphosphine selenide (31 mg) and TFA (3.5 mg) in 1 mL of

CH2Cl2 (14). The methyl 11-hydroxy-12-octadecenoate resulting from this treatment (Fig. 6) was

converted to the MC derivative and subjected to oxidative ozonolysis (14,15). Analysis of the methyl-

esterified product by GC-MS showed the formation of the MC derivative of dimethyl 2(S)-

hydroxydodecane-1,12-dioate, thus demonstrating that Compound 1 had the “S” configuration at C-11.

Another part of the methyl ester of Compound 1 (100 µg) was hydrogenated with 5 mg of palladium-

on-carbon in 1 mL of methanol affording methyl 11,12-dihydroxyoctadecanoate as the main product

(Fig. 6). The structure of this hydrogenolysis product was demonstrated by the mass spectrum

recorded on the TMS derivative showing prominent ions at m/z 443 (M+ - 31; loss of OCH3), 360 (M+

- 114; loss of OHC-(CH2)5-CH3), 287 (TMSO+=CH-(CH2)9-COOCH3), and 187 (TMSO+=CH-(CH2)5-

CH3). GC-MS analysis using standards of the corresponding derivatives of erythro- and threo-11,12-

dihydroxyoctadecanoic acids (retention times, 12.98 min (erythro) and 12.86 min (threo)) showed that

the BfEAS-derived material cochromatographed with the latter of these isomers, thus establishing the

threo relative configuration between the C-11/C-12 oxygens in Compound 1.

In conclusion, Compound 1 formed from 13(S)-HPODE in the presence of BfEAS was

identified as the cis-epoxy alcohol 12(R),13(S)-epoxy-11(S)-hydroxy-9(Z)-octadecenoic acid.

Compound 2 – GC-MS analysis of products (Me/TMS) formed upon incubation of 13(S)-

HPOTrE with BfEAS revealed a single predominant product, i.e. Compound 2 (Fig. 4D). The mass

spectrum (not illustrated) showed M+ at m/z 396 (0.02%), [M – Me]+ at m/z 381 (0.3%), and

TMSO+=CH-CH=CH-(CH2)7-COOCH3 at m/z 285 (73%). These data enabled us to ascribe the

structure of 12,13-epoxy-11-hydroxy-9,15-octadecadienoic acid to Compound 2, however the full

stereochemistry was not determined. Two minor products (< 1% of the product) were identified as the

α-ketol 13-hydroxy-12-oxo-9,15-ocatecadienoic acid and the cyclopentenone cis-12-oxo-10,15-

phytodienoic acid. Such AOS products were not detected after incubation of 13(S)-HPODE with

BfEAS.

Compound 4 – Three products appeared following incubation of 9(S)-HPODE with BfEAS as

determined by GC-MS analysis, i.e. Compounds 3, 4 and 5 (Fig. 7A); of these, Compound 4 was the

major product. The methyl ester derivatives of Compound 4 and 5 were obtained in pure form

following SP-HPLC using solvent systems of 2.5% 2-propanol in hexane (methyl ester of Compound

4) and 5% 2-propanol in hexane (methyl ester of Compound 5, two diastereomers).

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 9: Structural and mechanistical studies on hydroperoxide conversions

9

The mass spectrum (Fig. 7C) of Compound 4 (Me ester - TMS derivative) showed M+ at m/z

398 (0.1%); [M – Me]+ at m/z 383 (1%); [M – OHC(CH2)7COOCH3]+ at m/z 212 (3%) and

TMSO+=CH-CH=CH-(CH2)4-CH3 at m/z 199 (100%), as shown in the fragmentation scheme (Fig. 7C,

insert). The NMR spectral parameters of Compound 4 Me ester (suppl. Table 1) were nearly identical

to those of Compound 1. The spin-spin coupling value J12,13 = 11.0 Hz indicated that the double bond

had the "Z" configuration. The oxirane protons had cis assignment as seen from their coupling

constant value J9,10 = 4.3 Hz. The spin-spin coupling constant of the secondary alcohol methine (C11)

and the neighboring oxirane methine (C10), J10,11 = 7.7 Hz, did not provide information about the

erythro/threo relationship as mentioned above. However, the 2D-NOESY data exhibited a nuclear

Overhauser effect, a pronounced cross-peak between H10 and H11, which was in agreement with the

threo configuration.

An aliquot of the methyl ester of Compound 4 (100 µg) was subjected to catalytic

hydrogenation using Pt catalyst followed by deoxygenation as described above for the methyl ester of

Compound 1; this afforded methyl 11-hydroxy-9-octadecenoate. Derivatization of the material to its

MC derivative and oxidative ozonolysis generated the MC derivative of 2(S)-hydroxynonanoic acid

having a stereochemical purity of about 92%, thus demonstrating that the absolute configuration at C-

11 of Compound 4 was mainly "S". In another experiment, 100 µg of the methyl ester of Compound 4

was hydrogenated using palladium-on-carbon. The TMS derivative of the resulting hydrogenolysis

product, i.e. methyl 10,11-dihydroxyoctadecanoate, was analyzed by GC-MS using references of the

authentic erythro- and threo-10,11-diols (threo diol eluting 0.12 min ahead of the erythro diol). This

analysis showed that the 10,11-dihydroxyoctadecanoate derived from Compound 4 was mainly due to

the threo isomer, i.e. threo/erythro isomers appearing in a ratio of 93:7.

These results, demonstrated that Compound 4 was mainly due to 9(S),10(R)-epoxy-11(S)-

hydroxy-12(Z)-octadecenoic acid, although 7-8% appeared to exist with the 11(R) configuration.

Compound 5 – The mass spectrum of Compound 5 (Me/TMS) is shown in Fig. 7D. The

spectrum possessed M+ at m/z 470 (0.5%); [M – Me]+ at m/z 455 (0.4%); [M – TMSOH]+ at m/z 380

(22%), as well as other fragment ions depicted in the fragmentation scheme (Fig. 7D, insert). These

mass spectral data suggested the presence of a 9,14-diol structure.

The methyl ester of Compound 5 isolated as described above was subjected to NMR analyses.

The NMR data (suppl. Table 2), including the 1H-NMR, 2D-COSY, NOESY, HSQC and HMBC)

demonstrated the structure 9,14-dihydroxy-10(E),12(E)-octadecadienoic acid (Me ester). The signals

of protons H9/H14, H10/H13 and H11/H12 exhibited a pairwise coincidence due to the symmetrical

structure of two conjugated double bonds placed between two secondary alcohol functions. A complex

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 10: Structural and mechanistical studies on hydroperoxide conversions

10

multiplicity of signals of double bond methins in a 1H-NMR spectrum does not allow estimating the

spin-spin coupling constant values. However, the data of HSQC-TOCSY enabled us to measure the

spin constants J10,11 = J12,13 = 16.2 Hz. This showed that the double bonds at C10 and C12 had both the

"E" configuration.

Steric analysis of the methyl ester - MC derivative of Compound 5 (100 µg) showed the

formation of the MC derivatives of dimethyl 2(S)-hydroxydecane-1,10-dioate and methyl 2(R,S)-

hydroxyhexanoate, thus proving the presence in Compound 5 of a 9(S)-hydroxyl group, a racemic

alcohol function at C-14, and a conjugated diene structure located at C-10 - C-13. Together with the

NMR data, these results identified Compound 5 as 9(S),14(R,S)-dihydroxy-10(E),12(E)-

octadecadienoic acid.

Compound 3 – Compound 3 appeared as a minor component constituting about 3% of the

products formed upon incubation of 9(S)-HPODE with BfEAS (Fig. 7A). In contrast to all other

products formed by BfEAS, analysis of Compound 3 by GC-MS did not require methyl-esterification

indicating that the carboxyl group was derivatized. It was also the least polar of the products

encountered, and a convenient way to generate large amounts for structural work was to perform two-

phase incubations in the following way. 9(S)-HPODE (320 µM) was added to a mixture of 40 mL of

potassium phosphate buffer pH 7.4 containing BfEAS and 40 mL of hexane. The two-phase system

was vigorously shaken for 15 min and the hexane layer was taken to dryness. Analysis of the material

by GC-MS (before and after treatment with diazomethane and trimethylsilylating reagent) showed the

presence of Compound 3 as the sole product. Material for structural work was obtained by performing

several such incubations using totally 14 mg of 9(S)-HPODE followed by purification by SP-HPLC

with 0.3% 2-propanol in hexane as the solvent system; this afforded 2.1 mg of Compound 3 as a

colorless oil.

The mass spectrum of Compound 3 (Fig. 7B) showed M+ at m/z 294 (0.6%), [M – CH3(CH2)4]+

at m/z 223 (6%), [M – CH3(CH2)4CHCH=CHCH]+ at m/z 171 (13%), and [M –

CH3(CH2)4CHCH=CHCHO]+ at m/z 155 (49%). The proposed structure and its mass fragmentation

scheme is shown in Fig. 7B (insert). For further elucidation of the structure of Compound 3 we

recorded its NMR spectra.

The NMR spectrum of Compound 3 (Fig. 8 and suppl. Table 3) possessed some features

atypical for ordinary straight chain oxylipins. The heteronuclear multiple bond correlations revealed

the neighborhood of C1 and C13, thus indicating the existence of an intramolecular ester bond. The

H13 signal was strongly shifted downfield to 5.51 ppm due to the deshielding effect of the

neighboring carbonyl (C1), also confirming the ester link between C1 and C13. All methylene protons

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 11: Structural and mechanistical studies on hydroperoxide conversions

11

from H2 to H8 are chemically non-equivalent since they belong to a macrolactone ring. The oxirane

protons had a coupling constant J9,10 = 4.2 Hz, thus proving a cis-epoxide. The olefinic protons

exhibited J11,12 = 15.4 Hz, demonstrating that the double bond had the "E" configuration. Taken

together, the MS and NMR data enabled us to ascribe the structure cis-9,10-epoxy-11(E)-octadecen-

13-olide to Compound 3.

The stereochemical features of Compound 3 were determined by the conversions shown in Fig.

9. Compound 3 (500 µg) was dissolved in 0.5 mL of methanol and 4.5 mL of water and 50 µL of 2 M

HCl were added. After stirring for 10 min at 23oC, material was extracted with diethyl ether. An

aliquot was analyzed GC-MS as the TMS derivative showing two peaks of isomers in a 3:1 ratio. The

deduced molecular weight of this material was 312, corresponding to addition of one H2O to

Compound 3 (Fig. 9). Subsequent treatment with 0.2 M NaOH in 90% aqueous methanol at 23oC for

18 h liberated the carboxyl in its free form and afforded two major trihydroxy-octadecenoate isomers.

The methyl ester - TMS derivatives of those showed the following prominent mass spectral ions: m/z

545 (M+ - 15; loss of CH3), 460 (M+ - 100; rearrangement and loss of OHC-(CH2)4-CH3), 259 (100%;

TMSO+=CH-(CH2)7-COOCH3), and 155 (isomer a), and m/z 460 (M+ - 100; rearrangement and loss of

OHC-(CH2)4-CH3), 259 (TMSO+=CH-(CH2)7-COOCH3), and 173 (100%; TMSO+=CH-(CH2)4-CH3)

(isomer b) (Fig. 9). Using mass spectral ions and GLC retention times, the two isomers could be

structurally and stereochemically matched with isomeric trihydroxyoctadecenoates known since

previous work (15); this demonstrated that isomer a was identical to methyl 9(S),10(S),13(S)-

trihydroxy-11(E)-octadecenoate whereas isomer b was identical to methyl 9(S),12(R),13(S)-

trihydroxy-10(E)-octadecenoate. Independent support for the configurations at C-13 and C-9 of

isomers a and b was provided by oxidative ozonolysis which produced 2(S)-hydroxyheptanoic acid

and 2(S)-hydroxydecane-1,10-dioic acid as the main chiral fragments. On the basis of these findings

and the NMR data, it was concluded that Compound 3 had the structure 9(S),10(R)-epoxy-11(E)-

octadecen-13(S)-olide.

18O Labeling Experiments

In order to study the origin of the epoxy and hydroxy oxygen atoms in Compound 1 we

performed incubations of BfEAS with [18O2-hydroperoxy]13(S)-HPODE, and of unlabeled 13(S)-

HPODE in [18O]H2O. The products (Me/TMS) were analyzed by GC-MS. The mass spectrum of

derivatized Compound 1 biosynthesized from unlabelled 13-HPODE showed M+ at m/z 398 and [M –

Me]+ at m/z 383 (Fig. 4B). In contrast, Compound 1 (Me/TMS) formed upon BfEAS incubation with

[18O2]13-HPODE exhibited M+ at m/z 402 and [M – Me]+ at m/z 387 (Fig. 4C, Table 2), thus

indicating the incorporation of both 18O atoms from the hydroperoxide into the product. Selected ion

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 12: Structural and mechanistical studies on hydroperoxide conversions

12

monitoring (SIM) analyses of products (Me/TMS) and integration of the intensity of ions at m/z 402.4,

400.4 and 398.4 enabled us to quantify the molecular species of Compound 1 having 2, 1 and 0

incorporated 18O atoms, respectively (Table 2A). The species containing two 18O atoms constituted

94.93% of total integral intensities of three species. The estimated total isotopic content of 18O in

Compound 1 was 96.12%, while the [18O2]13-HPODE precursor contained 96.98% of 18O (Table 2B).

Thus, the experiments revealed nearly quantitative incorporation of two 18O atoms from [18O2]13-

HPODE into Compound 1. Upon incubation of BfEAS with unlabeled 13-HPODE in [18O]water (98% 18O) no incorporation of 18O into Compound 1 was observed as shown by SIM GC-MS.

Identification of products formed from linoleic acid incubated with amphioxus homogenate

Analysis of the derivatized reaction product formed from linoleic acid incubated with

amphioxus whole homogenate showed the presence of three major compounds, i.e. 9-hydroxy-10,12-

octadecadienoic acid, 9,10-epoxy-11-hydroxy-12-octadecenoic acid and 9,10-epoxy-13-hydroxy-11-

octadecenoic acid. Smaller amounts of 9-keto-10,12-octadecadienoic acid, isomeric 9,10,13- and

9,12,13-trihydroxyoctadecenoates, and 2-hydroxylinoleic acid were also observed. Because of the

limited amounts of material available, the stereochemistry of these products was not determined.

Linoleic acid is an endogenous fatty acid in amphioxus (18), and its metabolism by lipoxygenase

apparently involves 9-LOX activity.

DISCUSSION

The bond dissociation energy of the oxygen-oxygen bond of the hydroperoxy group is only

about 44 kcal/mol, and the facile homolytic cleavage of this bond accounts for most of the reactivity of

fatty acid hydroperoxides. The resulting alkoxy radical can undergo a number of transformations, one

of which consists of intramolecular attack at the neighboring unsaturated carbon atom forming a

delocalized epoxyallylic radical (7,8). Already 40 years ago structures produced by trapping of such

radicals were published, i.e. formation of epoxy adducts to tocopherol (19) as well as a specific epoxy

alcohol during heating of a fatty acid hydroperoxide (20). Since then, numerous agents have been

found to promote epoxy alcohol formation by the homolytic hydroperoxide cleavage/cyclization route,

e.g. ferrous ion and other metal ions, hemin, hemoglobin, and UV-light (7,8). Soybean lipoxygenase

operating under anaerobic conditions also produces epoxy alcohols together with other products (9).

Hepoxilins constitute a group of eicosanoids formed from arachidonic acid 12-hydroperoxide in the

presence of e.g. 12-lipoxygenase (10), and a specific conversion of arachidonic acid 12(R)-

hydroperoxide into the epoxy alcohol 11(R),12(R)-epoxy-8(R)-hydroxy-5(Z),9(E),14(Z)-eicosatrienoic

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 13: Structural and mechanistical studies on hydroperoxide conversions

13

acid is catalyzed by the hydroperoxide isomerase activity of epidermal lipoxygenase type 3 (eLOX3)

(11). In a recently published study, the cis-configured epoxy alcohol 8(R),9(S)-epoxy-10(S)-hydroxy-

11(Z),14(Z)-eicosadienoic acid was identified as the product formed from dihomo-γ-linolenic acid

incubated with an 8(R)-lipoxygenase from the coral Plexaura homomalla (12).

Earlier studies have shown that P450s (21) and AOS (22) can produce epoxy alcohols in

reactions which are considerably more specific compared to simple hematin-promoted hydroperoxide

degradations, however, a CYP74 clan enzyme acting as a true EAS was reported only recently (13). In

that study a recombinant CYP74 clan enzyme from amphioxus produced >95% of a cis-configured

epoxy alcohol when incubated with 13(S)-HPODE. The present paper reports further studies of this

enzyme regarding substrates, products and mechanism.

The hydroperoxides 13(S)-HPODE and 13(S)-HPOTrE were each converted into a single epoxy

alcohol by BfEAS, i.e. 12(R),13(S)-epoxy-11(S)-hydroxy-9(Z)-octadecenoic acid (Compound 1) and

12,13-epoxy-11-hydroxy-9,15-octadecadienoic acid (Compound 2), respectively. Studies with 13(S)-

[18O2]HPODE revealed that both hydroperoxide oxygens were fully retained in the epoxy alcohol

product, thus excluding mechanisms involving an epoxyallylic carbocation and incorporation of OH-

from water (23). The 9-LOX-derived hydroperoxide 9(S)-HPODE also produced an epoxy alcohol as

the main product, i.e. 9(S),10(R)-epoxy-11(S)-hydroxy-12(Z)-octadececenoic acid (Compound 4). The

general mechanism postulated for formation of epoxy alcohols by BfEAS (Fig. 10) involves initial

formation of an alkoxy radical by hydroperoxide O-O homolysis followed by cyclization into an

epoxyallylic radical and formation of the C-11 hydroxyl group by rebound of hydroperoxide oxygen

via Fe(IV)-OH. Two additional products formed from 9(S)-HPODE upon incubation with BfEAS

were characterized, i.e. Compound 5, which was identified as the diol 9(S),14(R,S)-dihydroxy-

10(E),12(E)-octadecadienoic acid, and Compound 3, which was identified as the macrolactone

9(S),10(R)-epoxy-11(E)-octadecen-13(S)-olide. It seems likely that formation of these products took

place from a common epoxyallylic carbocation intermediate formed by rearrangement and 1-electron-

oxidation of the epoxyallylic radical (Fig. 10). Loss of a proton from C-14 would lead to an

epoxydiene, which according to previous work (24,25) is quite unstable and suffers spontaneous

hydrolysis by solvent attack at C-14, thus forming the diol isolated. Alternatively, attack by the

carboxylate group at the C-13 carbocation will lead to formation of the macrolactone product.

Somewhat unexpectedly, this attack occurred in a stereospecific way, most likely when the

carbocation was still bound to the enzyme, and provided only the 13(S)-configured product. Another

example of stereospecific formation of an oxylipin macrolactone was reported very recently (26).

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 14: Structural and mechanistical studies on hydroperoxide conversions

14

Whereas the AOS and DES enzymes of the CYP74 family catalyze a formal dehydration of

their substrates, the EAS described in the present paper acts as a hydroperoxide isomerase. One more

member of the CYP74 family, i.e. HAS (also referred to as hydroperoxide lyase; belonging to the

CYP74B and CYP74C subgroups), has been shown to be a hydroperoxide isomerase, in this case

catalyzing formation of unstable fatty acid hemiacetals as shown in Fig. 1. Experiments with 18O2-

labeled hydroperoxides have revealed that both 18O atoms from hydroperoxides are incorporated into

hemiacetals (27-29); therefore the hydroperoxide homolysis/rebound mechanism is common to EAS

and HAS.

The possible function in amphioxus of the linoleic acid-derived CYP74 products encountered in

the present study is entirely unknown. When homogenates of amphioxus were incubated with linoleic

acid, exclusive formation of 9-LOX products were observed, therefore suggesting that Compounds 3-5

may be biologically relevant in amphioxus. However, linoleic acid is a minor fatty acid in amphioxus

(18), and it is possible that analogous or related products formed from the major amphioxus fatty acid,

i.e. docosahexaenoic acid (18), are of greater biological importance.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 15: Structural and mechanistical studies on hydroperoxide conversions

15

References

1. Brash, A.R. 2009. Mechanistic aspects of CYP74 allene oxide synthases and related cytochrome

P450 enzymes. Phytochemistry 70: 1522-1531.

2. Hamberg, M. 1995. Hydroperoxide isomerases. J. Lipid Med. Cell Signalling 12: 283-292.

3. Hamberg, M. 1987. Vanadium-catalyzed transformation of 13(S)-hydroperoxy-9(Z),11(E)-

octadecadienoic acid: structural studies on epoxy alcohols and trihydroxy acids. Chem. Phys. Lipids

43: 55-67.

4. Hamberg, M., and G. Hamberg. 1990. Hydroperoxide-dependent epoxidation of unsaturated

fatty acids in the broad bean (Vicia faba L.). Arch. Biochem. Biophys. 283: 409-416.

5. Blée, E., A.L. Wilcox, L.J. Marnett, and F. Schuber. 1993. Mechanism of reaction of fatty acid

hydroperoxides with soybean peroxygenase. J. Biol. Chem. 268: 1708-1715.

6. Hamberg, M., and U. Olsson. 2011. Efficient and specific conversion of 9-lipoxygenase

hydroperoxides in the beetroot. Formation of pinellic acid. Lipids 46: 873-878.

7. Gardner, H.W. 1989. Oxygen radical chemistry of polyunsaturated fatty acids. Free Rad. Biol.

& Med. 7: 65-86.

8. Dix, T.A. and L.J. Marnett. 1985. Conversion of linoleic acid hydroperoxide to hydroxy, keto,

epoxyhydroxy, and trihydroxy fatty acids by hematin. J. Biol. Chem. 260: 5351-5357.

9. Garssen, G.J., G.A. Veldink, J.F.G. Vliegenthart, and J. Boldingh. 1976. The formation of

threo-11-hydroxy-trans-12:13-epoxy-9-cis-octadecenoic acid by enzymatic isomerization of 13-L-

hydroperoxy-9-cis-11-trans-octadecadienoic acid by soybean lipoxygenase-1. Eur. J. Biochem. 62:

33-36.

10. Nigam, S., S. Patabhiraman, R. Ciccoli, G. Ishdorj, K. Schwarz, B. Petrucev, H. Kühn, and J.Z.

Haeggström. (2004). The rat leukocyte-type 12-lipoxygenase exhibits an intrinsic hepoxilin A3

synthase activity. J. Biol. Chem. 279: 29023-29030.

11. Yu, Z., C. Schneider, W.E. Boeglin, L.J. Marnett, and A.R. Brash. 2003. The lipoxygenase gene

ALOXE3 implicated in skin differentiation encodes a hydroperoxide isomerase. Proc. Natl. Acad. Sci.

USA 100: 9162-9167.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 16: Structural and mechanistical studies on hydroperoxide conversions

16

12. Jin, J., W.E. Boeglin, J.K. Cha, and A.R. Brash. 2012. 8R-Lipoxygenase-catalyzed synthesis of

a prominent cis-epoxyalcohol from dihomo-γ-linolenic acid: a distinctive transformation compared

with S-lipoxygenases. J. Lipid Res. 53: 292-299.

13. Lee, D.-S, P. Nioche, M. Hamberg, and C.S. Raman. 2008. Structural insights into the

evolutionary paths of oxylipin biosynthetic enzymes. Nature 455: 363-368.

14. Hamberg, M. 1992. A method for determination of the absolute stereochemistry of α,β-epoxy

alcohols derived from fatty acid hydroperoxides. Lipids 27: 1042-1046.

15. Hamberg, M. 1991. Regio- and stereochemical analysis of trihydroxyoctadecenoic acids derived

from linoleic acid 9- and 13-hydroperoxides. Lipids 26: 407-415.

16 Richardson, T.H., M-H. Hsu, T. Kronbach, H.J. Barnes, G. Chan, M.R. Waterman, B. Kemper,

and E.F. Johnson. 1993. Purification and characterization of recombinant-expressed cytochrome P450

2C3 from Escherichia coli: 2C3 encodes the 6β-hydroxylase deficient form of P450 3b. Arch.

Biochem. Biophys. 300: 510-516.

17. Scholz, J., F. Brodhun, E. Hornung, C. Herrfurth, M. Stumpe, A.K. Beike, B. Faltin, W. Frank,

R. Reski, and I. Feussner. 2012. Biosynthesis of allene oxides in Physcomitrella patens. BMC Plant

Biol. 12: 228.

18. Svetashev, V.I., L. Hefang, Y.F. Cai, S. Zuo-qing, C. Nian-hong, and J. Xin-ji. 1994. Fatty acid

composition and total lipid content in the lancelet Branchiostoma belcheri tsingtaoense (Tchang et

Koo). Comp. Biochem. Physiol. 108A: 325-329.

19. Gardner, H.W., K. Eskins, G.W. Grams, and G.E. Inglett. 1972. Radical addition of linoleic

hydroperoxides to α-tocopherol or the analogous hydroxychroman. Lipids 7: 324-333.

20. Hamberg, M., and B. Gotthammar. 1973. A new reaction of unsaturated fatty acid

hydroperoxides: formation of 11-hydroxy-12,13-epoxy-9-octadecenoic acid from 13-hydroperoxy-

9,11-octadecadienoic acid. Lipids 8: 737-743.

21. Chang, M.S., W.E. Boeglin, F.P. Guengerich, and A.R. Brash. 1996. Cytochrome P450-

dependent transformations of 15R- and 15S-hydroperoxyeicosatetraenoic acids: stereoselective

formation of epoxy alcohol products. Biochemistry 35: 464-471.

22. Song, W.-C., S.W. Baertschi, W.E. Boeglin, T.M. Harris, and A.R. Brash. 1993. Formation of

epoxyalcohols by a purified allene oxide synthase: implications for the mechanism of allene oxide

synthesis. J. Biol. Chem. 268: 6293-6298.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 17: Structural and mechanistical studies on hydroperoxide conversions

17

23. Gao, B., W.E. Boeglin, Y. Zheng, C. Schneider, and A.R. Brash. 2009. Evidence for an ionic

intermediate in the transformation of fatty acid hydroperoxide by a catalase-related allene oxide

synthase from the cyanobacterium Acaryochloris marina. J. Biol. Chem. 284: 22087-22098.

24. Hamberg, M. 1983. Autoxidation of linoleic acid: isolation and structure of four

dihydroxyoctadecadienoic acids. Biochim. Biophys. Acta 752: 353-356.

25. Niisuke, K., W.E. Boeglin, J.J. Murray, C. Schneider, and A.R. Brash. 2009. Biosynthesis of a

linoleic acid allylic epoxide: mechanistic comparison with its chemical synthesis and leukotriene A

biosynthesis. J. Lipid Res. 50: 1448-1455.

26. Choi, H., P.J. Proteau, T. Byrum, and W.H. Gerwick. 2012. Cymatherelactone and cymatherols

A-C, polycylic oxylipins from the marine brown alga Cymathere triplicata. Phytochemistry 73: 134-

141.

27. Grechkin A.N., L.S. Mukhtarova, and M. Hamberg. 2003. Detection of an enol intermediate in

the hydroperoxide lyase chain cleavage reaction. FEBS Lett. 2003. 549: 31-34.

28. Grechkin A.N., and M. Hamberg. 2004. The "heterolytic hydroperoxide lyase" is an isomerase

producing a short-lived fatty acid hemiacetal. Biochim. Biophys. Acta. 1636: 47-58.

29. Grechkin A.N., F. Brühlmann, L.S. Mukhtarova, Y.V. Gogolev, and M. Hamberg. 2006.

Hydroperoxide lyases (CYP74C and CYP74B) catalyze the homolytic isomerization of fatty acid

hydroperoxides into hemiacetals. Biochim. Biophys. Acta. 1761: 1419-1428.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 18: Structural and mechanistical studies on hydroperoxide conversions

18

ACKNOWLEDGMENTS

This work was supported by grant 12-04-01140-a from the Russian Foundation for Basic Research, a

grant from the Russian Academy of Sciences (program 'Molecular and Cell Biology'), a grant from the

program "Leading Scientific Schools", grant 14.740.11.0797 from the Federal Target Program (A.N.

Grechkin and B.I. Khairutdinov), and a grant from the Deutsche Forschungsgemeinschaft to Ivo

Feussner in the framework of the International Research Training Group (IRTG) 1422. Julia Scholz

was additionally supported by the Biomolecules program of the Göttingen Graduate School of

Neurosciences and Molecular Biology (GGNB).

The authors thank Gunvor Hamberg, Pia Meyer and Sabine Freitag for their skillful technical

assistance and Drs. Fahima Mukhitova and Anna Ogorodnikova for their help at separate stages of the

work.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 19: Structural and mechanistical studies on hydroperoxide conversions

19

TABLE 1. NMR spectral data for Compound 1 Me ester

Position number

13C chemical shifts (ppm); functional group

1H chemical shifts (ppm); multiplicity; coupling constant (Hz)

Heteronuclear multiple bond correlation

1 173.53; COOMe COOMe, H2, H3

2 34.24; CH2 2.11; t; 7.4 (H3) C3, C4, COO, COOMe, H3

3 25.36; CH2 1.54; m C2, C4, COO, COOMe, H2

4 29.51; CH2 1.15; m C2, C3, H2, H3

5 28.38-30.70; CH2 1.07-1.31; m

6 29.38; CH2 1.14; m H8a, H8b

7 30.01; CH2 1.21; m C9, H8a, H8b, H9

8a 28.48; CH2 1.92; m C6, C7, C9, H8b, H9

8b 2.01; m C7, C9,C10, H7, H8a

9 134.03; CH 5.44; dt; 11.0 (H10); 6.9 (H8a,b)

C7, C8, C10, H8a, H8b,

10 128.46; CH 5.49; ddt; 8.5 (H11); 1.2 (H8a,b)

C8, C9, C11, C12, H8a, H8b, H9, H11

11 66.77; CH 4.28; dd; 7.7 (H12) C10, C12, C13, H10, H12, OH,

12 60.74; CH 2.91; dd; 4.3 (H13) C10, C11, c13, C14, H10, H11, H13, H14

13 57.82; CH 2.74; ddd; 7.3 (H14a); 4.9 (H14b)

C14, C15, H11, H12, H13, H14, H15

14 29.09; CH2 1.42; m, AB C12, C13, C15, C16, H12, H13, H15a

15a 27.08; CH2 1.28; m C13, C14, C16, H13, H14, H17

15b 1.38; m C13, C14, C16, H13, H14, H17

16 32.16; CH2 1.23; m C16, C18, H14, H15, H18

17 23.08; CH2 1.21; m C15, C16, H15a, H15b, H18

18 14.39; CH3 0.87; t; 7.1 (H17) C16, C17, H16, H17

(1) 51.14; COOMe 3.37; s C2, COO, H2

(11) OH 1.80; d H11

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 20: Structural and mechanistical studies on hydroperoxide conversions

20

TABLE 2. 18O incorporation from [18O2-hydroperoxy]13(S)-HPODE into Compound 1 in the presence of BfEAS

A. GC-MS (SIM) quantification of Compound 1 species possessing 2, 1 or 0 atoms of 18O (M+ at m/z 402.4, 400.4 and 398.4, respectively).

Ion, m/z Integral intensities of molecular ion M+ peaks in SIM GC-MS chromatograms of 18O labelled and unlabelled Compound 1

BfEAS + [18O2-hydroperoxy]13-HPODE

BfEAS + unlabelled 13-HPODE

398.4 2.64±0.38 93.13±0.88

400.4 2.37±0.19* 6.72±0.56

402.4 94.93±0.56* 0.30±0.03

*Presented relative intensities of ions (m/z 400.4 and 402.4) in spectrum of 18O labelled product were corrected in accordance with the relative abundance of the same ions in the spectrum of unlabelled product.

B. The total isotopic content of 18O in Compound 1 and the precursor [18O2]13-HPOD

Total estimated 18O content in product 1 (100% corresponds to two 18O atoms incorporated)

Control (18O content in [18O2-hydroperoxy]13-HPODE)

96.12±0.58 96.98 ± 0.22%**

**[18O2]13-HPODE was successively reduced with sodium borohydride, methylated with diazomethane and trimethylsilylated. The resulting [18O]13-HODE was subjected to SIM GC-MS analyses. 18O content was estimated from the relative abundance of ion pairs 384.4 and 382.4, [M]+; 313.3 and 311.3 [M-C5H11]+; 227.2 and 225.2 [M-(CH2)8COOMe]+.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 21: Structural and mechanistical studies on hydroperoxide conversions

21

Figure legends

Figure 1: General mechanisms in the formation of CYP74 products illustrated with the

hydroperoxide 13(S)-HPODE as substrate. AOS, allene oxide synthase; DES, divinyl ether

synthase; HAS, hemiacetal synthase (earlier referred to as hydroperoxide lyase). The allene oxide and

hemiacetal products are both unstable and spontaneously converted to ketols and short chain

aldehydes, respectively.

Figure 2: SDS-PAGE analysis of BfEAS purification. BfEAS was heterologously expressed in E.

coli Bl21 star at 16°C for 3d and harvested by centrifugation. After cell disruption via sonification the

cell free extract was applied on a Ni-NTA column. Protein elution was performed with a linear

gradient of increasing imidazol concentration. Shown are the following fractions: unbound protein

(FT), washed proteins (W1, W2) and eluted proteins.

Figure 3: Spectroscopic analysis of recombinant BfEAS. Shown is a typical UV/vis-spectrum

(upper) and a CD-spectrum (lower) of purified BfEAS in sodium phosphate buffer (50 mM, pH 8.0).

The UV/vis-spectrum shows two major absorption maxima at 281 nm and 420 nm (γ, soret-band)

typical for heme proteins. Additionally three smaller maxima are present at 368 nm (δ), 539 nm (β)

and 565 nm (α). The CD-spectrum shows major minima at approx. 215 nm and 225 nm that are

characteristic for α-helical proteins.

Figure 4: GC-MS analyses of products (Me/TMS) formed from 13-hydroperoxides incubated

with BfEAS. A. GC-MS chromatogram of 13(S)-HPODE products. B. Mass spectrum of Compound

1; insert: fragmentation scheme. C. Mass spectrum of Compound 1 biosynthesized from [18O2-

hydroperoxy]13(S)-HPODE; insert: fragmentation scheme. D. GC-MS chromatogram of 13(S)-

HPOTrE products. "1", 12,13-epoxy-11-hydroxy-9-octadecenoic acid; "2", 12,13-epoxy-11-hydroxy-

9,15-octadecadienoic acid; "12-oxo-PDA", cis-12-oxo-10,15-phytodienoic acid; "α-ketol", 13-

hydroxy-12-oxo-9,15-octadecadienoic acid.

Figure 5: Structures of BfEAS products.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 22: Structural and mechanistical studies on hydroperoxide conversions

22

Figure 6: Chemical transformations carried out on Compound 1. TPP=Se, triphenylphosphine

selenide; TFA, trifluoroacetic acid; MCCl, (-)-menthoxycarbonyl chloride.

Figure 7: GC-MS analyses of products (Me/TMS) formed from 9(S)-HPODE incubated with

BfEAS. A. GC-MS chromatogram of 9(S)-HPODE products. B. Mass spectrum of Compound 3;

insert: fragmentation scheme. C. Mass spectrum of Compound 4; insert: fragmentation scheme. D.

Mass spectrum of Compound 5; insert: fragmentation scheme. "3", 9,10-epoxy-11-octadecen-13-olide;

"4", 9,10-epoxy-11-hydroxy-12-octadecenoic acid; "5", 9,14-dihydroxy-10,12-octadecadienoic acid.

Figure 8: NMR data for Compound 3 produced from 9(S)-HPODE upon incubation with

BfEAS. A. 2D-COSY spectrum. Top and left projections: 1H-NMR spectrum. Insert: structure of

Compound 3 with specified carbon chain numbering. B. 1H-13C-HSQC spectrum. Left projection: 13C-

NMR spectrum.

Figure 9. Chemical transformations carried out on Compound 3. The trihydroxyoctadecenoates

"isomer a" and "isomer b" were assigned the 9(S),10(S),13(S) and 9(S),12(R),13(S) absolute

configurations, respectively, by GC-MS analysis using authentic standards as described in ref. 15.

Figure 10. Reactions and mechanisms in the formation of products catalyzed by BfEAS. 13(S)-

HPODE (precursor of Compound 1): R = -(CH2)4CH3, R’ = -(CH2)6COOH; 13(S)-HPOTrE (precursor

of Compound 2): R = -CH2CH=CH-CH2CH3; R’ = -(CH2)6COOH; 9(S)-HPODE (precursor of

Compounds 4, 5 and 3): R = -(CH2)7COOH; R’ = -(CH2)3CH3.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 23: Structural and mechanistical studies on hydroperoxide conversions

23

Figure 1.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 24: Structural and mechanistical studies on hydroperoxide conversions

24

Figure 2.

18 mM 240 mM

kDa 116.0

66.2

45.0

◄ BfEAS

35.0

FT W1 W2 MW

Elution (increasing imidazol conc.)

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 25: Structural and mechanistical studies on hydroperoxide conversions

25

Figure 3.

190 200 210 220 230 240 250 260

-6

-4

-2

0

2

4

6

8

CD [m

deg]

wavelength [nm]

300 400 500 6000.0

0.2

0.4

0.6

0.8

1.0ab

sorp

tion

wavelength [nm]

281 nm

420 nm

539 nm 565 nm

368 nm

wavelength [nm]

abso

rptio

nCD

[mde

g]

wavelength [nm]

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 26: Structural and mechanistical studies on hydroperoxide conversions

26

Figure 4.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 27: Structural and mechanistical studies on hydroperoxide conversions

27

Figure 5.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 28: Structural and mechanistical studies on hydroperoxide conversions

28

Figure 6.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 29: Structural and mechanistical studies on hydroperoxide conversions

29

Figure 7.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 30: Structural and mechanistical studies on hydroperoxide conversions

30

Figure 8.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 31: Structural and mechanistical studies on hydroperoxide conversions

31

Figure 9.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from

Page 32: Structural and mechanistical studies on hydroperoxide conversions

32

Figure 10.

by guest, on April 4, 2018

ww

w.jlr.org

Dow

nloaded from