17
The Crystal Structure of Arabidopsis thaliana Allene Oxide Cyclase: Insights into the Oxylipin Cyclization Reaction W Eckhard Hofmann, a,1,2 Philipp Zerbe, b and Florian Schaller b,1,2 a Lehrstuhl fu ¨ r Biophysik, Ruhr-Universita ¨ t Bochum, D-44780 Bochum, Germany b Lehrstuhl fu ¨ r Pflanzenphysiologie, Ruhr-Universita ¨ t Bochum, D-44780 Bochum, Germany We describe the crystallization and structure elucidation of Arabidopsis thaliana allene oxide cyclase 2 (AOC2), a key enzyme in the biosynthesis of jasmonates. In a coupled reaction with allene oxide synthase, AOC2 releases the first cyclic and bio- logically active metabolite, 12-oxo-phytodienoic acid (OPDA). AOC2 (AT3G25770) folds into an eight-stranded antiparallel b-barrel with a C-terminal partial helical extension. The protein forms a hydrophobic binding cavity with two distinct polar patches. AOC2 is trimeric in crystals, in vitro and in planta. Based on the observed folding pattern, we assigned AOC2 as a low molecular weight member of the lipocalin family with enzymatic activity in plants. We determined the binding position of the competitive inhibitor vernolic acid (a substrate analog) in the binding pocket. Based on models for bound substrate 12,13- epoxy-9,11,15-octadecatrienoic acid and product OPDA, we propose a reaction scheme that explains the influence of the C15 double bond on reactivity. Reaction is promoted by anchimeric assistance through a conserved Glu residue. The transition state with a pentadienyl carbocation and an oxyanion is stabilized by a strongly bound water molecule and favorable pp interactions with aromatic residues in the cavity. Stereoselectivity results from steric restrictions to the necessary substrate isomerizations imposed by the protein. INTRODUCTION Since the initial discovery of methyl jasmonate (MeJA) as a secondary metabolite in essential oils of jasmine in 1962 (Demole et al., 1962), JAs have become accepted as a new class of plant hormone. In the early 1980s, their widespread occurrence throughout the plant kingdom (Meyer et al., 1984) and their growth-inhibitory (Dathe et al., 1981) and senescence-promoting activities (Ueda and Kato, 1980) were established. It has become increasingly clear, however, that biological activity is not limited to JA but extends to, and even differs among, its many metab- olites and conjugates as well as its cyclopentenone precursors (Kramell et al., 1997; Stintzi et al., 2001). Because of the different biological function of each member of the group of JAs, the enzymes of JA biosynthesis and metabolism may thus have a regulatory function in controlling the activity and relative levels of different signaling molecules. Research in recent years generally confirmed the Vick and Zimmerman pathway of JA biosynthesis (the octadecanoid path- way) and made considerable progress with respect to the bio- chemistry of the enzymes involved as well as the molecular organization and regulation of the pathway. The early elucidation of the JA biosynthetic pathway and the demonstration of the growth-inhibiting and senescence-promoting activity were followed by the discovery that JAs are involved in plant defense reactions, which greatly stimulated the interest in these compounds as plant signaling molecules. A role in plant defense was first shown by Farmer and Ryan, who demonstrated the induction of proteinase inhibitors by MeJA and JA as part of the defense response against herbivorous insects (Farmer and Ryan, 1990; Farmer et al., 1991). JAs were then shown to be active inducers of antimicrobial phytoalexins by Gundlach et al. (1992), and subsequent work clearly established their defense gene–inducing activity for induced resistance against insect pred- ators and pathogens (for a review, see Reymond and Farmer, 1998; Weiler et al., 1998; Ble ´ e, 2002). Such a role was unequiv- ocally confirmed by the analysis of mutants compromised in either the synthesis or the perception of JA signals (for a review, see Schaller et al., 2005). The pathway of JA biosynthesis is shown in Figure 1. Biosyn- thesis is believed to start with the oxygenation of free a-linolenic acid (LA), which is converted to (9Z,11E,15Z,13S)-13-hydroper- oxy-9,11,15-octadecatrienoic acid [13(S)-HPOT] in a reaction catalyzed by 13-lipoxygenase. The presumed release of LA from membrane lipids through the action of a lipase may be triggered by local or systemic signals, such as oligogalacturonides, chitosan, systemin (Farmer and Ryan, 1992; Mueller et al., 1993; Narva ´ ez- Va ´ squez et al., 1999), or wounding (Conconi et al., 1996; Narva ´ ez-Va ´ squez et al., 1999). The 13(S)-HPOT serves as a substrate for several enzymes (Figure 1), such as divinylether synthase, lipoxygenase, hydroperoxide reductase, epoxyalco- holsynthase, peroxygenase, hydroperoxide lyase, or allene ox- ide synthase (AOS) (Ble ´e and Joyard, 1996; Vick and Zimmerman, 1981, 1987). AOS converts 13(S)-HPOT to the 1 These authors contributed equally to this work. 2 To whom correspondence should be addressed. E-mail eckhard. [email protected] or fl[email protected]; fax 49-234-32-14748 or 49-234-32-14187. The authors responsible for distribution of materials integral to the findings presented in this article in accordance with the policy described in the Instructions for Authors (www.plantcell.org) are: Eckhard Hofmann ([email protected]) and Florian Schaller (florian. [email protected]). W Online version contains Web-only data. www.plantcell.org/cgi/doi/10.1105/tpc.106.043984 The Plant Cell, Vol. 18, 3201–3217, November 2006, www.plantcell.org ª 2006 American Society of Plant Biologists Downloaded from https://academic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

The Crystal Structure of Arabidopsis thaliana Allene Oxide ...which produces both allene oxides from the respective 13(S)-hydroperoxy fatty acids (18:3 and 18:2, respectively). It

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

  • The Crystal Structure of Arabidopsis thaliana Allene OxideCyclase: Insights into the Oxylipin Cyclization Reaction W

    Eckhard Hofmann,a,1,2 Philipp Zerbe,b and Florian Schallerb,1,2

    a Lehrstuhl für Biophysik, Ruhr-Universität Bochum, D-44780 Bochum, Germanyb Lehrstuhl für Pflanzenphysiologie, Ruhr-Universität Bochum, D-44780 Bochum, Germany

    We describe the crystallization and structure elucidation of Arabidopsis thaliana allene oxide cyclase 2 (AOC2), a key enzyme

    in the biosynthesis of jasmonates. In a coupled reaction with allene oxide synthase, AOC2 releases the first cyclic and bio-

    logically active metabolite, 12-oxo-phytodienoic acid (OPDA). AOC2 (AT3G25770) folds into an eight-stranded antiparallel

    b-barrel with a C-terminal partial helical extension. The protein forms a hydrophobic binding cavity with two distinct polar

    patches. AOC2 is trimeric in crystals, in vitro and in planta. Based on the observed folding pattern, we assigned AOC2 as a low

    molecular weight member of the lipocalin family with enzymatic activity in plants. We determined the binding position of the

    competitive inhibitor vernolic acid (a substrate analog) in the binding pocket. Based on models for bound substrate 12,13-

    epoxy-9,11,15-octadecatrienoic acid and product OPDA, we propose a reaction scheme that explains the influence of the C15

    double bond on reactivity. Reaction is promoted by anchimeric assistance through a conserved Glu residue. The transition

    state with a pentadienyl carbocation and an oxyanion is stabilized by a strongly bound water molecule and favorable p–p

    interactions with aromatic residues in the cavity. Stereoselectivity results from steric restrictions to the necessary substrate

    isomerizations imposed by the protein.

    INTRODUCTION

    Since the initial discovery of methyl jasmonate (MeJA) as a

    secondary metabolite in essential oils of jasmine in 1962 (Demole

    et al., 1962), JAs have become accepted as a new class of plant

    hormone. In the early 1980s, their widespread occurrence

    throughout the plant kingdom (Meyer et al., 1984) and their

    growth-inhibitory (Dathe et al., 1981) and senescence-promoting

    activities (Ueda and Kato, 1980) were established. It has become

    increasingly clear, however, that biological activity is not limited

    to JA but extends to, and even differs among, its many metab-

    olites and conjugates as well as its cyclopentenone precursors

    (Kramell et al., 1997; Stintzi et al., 2001). Because of the different

    biological function of each member of the group of JAs, the

    enzymes of JA biosynthesis and metabolism may thus have a

    regulatory function in controlling the activity and relative levels of

    different signaling molecules.

    Research in recent years generally confirmed the Vick and

    Zimmerman pathway of JA biosynthesis (the octadecanoid path-

    way) and made considerable progress with respect to the bio-

    chemistry of the enzymes involved as well as the molecular

    organization and regulation of the pathway.

    The early elucidation of the JA biosynthetic pathway and the

    demonstration of the growth-inhibiting and senescence-promoting

    activity were followed by the discovery that JAs are involved in

    plant defense reactions, which greatly stimulated the interest in

    these compounds as plant signaling molecules. A role in plant

    defense was first shown by Farmer and Ryan, who demonstrated

    the induction of proteinase inhibitors by MeJA and JA as part of

    the defense response against herbivorous insects (Farmer and

    Ryan, 1990; Farmer et al., 1991). JAs were then shown to be

    active inducers of antimicrobial phytoalexins by Gundlach et al.

    (1992), and subsequent work clearly established their defense

    gene–inducing activity for induced resistance against insect pred-

    ators and pathogens (for a review, see Reymond and Farmer,

    1998; Weiler et al., 1998; Blée, 2002). Such a role was unequiv-

    ocally confirmed by the analysis of mutants compromised in

    either the synthesis or the perception of JA signals (for a review,

    see Schaller et al., 2005).

    The pathway of JA biosynthesis is shown in Figure 1. Biosyn-

    thesis is believed to start with the oxygenation of free a-linolenic

    acid (LA), which is converted to (9Z,11E,15Z,13S)-13-hydroper-

    oxy-9,11,15-octadecatrienoic acid [13(S)-HPOT] in a reaction

    catalyzed by 13-lipoxygenase. The presumed release of LA from

    membrane lipids through the action of a lipase may be triggered

    by local or systemic signals, such as oligogalacturonides, chitosan,

    systemin (Farmer and Ryan, 1992; Mueller et al., 1993; Narváez-

    Vásquez et al., 1999), or wounding (Conconi et al., 1996;

    Narváez-Vásquez et al., 1999). The 13(S)-HPOT serves as a

    substrate for several enzymes (Figure 1), such as divinylether

    synthase, lipoxygenase, hydroperoxide reductase, epoxyalco-

    holsynthase, peroxygenase, hydroperoxide lyase, or allene ox-

    ide synthase (AOS) (Blée and Joyard, 1996; Vick and

    Zimmerman, 1981, 1987). AOS converts 13(S)-HPOT to the

    1 These authors contributed equally to this work.2 To whom correspondence should be addressed. E-mail [email protected] or [email protected]; fax 49-234-32-14748or 49-234-32-14187.The authors responsible for distribution of materials integral to thefindings presented in this article in accordance with the policy describedin the Instructions for Authors (www.plantcell.org) are: EckhardHofmann ([email protected]) and Florian Schaller ([email protected]).W Online version contains Web-only data.www.plantcell.org/cgi/doi/10.1105/tpc.106.043984

    The Plant Cell, Vol. 18, 3201–3217, November 2006, www.plantcell.org ª 2006 American Society of Plant Biologists

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • unstable epoxide 12,13(S)-epoxy-9(Z),11,15(Z)-octadecatrienoic

    acid (12,13-EOT), which is cyclized by allene oxide cyclase

    (AOC) to the first cyclic and biologically active compound of the

    pathway, 12-oxo-phytodienoic acid (OPDA). Reduction of the

    10,11-double bond by an NADPH-dependent OPDA reductase

    then yields 3-oxo-2(29(Z)-pentenyl)-cyclopentane-1-octanoicacid (OPC-8:0), which is believed to undergo three cycles of

    b-oxidation to yield the end product of the pathway [i.e., JA with

    (3R,7S)-configuration; (þ)-7-iso-JA] (Vick and Zimmerman,1984). The biosynthesis of JA appears to involve two different

    compartments: the conversion of LA to OPDA is localized in the

    chloroplasts (Vick and Zimmerman, 1987; Song et al., 1993;

    Laudert et al., 1997), while the reduction of OPDA to OPC-8:0

    (Schaller and Weiler, 1997a, 1997b; Schaller et al., 2000;

    Strassner et al., 2002) and the three steps of b-oxidation (i.e.,

    conversion of OPC-8:0 to JA) occur in peroxisomes (Gerhardt,

    1983; Vick and Zimmerman, 1984; Li et al., 2005). Like JA,

    OPDA may accumulate to substantial amounts in plant tissue

    (Stelmach et al., 1998). By contrast, OPC-8:0 occurs only in trace

    amounts.

    Vick and Zimmerman (1981) postulated the existence of a

    hydroperoxide cyclase involved in the conversion of 13(S)-HPOT

    into OPDA. The allene oxide 12,13-EOT was shown to serve as

    the immediate precursor of OPDA in plants (Baertschi et al.,

    1988; Crombie and Morgan, 1988; Hamberg and Hughes, 1988).

    Hamberg and Fahlstadius (1990) showed that this cyclization

    reaction is catalyzed by AOC, a soluble enzyme in maize (Zea

    mays). AOC was subsequently purified as an apparent dimer of

    47 kD from maize kernels (Ziegler et al., 1997) and characterized

    with respect to its substrate specificity: the enzyme accepted

    12,13(S)-epoxylinolenic acid but not 12,13(S)-epoxylinoleic acid

    as a substrate (Ziegler et al., 1999). This is in contrast with AOS,

    which produces both allene oxides from the respective 13(S)-

    hydroperoxy fatty acids (18:3 and 18:2, respectively). It thus

    appears that AOC confers additional specificity to the octade-

    canoid biosynthetic pathway.

    An interesting aspect of the AOC reaction is the apparent

    competition between the spontaneous decomposition of its

    substrate, the unstable allene oxide, to form a- and g-ketols and

    racemic cis-OPDA, and the enzyme-catalyzed formation of op-

    tically pure (9S,13S)-OPDA [i.e., cis(þ)-OPDA], the first pathwayintermediate having the characteristic pentacyclic ring structure

    of JAs (Figure 1) and biological activity. The extremely short half-

    life of allene oxides (half-time

  • biochemical and structural studies on AOC2 because this iso-

    zyme has been shown to be highly expressed in planta and was

    described as being the most active of the four Arabidopsis AOCs

    (Stenzel et al., 2003b).

    RESULTS

    Overexpression, Purification, and Crystallization

    of Arabidopsis AOC2

    Because of relatively weak expression levels of published con-

    structs (C. Wasternack, personal communication) we cloned

    differently truncated versions of AOC2 in pET 21b(þ) (Novagen/Merck His6-tag, C-terminally) and pQE30 (Qiagen His6-tag,

    N-terminally) and analyzed expression levels of insoluble and

    soluble AOC2. Constructs beginning at base 132 showed the

    highest expression of soluble protein in both vectors and were

    subsequently used.

    Using SDS-PAGE analysis, we observed a strong influence of

    the His6-tag location on the oligomerization state of the protein.

    While AOC2 with a C-terminal His6-tag runs as a monomer with

    an apparent mass of ;22 kD, the N-terminally tagged proteinshows a dominant band at a molecular mass of ;60 kD, con-sistent with an SDS-stable trimer formation. The identity of the

    bands could be confirmed by immunoblot analyses using anti-

    bodies raised against AOC2 or the His6-tag.

    A similar pattern to that of the N-terminally tagged protein has

    been observed in immunoblot analyses of native AOCs from plant

    extracts (P. Zerbe and F. Schaller, unpublished data). This clearly

    shows that the trimeric form is present in planta as well and is not

    an artifact due to addition of the N-terminal His6-tag. Addition of

    a C-terminal His6-tag seems to destabilize the trimers.

    Overall Structure

    The structure of Arabidopsis AOC2 has been determined by

    multiwavelength anomalous dispersion with selenomethionine

    (SeMet)-labeled protein crystals of the space group P212121 and

    was refined to a resolution of 1.5 Å (SeMet-AOC2). Additionally,

    we solved the structure of the unlabeled protein in space group

    P21 at a resolution of 1.8 Å (WT-AOC2). In both space groups,

    AOC2 forms closely packed trimers. In space group P21, we find

    two copies of this AOC2 trimer, which superimpose with a root

    mean square deviation (RMSD) of 0.33 Å for all Ca atoms. Both

    align well with the single trimer found in the SeMet-AOC2 struc-

    ture with RMSDs of 0.40 and 0.44 Å, respectively. Data collection

    and refinement statistics are presented in Table 1.

    Figure 2 shows the structure of SeMet-AOC2 as a ribbon

    diagram together with a topology diagram. The prominent struc-

    tural feature of AOC2 is the eight-stranded antiparallel b-barrel of

    ;40-Å height formed by residues 17 to 147 (numbered accord-ing to the recombinantly expressed protein). Based on the offset

    observed in the hydrogen bonding pattern upon closure of the

    barrel, we can assign a shear number of 10, which is a measure

    for the tilt of the b-strands relative to the barrel axis (Murzin,

    1994). The central cross section is slightly elliptical with axes of

    ;14 and 18 Å (Figure 2B). The 41 C-terminal residues cover the

    bottom and one side of the barrel with two one-turn 3/10 helices

    and a short a-helix connected by random coil stretches.

    The starting residue of the first b-strand (Glu-17, or Glu-82 in

    the unprocessed protein) corresponds to the predicted cleavage

    site of the transit peptide (ChloroP; Emanuelsson et al., 1999). In

    our construct, we included five additional residues, two of which

    are well defined in the density. The first 14 residues, including the

    N-terminal His6-tag and some vector-specific residues are dis-

    ordered in the crystal structure.

    The barrel forms an elongated cavity, which is lined mostly by

    aromatic and hydrophobic residues and reaches ;14 Å into theprotein (Figure 2C). Three areas of the cavity surface are note-

    worthy. First, the conserved Glu-23 is positioned at the very

    bottom of the cavity, introducing a negative charge. Second, an

    additional polar patch is formed by Ser-31, Asn-25, and Asn-53

    on one side of the cavity. These three residues together with

    the main chain nitrogen from Pro-32 coordinate a tightly bound

    water molecule that is found in all our AOC2 structures. Third, we

    found Cys-71 on the opposing wall of the cavity. These residues

    are strictly conserved among all AOC sequences in the EBI

    UNIRef100 database (Figure 3).

    Quarternary Structure (Trimer)

    In complete agreement with our biochemical data, we found

    AOC2 to form trimers when crystallized (Figure 4). The barrel

    axes of the monomers are tilted ;308 with respect to the trimeraxis and pack closely with their barrel walls. Trimerization buries

    ;2000 A2 of the barrel surface and is the only interaction in thecrystal that is identical for all copies of the molecule. Crystal

    contacts cover a significantly smaller surface (800 to 900 A2 per

    monomer) and are not identical for all monomers. Identical

    trimerization is observed in both the WT-AOC2 and SeMet-

    AOC2 crystals, which have different space groups and packing

    interactions. Residues involved in trimerization are highly con-

    served in known AOCs (see Supplemental Figure 1 online). The

    majority of the C-terminal part of the protein is not involved in the

    interaction of the monomers (Figure 4), but the C terminus (Asn-

    188) together with Arg-29 forms a strong salt bridge to Glu-75 of

    a neighboring monomer.

    Structural Comparison with Other Proteins

    According to the rules underlying the SCOP database (Murzin

    et al., 1995), the AOC2 structure with an eight-stranded strictly

    antiparallel b-barrel with a shear number of 10 would be either

    assigned to the streptavidin fold or to a new fold due to the to-

    pological differences in the orientation of the barrel. For func-

    tional analysis, it is more fruitful to discuss its relation to protein

    families. AOC2 certainly belongs to the superfamily of calycins

    (Flower et al., 2000), which are characterized by a central

    b-barrel forming a hydrophobic cavity for binding of small lipo-

    philic compounds. Major members of this superfamily are the

    triabins, metalloproteinase inhibitors, the fatty acid binding pro-

    teins, the avidins, and the lipocalins. Of these, the lipocalins most

    closely resemble the tertiary structure and fold of AOC2. Indeed,

    a DALI (Holm and Sander, 1994) structural similarity search

    clearly assigns AOC to the lipocalin fold, finding as the top hits

    Crystal Structure of AOC2 3203

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • retinol binding protein (RBP; 1AQB) and nitrophorin (1NP1) each

    with a Z-score of 3.2.

    The lipocalin protein family (Pervaiz and Brew, 1985, 1987;

    Grzyb et al., 2006) is a large group of small extracellular proteins

    with great diversity at the sequence level but mostly with either

    three characteristic short conserved sequence motifs (SCR I-III;

    the kernel lipocalins) or else one conserved motif (SCR I; the

    outlier lipocalins). However, the crystal structures are highly

    conserved and comprise a single eight-stranded continuously

    hydrogen-bonded antiparallel b-barrel with an integral ligand

    binding site (Figure 5). The structural basis for these SCRs is a

    clustering of conserved residues around a central Trp in SCR I at

    the beginning of b-strand A, which stabilizes barrel closure. Sev-

    eral members of the lipocalin family bind to specific cell surface

    receptors or form complexes with soluble macromolecules. Ini-

    tially, lipocalins were described as transport proteins; however, it

    has become clear that the lipocalins exhibit a multitude of func-

    tions, with roles in retinol transport, olfaction, and pheromone

    transport, prostaglandin biosynthesis, invertebrate cryptic col-

    oration, and a role in the xanthophyll cycle. In addition to the

    central b-barrel, the C-terminal part of the protein is usually ap-

    pended by a characteristic a-helix. One end of the barrel is

    closed by the N terminus that traverses the base of the barrel

    (Figure 5). At the opposite end, the b-barrel is typically open to

    solvent and provides access to the cavity. These binding sites of

    the lipocalins may have very different shapes (e.g., with a ligand

    pocket like an extended cave with lobes, a ligand pocket deep

    within the hydrophobic core, with loops encapsulating the ligand,

    or a binding site with a wide, funnel-like opening to the solvent)

    (for a review, see Schlehuber and Skerra, 2005).

    Automated superposition of AOC2 with the two lipocalin hits

    from the DALI search leads to RMSD values of 2.6 Å for 74

    matched Ca for 1AQB and 3.6 Å for 100 matched Ca for 1NP1.

    The superposition with RBP as an archetypal representative of

    the lipocalin family is shown in Figure 5.

    Stretches of superposed residues all fall within the b-strands,

    therefore aligning the barrel walls well, while the loop regions

    show large differences. AOC2 does not show the classical

    C-terminal helix, and the barrel is not closed by the preceding

    sequence but by the extended C terminus. Most striking is the

    Table 1. Data Collection and Refinement Statisticsa

    Data Set Peak

    Inflection

    Point Remote Inhibitor Wild Type

    Beamline ESRF ID13.1 In house SLS PXII

    Resolution (Å) 37–1.5 (1.6–1.5) 38–1.7 (1.8–1.7) 50–1.8 (2.1–1.8)

    Cell parameters (a, b, c; Å)

    (a, b, g; 8)

    64.48, 99.95, 106.15

    90 90 90

    64.5, 99.8, 105.9

    90 90 90

    67.12, 104.59, 86.96

    90 95.0 90

    Space group P212121 P212121 P21Wavelength 0.97935 0.97955 0.9789 1.54 0.950

    Completeness (%) 95.9 (94.2) 95.9 (94.2) 95.9 (94.2) 93.5 (89.8) 97.7 (93.8)

    Multiplicity 7.8 (7.8) 7.8 (7.8) 7.8 (7.8) 4.2 (4.2) 4.3 (3.4)

    Average I/sI 19.3 (5.3) 19.3 (5.3) 18.6 (4.9) 15.2 (4.2) 11.3 (4.8)

    Rsym (%) 7.0 (38.8) 7.0 (39.1) 7.4 (42.7) 7.9 (34.5) 7.4 (22.0)

    Rmeas (%)b 7.5 (41.6) 7.5 (41.9) 7.9 (45.8) 9.1 (39.5) 8.3 (25.8)

    Rmrgd-F (%)b 7.0 (26.1) 7.0 (26.1) 7.6 (29.0) 9.5 (24.2) 8.6 (28.2)

    Phasing

    FOM (RESOLVE)c 0.86

    Refinement

    Resolution (Å) 37–1.5

    (1.54–1.5)

    73–1.7 (1.74–1.7) 86–1.8 (1.85–1.8)

    Rcryst (%) 17.3 (18.8) 19.8 (29.7) 20.7 (43.4)

    Rfree (%)d 19.1 (21.8) 23.3 (36.7) 24.8 (51.2)

    Number of atoms

    Protein 4194 4124 8179

    Inhibitor 21

    Glycerol 42 30 30

    Water/ion 556 369 660/2

    Number of side chains with

    alternate conformations

    19 6 5

    Average B-factor protein (inhibitor) 13.5 12.8 (48.2) 23.8

    RMSD from ideality (protein atoms)

    Bonds (Å) 0.008 0.021 0.022

    Angles (8) 1.224 1.838 1.757

    a Data in parentheses represent values in the highest resolution bin.b For definition of Rmeas and Rmrgd-F, see Diederichs and Karplus (1997).c FOM calculated after density modification and threefold averaging in RESOLVE.d Rfree calculated from 5% of data omitted from refinement.

    3204 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • different position of the termini with respect to the barrel orien-

    tation. In the orientation shown in Figure 5, the origin of the closed

    b-sheet is on the top for the lipocalins, whereas it is on the bottom

    side for AOC2. This topological difference could be the result of a

    strand exchange in which either the first or last b-strand (strand A

    or strand H in lipocalin nomenclature) has been replaced by C- or

    N-terminal sequence stretches, respectively. Notably, the se-

    quence DLVPFTNKLY (residues 47 to 56) from AOC2 could be

    automatically aligned to SCR I in pig RPB (accession number

    1AQB). While we can observe a weak sequence similarity in this

    stretch, the conserved Gly and Trp (residues 22 and 24, respec-

    tively, in RBP) are missing in the AOC sequence. Interestingly, the

    region identified in AOC2 falls within the second b-strand and

    superimposes spatially with the location of SCR I in b-strand A of

    the lipocalins (Figure 5).

    Recently, the structure of a coral AOS domain has been

    reported (Oldham et al., 2005). In soft corals, no AOC has been

    described so far, so an intriguing possibility would be that coral

    AOS catalyzes both epoxidation and cyclization reactions. In-

    deed, coral AOS has been found to have a catalase-type fold,

    which includes an eight-stranded antiparallel barrel. However,

    there has been no report that the barrel fulfils a catalytic role in

    cyclization in this case, and the interior is tightly packed with

    protein side chains, preventing a similar substrate binding as in

    lipocalins.

    Substrate Specificity

    To gain more information about the catalytic mechanism, we

    tested 13(S)-HPOT and 13(S)-hydroperoxy-octadecadienoic acid

    (13(S)-HPOD) as substrates for Arabidopsis AOC2 and AOS. We

    found that with 13(S)-HPOD as substrate, only negligible amounts

    of cyclic product were formed (data not shown). Since AOS

    makes use of linoleic acid and linolenic acid derivatives (Siedow,

    1991; Song and Brash, 1991; Pan et al., 1995; Laudert et al., 1996),

    AOC2 is apparently incapable of cyclizing the corresponding

    Figure 2. Structure of the AOC2 Monomer.

    (A) Side view of the AOC2 monomer shown in ribbon representation, color changing from blue (N terminus) to red (C terminus).

    (B) Top view, rotated by 908 toward the viewer with respect to (A). The same representation is used as in (A).

    (C) Molecular surface representation of the proposed binding pocket. The orientation and color scheme are similar to (A). Also shown are the residues

    discussed in the text in stick representation. Water 75 is shown as a red sphere.

    (D) Topology diagram in the same color coding as in (A). b-Strands are shown as arrows (with numbers indicating the first and last residues) and helices

    as small cylinders.

    Crystal Structure of AOC2 3205

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • epoxide, which was indicated before by studies of Ziegler et al.

    (1997) on AOC from maize seeds. Therefore, the double bond at

    position 15 seems to be essential for the enantioselective cycli-

    zation by AOC2. This is in agreement with a proposal for the

    chemical cyclization reaction, whereby the 15(Z) double bond

    facilitates the opening of the oxirane ring of the allene oxide, and

    the resulting C-13 carbocation is stabilized by the p electrons of

    the 15(Z) double bond (Grechkin, 1994).

    Substrate Binding Site and Biochemical Analysis of

    Mutated AOC2 Protein

    In 1990, a competitive inhibitor of the AOC of maize [(þ/�)-cis-12,13-epoxy-9(Z)-octadecenoic acid ¼ vernolic acid] was de-scribed (Hamberg and Fahlstadius, 1990). Vernolic acid is a

    naturally occurring fatty acid enriched especially in the seed oil

    (50 to 90% [w/w] of the total fatty acids) of several Asteraceae

    genera (Gunstone, 1954; Badami and Patil, 1980) and Euphor-

    biaceae species (Kleinman et al., 1965; Spitzer et al., 1996) and is

    a substrate analog of 12,13-EOT only missing the C11 double

    bond. We tested vernolic acid for its inhibiting effects on Arabi-

    dopsis AOC2. In our coupled enzymatic assay, total activity was

    downregulated by 50%, with some loss of stereoselectivity

    (Figure 6).

    To determine the location of the active site and to gain insight

    into possible substrate binding modes, we soaked SeMet-AOC2

    crystals with vernolic acid and solved the structure of the com-

    plex at 1.7-Å resolution (for details, see Methods and Table 1). As

    expected by the structural similarity to the lipocalins, we found

    vernolic acid bound in the hydrophobic barrel cavity (Figure 7A).

    In the observed binding mode, the charged carboxylic head

    group is located outside the cavity on the protein surface, while

    the more hydrophobic part of the molecule with the oxirane is

    buried deep in the barrel. The molecule is in the extended con-

    formation, which is not directly compatible with the cyclization

    reaction. The dominant molecular interactions with the protein

    are shown in Figure 7B. Overall the protein contributes mostly

    aromatic and hydrophobic residues to binding. The only polar/

    charged contacts are with Glu-23 close to the C15 double bond

    and with the tightly bound water molecule toward the oxirane.

    Figure 3. Sequence Alignment of Known AOCs.

    Representation of secondary structure elements and numbering is based on the SeMet-AOC2 structure. Asterisks mark residues modeled with

    alternate conformations. The following AOC2 sequences are listed (accession numbers in parentheses): Arabidopsis, AOC2_AT (Q9LS02), AOC1_AT

    (Q9LS03), AOC3_AT (Q9LS01), and AOC4_AT (Q93ZC5); Pisum sativum, AOC_PS (Q3LI84); Medicago truncatula, AOCa_MT (Q711Q9) and AOCb_MT

    (Q599T8); Oryza sativa, AOC_OS (Q8L6H4); Nicotiana tabacum, AOC_NT (Q711R1); Hordeum vulgare, AOC_HV (Q711R0); Zea mays, AOC_ZM

    (Q6RW09); Humulus lupulus, AOC1_HL (Q68IP7); AOC4_HL (Q68IP6); Bruguiera sexangula, MAN_BS (Q9AYT8); Lycopersicon esculentum, AOC_LE

    (Q9LEG5); S. tuberosum, AOC_ST (Q8H1X5); Physcomitrella patens, AOCa _PP (Q8GS38) and AOCb_PP (Q8H0N6). Figure was produced using the

    ESPript server (Gouet et al., 1999).

    3206 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • Based on the observed protein–inhibitor interactions, we de-

    cided to investigate the influence of single amino acid substitutions

    on the activity of AOC2. Phe-85 seems to contribute to binding of

    the substrate by forming a hydrophobic slide, which at the same

    time restricts the conformational freedom of the substrate during

    cyclization and thereby could render the reaction stereoselec-

    tive. With PCR-based mutagenesis, we exchanged Phe-85 with

    Leu (F85L) and Ala (F85A). Both mutants can be purified in similar

    amounts as the wild-type protein and show no overall folding

    defect as judged by circular dichroism (data not shown). The

    F85L mutant was successfully crystallized and showed no

    structural deviations from the wild-type protein. In the coupled

    AOS/AOC2 activity test, the F85L mutant showed a moderately

    reduced total activity and a slight loss of stereoselectivity (Figure

    6). For the F85A mutant, this result was even more pronounced:

    here, total activity was reduced ;50%, and the stereoselectivitywas reduced to a ratio of 1:4. As this mutant was not as highly

    purified as the other proteins, the absolute activity values have to

    be treated with some caution.

    DISCUSSION

    Comparison with Other Arabidopsis AOC Structures

    During the preparation of this manuscript, the Center for Eukary-

    otic Structural Genomics consortium (Madison, WI) deposited

    the coordinates of ligand-free AOC2 (CESG-AOC2, entry 1Z8K)

    and ligand-free AOC1 (CESG-AOC1, entry 1ZVC) to the Protein

    Data Bank (G.E. Wesenberg, G.N. Phillips Jr., E. Bitto, C.A.

    Bingman, and S.T.M. Allard, unpublished data). Although there is

    no publication yet on these structures, it is interesting to compare

    our results with the available coordinates.

    CESG-AOC2 has been solved using SeMet-labeled protein

    and was refined at 1.7-Å resolution in the same orthorhombic

    space group as our SeMet-AOC2 structure. Both models super-

    impose very well with an RMSD of only 0.31 Å for 173 Ca atoms.

    A stereoview of the superposition is shown in Figure 8. Closer

    inspection does not reveal any significant structural differences;

    specifically, the trimerization and the tightly bound water molecule

    inside the cavity are also found in the CESG-AOC2 structure.

    More interesting is the comparison with CESG-AOC1. The

    structure has been solved at a resolution of 1.8 Å in a hexagonal

    space group. In these crystals, the same mode of trimerization is

    observed for the protein, with the trimer axis now corresponding

    to a crystallographic symmetry. Again, we find the same tightly

    bound water molecule in the proposed active site of the enzyme.

    Sequence comparison to AOC2 shows a total of 10 mutations

    Figure 4. Trimeric Form of Arabidopsis AOC2.

    Ribbon representation of the AOC2 trimer. Residues 15 to 147 are shown

    in gray, and residues 148 to 188 are colored red. The view is along the

    trimer axis.

    Figure 5. Structural Superposition of AOC2 and RBP.

    Stereo diagram of pig RBP (Protein Data Bank accession code 1AQB) with the bound ligand retinol superposed with AOC2. RBP is drawn as a yellow

    ribbon, and the bound retinol is shown in van der Waals representation in purple and red. AOC2 is shown as a gray ribbon. To illustrate the discussed

    strand exchange, N-terminal portions of both proteins are colored blue (residues preceding the first b-strand in RBP and in the first b-strand in AOC2).

    C-terminal stretches of the proteins are colored maroon (including the last b-strand for RBP) and red (after the last residue in the barrel for AOC2). For

    clarity, the last 29 residues of RBP have been omitted, which include the conserved lipocalin helix mentioned in the text.

    Crystal Structure of AOC2 3207

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • (depicted in stick model in Figure 8), most of which can be

    classified as conservative exchanges, while in three cases a

    charge reversal occurs. Introduction of a positive charge at

    position 42 (Met to Arg) in the first turn is counterbalanced by the

    exchange of Lys-81 to Asn, which is located on the neighboring

    turn after the third b-strand. While we can observe a slight

    displacement of the main chain in that region, this is likely to be a

    result of different crystal packing, as Leu-41 and Met-42 are

    involved in crystal contacts in the case of AOC2. The preceding

    three residues (corresponding to residues 39 to 41 in AOC2) have

    not been modeled in CESG-AOC1, probably due to disorder.

    Another exchange involving a reversal in local charge is located

    on the opposite side of the protein at residue 178 due to the

    exchange of Glu to Lys.

    None of the changes seem to be of potential relevance for the

    ligand binding site. Indeed, we found no significant differences in

    either total activity, stereo, or substrate specificity for all four

    AOC isoforms (AOC1-4) in our activity assays (P. Zerbe, V.

    Effenberger, and F. Schaller, unpublished data). This is in con-

    trast with the observation of Stenzel et al. (2003b), who reported

    AOC2 having the highest activity but without quantifying this.

    A differential expression pattern of AOCs is clearly established.

    Upon wounding and application of jasmonic acid, expression of

    AOC2 is upregulated to a higher degree than that of AOC1

    (Stenzel et al., 2003b), whereas during senescence, a stronger

    upregulation of expression has been observed for AOC1 than for

    AOC2 (He et al., 2002). The other two isoforms, AOC3 and AOC4,

    seem not to be as strongly upregulated or sometimes even

    downregulated. The small structural differences, which mainly

    involve a change in the surface potential pattern between AOC1

    and AOC2, might therefore represent a subtle optimization for

    interactions of these proteins with different partners either in

    different cellular locations or physiological states of the plant. For

    a complete understanding of these regulatory processes, more

    data on the expression profiles for all four enzymes in planta are

    necessary. Meanwhile, the structure elucidation of AOC3, whose

    crystallization has already been indicated by the CESG in the

    structural genomics target database (http://targetdb.pdb.org/),

    might shed some more light on the structural variations between

    the isoforms.

    Structural Relation to the Lipocalin Family

    While a database search with the PROSITE lipocalin signature

    motif results in 113 hits from plants, there are to our knowledge

    only four examples of plant lipocalins explicitly assigned in the

    literature: the small temperature-induced lipocalin TIL (Charron

    et al., 2002), the two xanthophyll cycle enzymes violaxanthin

    deepoxidase (VDE) and zeaxanthin epoxidase (ZEP) (Burgos

    et al., 1998), and the recently identified chloroplast lipocalin CHL

    (Charron et al., 2005). The last three proteins are located in the

    chloroplast. While all four proteins share homology in the struc-

    turally conserved regions, especially in SCR I with the invariant

    Gly and Trp present, homology in SCR II and III is much weaker,

    suggesting that these plant proteins would fall in the class of

    outlier lipocalins (Burgos et al., 1998). The other hits produced by

    the automated search are either false positive hits or the clas-

    sification as lipocalins has not yet been supported by analysis of

    functional data.

    As AOC2 can be assigned structurally to the lipocalin fold, this

    would allow placement in one of several protein families showing

    this general folding pattern or even setting AOC2 aside as a new

    structural motif on its own. Considering the observed topology of

    the barrel and functional similarities, we propose to consider

    AOC2 as a distant member of the lipocalin family in plants. The

    most relevant differences are discussed below.

    The different placement of the termini in the superposition is

    striking (Figure 5). The barrel organization can be described right-

    or left-handed in the following way: if the thumb points upwards

    in the direction of the first b-strand, the bent fingers of the re-

    spective hand should indicate the sequential organization of the

    successive b-strands. By this definition, all lipocalin family mem-

    bers known so far (to our knowledge, this extends to all structures

    assigned to the lipocalin fold) form a left-handed barrel, while the

    AOC2 calyx is right-handed. In both cases, possible ligand

    binding sites are inside the palm and are accessible from above.

    As mentioned before, N- or C-terminal strand insertion with

    concomitant strand removal on the opposite end would auto-

    matically convert right- to left-handed barrels and vice versa.

    Along these lines, the first b-strand in AOC2 (blue in Figure 5)

    could have been formed by the insertion of the longer N-terminal

    part of the classical lipocalins (also colored blue) into the barrel.

    This would push b-strand H of the lipocalins (shown in maroon)

    out of the continuous b-sheet. The resulting differences in the

    structural organization would reduce the central role of the

    tryptophane in SCR I, and a loss of similarity in the SCRs would

    be obvious. Indeed, the observed weak similarity to SCR I in the

    second b-strand could be indicative of such an early process.

    Figure 6. Enzymatic Activity of AOC2 Mutants and AOC2 with the

    Inhibitor Vernolic Acid.

    For determination of enzymatic activity, 5 mg of AOS and 10 mg of AOC2

    were mixed in 1 mL of 10 mM PPi buffer, pH 7.0. After addition of 100 mg

    of HPOT as substrate, samples were incubated at room temperature

    for 15 min. Synthesized OPDA was extracted with ethyl acetate and

    analyzed via gas chromatography–mass spectrometry (GC-MS) as

    described in Methods. Shown are the results of the coupled enzymatic

    assay for AOS alone (1), AOS and AOC2 (2), AOS and AOC2 in presence

    of 160 mM inhibitor vernolic acid (3), and the mutants F85L (4) and F85A

    (5). The results are presented as the relative amount of synthesized

    OPDA compared with the yield with AOC2. The gray bars show total

    synthesized OPDA, whereas the black and white bars represent the yield

    of (þ)- and (�)-enantiomer, respectively. Error bars represent SD basedon three independent triplicate measurements (1, 2, 4, and 5) or a

    duplicate measurement (3).

    3208 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • Figure 7. Structure of AOC2 with Bound Inhibitor Vernolic Acid.

    (A) In the stereoview, monomer B is shown in gray ribbon representation, vernolic acid (VA) as blue stick model, and Water75 as red sphere. Residues

    directly involved in interactions with either the inhibitor or Water75 are depicted in yellow stick representation and labeled. Phe-85 is colored red. Omit

    difference density for vernolic acid (see Methods) at 2.4 s is drawn as red chicken wire. The view is rotated ;908 around the vertical with respect toFigure 2A.

    (B) Ligplot sketch showing the molecular interactions between the protein and the bound inhibitor vernolic acid. Contacts to the conserved Water75 are

    shown in green dashed lines with distances, and hydrophobic contacts are represented by opposing red spoked arcs. The interaction of Glu-23 is

    represented as a yellow arc. Vernolic acid is shown in purple.

    Crystal Structure of AOC2 3209

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • This divergence in structure from the classical fold is not

    unexpected. Investigation of the exon-intron structure of lip-

    ocalins clearly places the plant lipocalins together with Dictos-

    telium lipocalins into a gene structure–related group distinct from

    lipocalins from bacteria, insects, or animals (Sanchez et al.,

    2003). Since the early acquisition of the ancestor gene, it has

    been duplicated and adapted to different tasks. While TIL still can

    be counted as a genuine lipocalin, VDE evolved to be catalytically

    active in deepoxidation and acquired additions to the central

    barrel structure. With a size of ;350 amino acids, VDEs aremuch larger than the typical lipocalin monomer, which consists

    of 160 to 180 amino acids (Flower, 1994; Flower et al., 1995;

    Burgos et al., 1998). The same is true for ZEP, for which threading

    methods don’t even predict a barrel structure. Therefore, the

    peculiar architecture of these proteins raised doubt as to whether

    they truly belong to the lipocalin family (Ganfornina et al., 2000;

    Salier, 2000).

    Up to now, not enough representatives of plant lipocalin genes

    have been identified to decide whether the clade of plant

    lipocalins does indeed show a much higher variability as the

    better-characterized lipocalins from bacteria, insects, or ani-

    mals.

    On the other hand, one cannot exclude the possibility that we

    in fact do observe an example of convergent evolution. This is an

    ongoing debate given the very low sequence homology among

    lipocalins and especially for the outlier lipocalins, for which only

    sequence similarity in SCR I is detectable anyway (Grzyb et al.,

    2006).

    Oligomerization of AOC2

    In former investigations, AOC2 has been characterized to form

    dimers based on size exclusion chromatography (Ziegler et al.,

    1997), but several lines of evidence point toward a trimer as the

    physiological relevant oligomerization state of the protein. First,

    we can detect trimers of recombinant AOC2 or AOC2 from plant

    extracts in SDS page or immunoblots. Secondly, we observe

    very large buried surface areas upon trimerization that include

    three salt bridges involving the terminal carboxy group and

    several strong hydrogen bonds. This is in agreement with the loss

    of trimer formation for the construct with the C-terminal His-tag,

    which would likely interfere with the formation of the salt bridges.

    Taken together, this pattern is indicative of a stable protein–

    protein interaction. Importantly, residues involved in these interac-

    tions are predominantly conserved in all known AOC sequences.

    Additionally, while both wild-type AOC2 and the CESG-AOC1

    from Arabidopsis crystallize in different space groups and have

    therefore a modified crystal packing organization, in both cases

    the same mode of trimerization is observed.

    AOC from maize has been shown to have an apparent molec-

    ular mass of 47 kD in size exclusion chromatography and was

    discussed as a dimer (Ziegler et al., 1997). We found a similar

    behavior for AOC2 of Arabidopsis (data not shown), but based

    on our additional structural and biochemical data, we attribute

    this discrepancy of 25% to the calculated molecular weight of a

    trimer to an atypical retardation of this protein on the size

    exclusion matrix.

    We therefore expect the trimeric form to be the physiological

    relevant oligomerization state for AOCs of Arabidopsis and maize.

    Nevertheless, this SDS-stable trimer formation seems not to be

    absolutely required, since both the C- and N-terminally tagged

    proteins show similar enzymatic activities. However, trimer for-

    mation might improve overall protein stability.

    Active Site and Reaction Mechanism

    Based on the structural similarity of AOC2 to the lipocalins, we

    propose that the active site is located inside the barrel cavity.

    Figure 8. Structures of AOC2, CESG-AOC2, and CESG-AOC1

    Shown is a stereoview of the Ca trace of AOC2 (in gray), CESG-AOC2 (accession code 1Z8K; in red), and CESG-AOC1 (accession code 1ZVC; in blue).

    The 10 residues of AOC2 that differ in the AOC1 sequence are depicted in yellow stick representation.

    3210 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • While most lipocalins just bind small hydrophobic substrates in

    the interior, very few are known to be enzymatically active. One ex-

    ample is noteworthy: human b-trace or prostaglandinsynthase-D

    catalyzes the conversion of prostaglandin H2 to prostaglandin

    D2. As in the case of AOC2, the substrate is unstable in water and

    catalysis in a protected environment seems to be favorable.

    Here, the active site Cys is located in the barrel interior (Nagata

    et al., 1991).

    Indeed, in AOC2 we find the competitive inhibitor vernolic acid

    to be bound inside the barrel of one monomer. We don’t observe

    any induced fit mechanism, as the protein scaffold superposes

    with an RMSD of 0.33 Å for all 174 Ca atoms with SeMet-AOC2.

    Even on the level of side chains only marginal differences are

    present both in comparison to SeMet-AOC2 and to the other two

    monomers without bound inhibitor.

    Surprisingly, the carboxylic end of the inhibitor does not seem

    to be tightly bound to the protein, as there is no clear electron

    density for the first five carbon units. This is in contrast with the

    observation of Ziegler and coworkers that methylation of this

    group abolishes the inhibitory effect for AOC from maize (Ziegler

    et al., 1997). One possible explanation is the exchange of Leu-27

    of AOC2 to Arg in maize AOC, which would be positioned to form

    a strong salt bridge to the inhibitor in this protein. Another

    possibility would be that the carboxylic moiety is involved in the

    interaction with AOS for both proteins. Because we don’t have

    data on the influence of methylation on inhibition of AOC2, a

    decision between both cases is not possible yet. Work is in

    progress to obtain comparable data on both proteins with the

    different forms of the inhibitor.

    The finding that the location of the epoxy group in the 12,13-

    position of vernolic acid is essential for binding to AOC from

    maize (Ziegler et al., 1997) is consistent with the observed bind-

    ing position in our structure. Any shift of the epoxy group would

    inhibit the formation of the hydrogen bond to the conserved

    water molecule.

    To allow a discussion of possible reaction mechanisms, we

    tentatively modeled 12,13-EOT (substrate) and OPDA (product)

    molecules into the pocket in the same binding mode as vernolic

    acid (Figure 9A). Based on the resulting arrangement of protein

    side chains around the bound substrate, we postulate the

    following reaction mechanism for the cyclization reaction inside

    the protein (Figure 10).

    After binding into the pocket, the opening of the oxirane ring

    and the subsequent formation of the classical pentadienyl cation

    will be controlled by the protein environment. The influence of the

    negative charge of Glu-23 leads to a delocalization of the C15

    double bond toward the oxirane, which promotes its opening and

    the cyclization reaction by the mechanism of anchimeric assis-

    tance (Figure 10A). Formation of the oxyanion is stabilized by the

    tightly bound water molecule Water75. For the subsequent peri-

    cyclic ring closure, a conformational change around the C10-C11

    bond from trans to cis geometry is necessary, producing a non-

    planar ring-like pentadienyl carbocation (Figure 10B). Phe-51 is

    positioned to stabilize this by p–p interaction. Additionally, Cys-71

    is in a favorable position to stabilize the positive carbocation. This

    trans-cis isomerization around C10-C11 has to be accompanied

    by a cis-trans rotation around the C8-C9 bond due to the steric

    restrictions in the protein cavity. Additionally, the conformational

    change is promoted by the hydrophobic effect, as it buries more of

    the hydrocarbon tail of the molecule inside the cavity.

    The proposed reaction scheme assumes as a final step a

    classical conrotary pericyclic ring closure, according to the

    Woodward-Hoffmann rules (Figure 10C). The absolute stereo-

    selectivity of the reaction follows from the steric restrictions by

    the protein environment on the formation of the cis intermediate.

    Phe-85, Val-45, and Tyr-105 especially seem to form a greasy

    slide, which at the same time facilitates and restricts the confor-

    mational change of the hydrocarbon tail. Indeed, the F85L and

    F85A mutants show a reduced reactivity and stereoselectivity of

    the reaction. The latter effect would be explained by the removal

    Figure 9. Substrate and Product Modeled into the Binding Pocket of AOC2 and Comparison with PGHS-1.

    (A) Shown are 12,13-EOT and OPDA as green and blue sticks, respectively. The protein is shown in a similar representation as in Figure 7. Also drawn is

    the molecular protein surface colored according to atom types. Possible hydrogen bonds between substrate and protein are shown as dashed green

    lines.

    (B) View into the cyclooxygenase active site of PGHS-1 from sheep (Ovis aris) (accession number 1DIY) with bound substrate arachidonic acid (AA). A

    similar representation is chosen as compared with (A). Also shown are the residues discussed in the text.

    Crystal Structure of AOC2 3211

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • of some of the steric restrictions on the trans-cis rotation around

    C10-C11. A rotation in the opposite direction would automati-

    cally result in the opposite stereoisomer of the product.

    Our proposed mechanism explains the importance of the

    C15 double bond for reactivity along the lines of the proposal of

    Grechkin (1994) by the mechanism of anchimeric assistance. It

    also rationalizes the impact of the protein environment on the

    stereoselectivity of the cyclization reaction. It is clear by the

    packing in the cavity that changes in the position of the epoxide

    group will affect binding affinities of possible substrates.

    A similar biosynthesis of cyclic hormones can be found in

    mammals during formation of prostanoids, which are structurally

    similar to JAs. One of the enzymes involved is the Lipocalin-type

    prostaglandin D synthase, which has been shown to be an ex-

    ample for an enzymatically active lipocalin (Urade and Hayaishi,

    2000). The enzyme responsible for the first committed step in

    prostanoid catalysis from the linear precursor is prostaglandin

    endoperoxide synthase (PGHS). Due to its medical relevance as

    target for anti-inflammatory drugs like aspirin, it has been very

    well characterized. PGHS catalyzes both the cyclization of

    arachidonic acid to prostaglandin G2 and the successive per-

    oxidation of PGG2 to prostaglandin H2. While PGHS shows a

    completely different overall architecture than AOC, the functional

    homology makes a comparison of the active sites worthwhile.

    PGHS is a dimeric monotopic membrane protein with a

    molecular mass of ;70 kD (monomer), which is located in theendoplasmic reticulum and nuclear envelopes. Besides an

    N-terminal EGF domain, the protein is mainly a-helical and has

    two distinct active sites for the two catalyzed reaction steps

    (Picot et al., 1994; for a review, see Garavito et al., 2002). In

    the cyclooxygenase active site, arachidonic acid binds with the

    carboxylic end, forming salt bridges to a conserved Arg residue.

    The rest of the molecule is tightly packed into an L-shaped cavity

    lined mainly with hydrophobic and aromatic residues, as found in

    the AOC2 binding pocket (Figure 9B). The catalytically active Tyr

    is positioned at the kink of the cavity for abstraction of the 13proS

    hydrogen, the initiating step of the reaction. Interestingly, the

    only other polar residue is situated on the opposite side of the

    polycarbon chain, similar to the position of Cys-71 in the case of

    AOC2. This has been shown to be critical in determining the

    stereochemistry of the product (Schneider et al., 2002). It has

    been found that the packing at the v-end of the substrate

    molecule is very important for maintaining the catalytic selectivity

    and activity (Rowlinson et al., 1999; Malkowski et al., 2001). While

    increasing the size of residues in this region can abolish activity

    due to steric inhibition of binding, mutations increasing the vol-

    ume of the binding pocket mostly diminish activity, as optimal

    positioning of substrate is not enforced. It has been proposed

    that the carboxylic end stays fixed during catalysis in PGHS,

    while the v-end has to move closer to allow for the substantial

    conformational changes necessary for ring closure (Garavito

    et al., 2002). In the case of AOC2 from Arabidopsis, a movement

    of the carboxyl end of the substrate along our proposed reaction

    scheme seems to be more likely. While the relative flexibility of

    the carboxy group of vernolic acid might not represent the

    situation for the substrate 12,13-EOT in planta, there are no

    positively charged residues appropriately placed to form such a

    strong salt linkage as observed in the case of PGHS.

    Here, we find two completely different structures showing the

    same binding principles. A hydrophobic mold binds the sub-

    strate exactly in the right position to a single catalytically critical

    polar site without apparent induced fit during binding. While the

    exact reaction schemes are very different, the hydrophobic mold

    in both cases also ensures the stereospecificity of the reaction.

    As the comparison with PGHS shows, residues in the vicinity of

    the substrate binding site are probably highly optimized to con-

    trol the reaction in the cavity. Structural and functional analyses

    of additional point mutants will help to clarify the individual

    influence of the key amino acids in the binding pocket on the

    proposed mechanism and the stereoselectivity of the reaction.

    Figure 10. Schematic Overview of the Proposed Cyclization Reaction.

    The AOC2 catalyzed cyclization reaction involves the opening of the

    oxirane ring and a subsequent conrotary pericyclic ring closure (see text

    for details). The fatty acid moiety of 12,13-EOT [(CH2)7COOH] is labeled

    as R, and the schematic binding pocket of AOC2 is displayed in gray.

    3212 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • METHODS

    Protein Expression and Purification

    A truncated version of AOC2 from Arabidopsis thaliana has been cloned

    into the vector pQE30 (Qiagen) and pET21b (Novagen/Merck) to yield

    a fusion protein (20.2 kD) in which the first 77 amino acids (the pre-

    dicted transit signal) have been replaced by a His6-tag (sequence

    MRGSHHHHHHRS) or in which the His6-tag has been coupled with the

    C terminus of the truncated protein. The resulting constructs were

    transformed into Escherichia coli strain M15. Cells were grown in 2YT

    media and induced at an OD of 0.5 to 0.6 with addition of 0.2 mM

    isopropylthio-b-galactoside and 0.2% arabinose, respectively. E. coli

    with the pQE30 construct were harvested after an induction time of ;5 hat 378C and 220 rpm. Bacteria with the pET21b construct were induced

    for a further 30 to 40 h at 258C. Cells were broken by ultrasonication, and

    the fusion protein was purified by affinity chromatography (Ni-NTA;

    Qiagen) and concentrated with Centricon concentrators (5000 D cutoff;

    Millipore). The yield of purified protein exceeded 10 mg/L culture.

    SeMet-labeled N-terminally His6-tagged AOC2 was produced by

    growth of M15 cells in M9 minimal medium supplemented with SeMet

    as described (Van Duyne et al., 1993). The mutated proteins of AOC2

    were expressed in the pQE vector under the same conditions as the

    native AOC2.

    AOS (pQE30-AOS; Laudert et al., 1996) was expressed and purified in

    accordance with Oh and Murofushi (2002). Protein expression was

    induced at an OD600 of ;0.5 by addition of isopropylthio-b-galactosideto a final concentration of 0.4 mM and further incubation at 168C and 150

    rpm overnight. Cells were broken by ultrasonication, and recombinant

    AOS was purified via Ni-NTA agarose chromatography (Qiagen) by

    adding Triton X-100 up to a final concentration of 0.1% (v/v) to the

    recommended buffers.

    Generation of Mutant Proteins of AOC2

    The generation of AOC2 mutants was done via PCR according to the

    sequential PCR step method of the current protocols in molecular

    biology. Phe at position 85 was exchanged to Ala (primers: 59-GAAAGGT-

    GAAAGAGCCGAAGCTACTTATA-39and 59-TATAAGTAGCTTCGGCTC-

    TTTCACCTTTC-39) and Leu (primers: 59-GAAAGGTGAAAGACTCGAA-

    GCTACTTATA-39and 59-TATAAGTAGCTTCGAGTCTTTCACCTTTC-39).

    The N- and C-terminal primers were 59-TATGGATCCCCAAGCAAAGTT-

    CAAGAACTG-39 (BamHI restriction site underlined) and 59-TATGTC-

    GACTTAGTTGGTATAGTTACTTATAAC-39 (SalI restriction site under-

    lined). The mutated cDNA was cloned afterwards into the expression

    vector pQE30, and the protein was expressed under standard conditions.

    Gel Electrophoresis and Protein Immunoblotting

    Denaturing gel electrophoresis was performed according to Laemmli

    (1970). The discontinuous systems consisted of 4% stacking gels and

    12.5% resolving gels. Protein blotting onto nitrocellulose was done

    electrophoretically overnight (48C, 64 mA) as described by Towbin et al.

    (1979). Immunodetection followed standard procedures (Parets-Soler

    et al., 1990) with goat anti-rabbit immunoglobin-conjugated alkaline

    phosphatase and the enzyme substrates 4-nitrotetrazolium blue and

    5-bromo-4-chloro-3-indolyl phosphate.

    Determination of Total Enzymatic Activity via

    Chiral Capillary GC-MS

    For activity analysis, 5 to 20 mg of purified AOS and AOCs (weight-based

    ratio 1:2) were added to 1 mL of 10 mM PPi buffer, pH 7.0. The reaction

    was started by addition of 100 mg of HPOT. After incubation for 15 min at

    room temperature, the reaction was stopped by acidifying with 1 M HCl to

    pH 3.0. The synthesized OPDA was extracted twice with 2 volumes of

    ethyl acetate and taken to dryness. For GC-MS analysis, 1 nmol of 2[H]5-

    OPDA was added before extraction as an internal standard. Dried

    samples were incubated in 250 mL of 0.1 M KOH for 45 min at room

    temperature to convert the synthesized OPDA into trans-configuration.

    After a second extraction step, OPDA was dissolved in methanol and

    methylated with ethereal diazomethane (Hamberg and Fahlstadius,

    1990). The dried fractions were then redissolved in 50 to 100 mL of

    chloroform, and 1 mL of the sample was injected into a Varian GC 3400

    gas chromatograph in splitless mode in direct connection to a MAT

    Magnum ion trap mass spectrometer (Finnigan) using a chemical ioniza-

    tion mode with methanol as reactant gas. Separations of the (þ)- and the(�)-enantiomer of OPDA were made on a b-Dex120 column (30 m 3 0.25mm 3 0.15 mm stationary-phase thickness) coated with 20% permethyl-

    b-cyclodextrin in SPB-35 (Supelco) with He carrier gas at 1 mL/min (gas

    prepressure of 80 kPa) using the following temperature program: injector

    temperature of 2208C, 1 min isothermally at 508C, with 108C/min up to

    1908C, 75 min isothermally at 1908C, with 38C/min up to 2208C, 15 min

    isothermally at 2208C, and transfer line temperature 2208C. For details,

    see Laudert et al. (1997) and Schaller et al. (1998). Quantitation of AOC

    enzymatic activity was based on (þ)-OPDA-to-(�)-OPDA ratios derivedfrom integration of total ion current traces.

    Crystallization

    Initial crystallization screening for AOC2 was performed using vapor

    diffusion against the Wizard I and II kits from Emerald BioSystems. Protein

    crystals were found with condition WI-28, which was subsequently

    optimized. Using 14 to 16% PEG-8000, 6% tert-butanol, and 0.1 M

    MES, pH 5.0 to 6.0, as reservoir solution in either hanging- or sitting-drop

    setups (4 mg/mL protein concentration), we were able to grow crystals

    (space group P21) up to a size of 250 3 150 3 50 mm, which diffract to

    ;1.7-Å resolution.AOC2-SeMet crystals (space group P212121) were grown from reser-

    voir solutions containing 10% PEG-3350, 200 mM NaCl, and 100 mM

    phosphate-citrate, pH 4.2, with a protein concentration of 13.5 mg/mL.

    They typically reach a size of 250 3 200 3 100 mm and diffract to better

    than 1.3-Å resolution. Crystals grew at 188C in ;1 week.For soaking experiments, AOC2-SeMet crystals were transferred for

    1 to 2 d into appropriate mother liquor saturated with inhibitor.

    X-Ray Diffraction Data Collection

    Data collection was performed at 100K on flash-frozen crystals at the

    Swiss Light Source (Villigen, Switzerland) at beamline PXII using a MAR

    CCD detector. Cryoprotection was achieved by briefly soaking the

    crystals in paraffin oil or in mother liquor supplemented with glycerol.

    For SeMet-AOC2 crystals, MAD data at the Se edge were collected at the

    European Synchroton Radiation Facility (Grenoble, France) on beamline

    ID13.1 at 100K using a MAR CCD detector. Data of inhibitor-soaked

    SeMet-AOC2 crystals were collected in house at 100K on a Bruker FR591

    rotating anode with Montel multilayer optics and a MAR345dtb detector

    system.

    Data were processed using the XDS package (Kabsch, 1993). Statistics

    of the different data sets are shown in Table 1.

    Structure Solution and Refinement

    SeMet-AOC2

    Self-rotation functions indicated the presence of a threefold noncrystallo-

    graphic symmetry. Search for the Se substructure and initial phasing

    were done with the programs of the SHELX package (Sheldrick and

    Crystal Structure of AOC2 3213

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • Schneider, 1997) using the hkl2map interface (Pape and Schneider, 2004)

    with data to 2-Å resolution. SHELXD identified four major peaks, two of

    which in retrospect correspond to two copies of SeMet42. The other two

    are a double peak corresponding to the third copy of SeMet42. This

    solution showed correlation coefficients of 41 for all data and 32 for weak

    data. Phasing with SHELXE using data to 3 Å clearly identified the correct

    hand and already resulted in a well-interpretable map. Fifty-eight cycles

    of density modification using an estimated solvent content of 0.4 resulted

    in density contrast of 0.73 (0.38), connectivity of 0.95 (0.91), and a

    pseudo-free correlation coefficient of 83.2 (63.3) (values for the wrong

    hand in parentheses).

    Density modification (using the threefold symmetry), phase extension,

    and auto building in RESOLVE (Terwilliger, 2000) could assign 387 of the

    expected 564 (3 3 188) residues, including side chains. An additional 44

    residues were placed without side chains.

    For refinement, the peak wavelength data set was used. Five percent of

    the data were randomly assigned to the Rfree test set and omitted from

    refinement.

    The resulting model was further improved in several cycles of refine-

    ment, automated water placement, and manual rebuilding using pro-

    grams of the CCP4 package (Collaborative Computational Project,

    Number 4, 1994) and ARP/wARP (Cohen et al., 2004).

    The final model contains three protein monomers composed of resi-

    dues 15 to 188. The first 14 residues, including the His6-tag, could not be

    placed due to missing density. Based on clear density, a total of 19

    residues had to be modeled with alternate conformations. Additionally,

    three glycerol and 556 water molecules were included. Part of the

    experimental electron density map after RESOLVE with the final model

    is shown in Supplemental Figure 2 online.

    WT-AOC2

    The structure of WT-AOC2 was solved using the refined coordinates of

    SeMet-AOC2 after removal of solvent and alternate conformations and

    the mutation of SeMet to Met as a search model. Two trimers were found

    in the asymmetric unit of the monoclinic unit cell. After several rounds of

    refinement in REFMAC (Murshudov et al., 1997) and manual rebuilding

    and water picking in COOT (Emsley and Cowtan, 2004), the final model

    consists of 1036 amino acids, five glycerol molecules, two sodium ions,

    and 660 water molecules. Gln-102 had to be modeled in alternate

    conformations in five of the six copies of AOC2 in the asymmetric unit.

    Weak NCS restraints were used throughout the refinement.

    SeMet-AOC2 Inhibitor

    For analysis of the inhibitor data, one copy of the 2BRJ model without

    alternate conformations and solvent molecules was placed in the unit cell.

    After several rounds of refinement and water picking, extended residual

    density in monomer B was fitted with a model of the inhibitor vernolic acid.

    The initial atomic coordinates and the necessary geometry definitions

    were created using the PRODRG server (Schuettelkopf and van Aalten,

    2004). At the end of the refinement, the occupancy of the inhibitor was set

    to 0.25, 0.5, 0.75, and 1 and the structure refined using identical pro-

    tocols. With full occupancy, the best R-factors were reached. While we

    observed clear density for the majority of the inhibitor, the carboxylic end

    up to C5 did not show clear density due to disorder (Figure 7A). The shape

    of the resulting difference density together with the higher flexibility of the

    inhibitor in comparison to the surrounding protein atoms also points to

    different minor binding modes of the inhibitor (possibly also the second

    enantiomer), which were not modeled.

    Elongated residual density in monomer C was judged too weak to

    safely allow positioning of the inhibitor. In monomer A, the entrance to the

    binding site is completely blocked by Leu-41 from a crystallographically

    related molecule. Therefore, binding of inhibitor during soaking in this

    monomer is prohibited. Weak NCS restraints were used in the initial

    stages of the refinement but were released after inclusion of the inhibitor.

    For preparation of the omit maps, vernolic acid was removed from the

    final model and random positional shifts (average RMSD 0.12 Å) were

    applied to the remaining coordinates to remove model bias. After another

    round of refinement with REFMAC, an mFo-DFc difference omit map was

    calculated.

    Figure Preparation

    Figures were generated using Pymol (http://www.pymol.org). To allow

    coloring of residues by conservation, an alignment of 22 different AOCs

    found with a National Center for Biotechnology Information BLAST search

    in the UNIREF100 database was analyzed using the ESPript server (Gouet

    et al., 1999). In the resulting coordinate file, the similarity score has been

    assigned to the B-factor column, ranging from 100 (fully conserved) to 0

    (no conservation). This range was assigned to the color range red to blue

    for figure generation.

    Superposition of molecules was either performed in COOT (Emsley and

    Cowtan, 2004) or using the SuperPose server (Maiti et al., 2004).

    Accession Numbers

    Coordinates and structure factors can be found in the Protein Data Bank

    under accession numbers 2BRJ, 2GIN, and 2DIO for the AOC2-SeMet,

    WT-AOC2, and AOC2-SeMet inhibitor structure, respectively.

    Supplemental Data

    The following materials are available in the online version of this article.

    Supplemental Figure 1. Conservation of the Trimer Interface.

    Supplemental Figure 2. MAD-Solvent-Flattened Electron Density of

    AOC2.

    ACKNOWLEDGMENTS

    We thank C. Kötting for helpful discussions, K. Diederichs and D.R.

    Madden for comments on the manuscript, and E.W. Weiler for continu-

    ing support and interest in the work. Part of this work was performed at

    the Swiss Light Source, Beamline X10SA, Paul Scherrer Institute

    (Villigen, Switzerland) and at the European Synchroton Radiation Facil-

    ity, Beamline ID13.1 (Grenoble, France). We would especially like to

    acknowledge the help of the respective beamline staff and of our

    colleagues from the Max Planck Institutes for Medical Research

    (Heidelberg) and Molecular Physiology (Dortmund) during data collec-

    tion. We also acknowledge financial support from the Deutsche For-

    schungsgemeinschaft by Grants SFB480-C6 (E.H.), HO2600 (E.H.), and

    SCHA939 (F.S.).

    Received May 10, 2006; revised August 4, 2006; accepted October 17,

    2006; published November 3, 2006.

    REFERENCES

    Badami, R.C., and Patil, K.B. (1980). Structure and occurrence of

    unusual fatty acids in minor seed oils. Prog. Lipid Res. 19, 119–153.

    Baertschi, S.W., Ingram, C.D., Harris, T.M., and Brash, A.R. (1988).

    Absolute configuration of cis-12-oxophytodienoic acid of flaxseed:

    3214 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • Implications for the mechanism of biosynthesis from the 13(S)-

    hydroperoxide of linolenic acid. Biochemistry 27, 18–24.

    Blée, E. (2002). Impact of phyto-oxylipins in plant defense. Trends Plant

    Sci. 7, 315–322.

    Blée, E., and Joyard, J. (1996). Envelope membranes from spinach

    chloroplasts are a site of metabolism of fatty acid hydroperoxides.

    Plant Physiol. 102, 821–828.

    Brash, A.R., Baertschi, S.W., Ingram, C.D., and Harris, T.M. (1988).

    Isolation and characterization of natural allene oxides: Unstable

    intermediates in the metabolism of lipid hydroperoxides. Proc. Natl.

    Acad. Sci. USA 85, 3382–3386.

    Burgos, R.C., Hieber, A.D., and Yamamoto, Y. (1998). Xanthophyll

    cycle enzymes are members of the lipocalin family, the first identified

    from plants. J. Biol. Chem. 273, 15321–15324.

    Charron, J.-B.F., Breton, G., Badawi, M., and Sarhan, F. (2002). Mo-

    lecular and structural analyses of a novel temperature stress-induced

    lipocalin from wheat and Arabidopsis. FEBS Lett. 517, 129–132.

    Charron, J.-B.F., Ouellet, F., Pelletier, M., Danyluk, J., Chauve, C.,

    and Sarhan, F. (2005). Identification, expression, and evolutionary

    analyses of plant lipocalins. Plant Physiol. 139, 2017–2028.

    Cohen, S.X., Morris, R.J., Fernandez, F.J., Ben Jelloul, M., Kakaris,

    M., Parthasarathy, V., Lamzin, V.S., Kleywegt, G.J., and Perrakis,

    A. (2004). Towards complete validated models in the next generation

    of ARP/wARP. Acta Crystallogr. D60, 2222–2229.

    Collaborative Computational Project, Number 4 (1994). The CCP4

    suite: Programs for protein crystallography. Acta Crystallogr. D Biol.

    Crystallogr. D50, 760–763.

    Conconi, A., Miquel, M., Browse, J.A., and Ryan, C.A. (1996). Intra-

    cellular levels of free linolenic and linoleic acids increase in tomato

    leaves in response to wounding. Plant Physiol. 111, 797–803.

    Crombie, L., and Morgan, D.O. (1988). Formation of acylic a- and

    g-ketols and 12-oxophytodienoic acid from linolenic acid 13-hydro-

    peroxide by a flax enzyme preparation. Evidence for a single enzyme

    leading to a common allene epoxide intermediate. J. Chem. Soc.

    Perkin Trans. 8, 558–560.

    Dathe, W., Rönsch, H., Preiss, A., Schade, W., Sembdner, G., and

    Schreiber, K. (1981). Endogenous plant hormones of the broad bean,

    Vicia faba L. (-)-Jasmonic acid, a plant growth inhibitor in pericarp.

    Planta 153, 530–535.

    Demole, E., Lederer, E., and Mercier, D. (1962). Isolement et déter-

    mination de la structure du jasmonate de méthyle, constituant odor-

    ant charactéristique de l’essence de jasmin. Helv. Chim. Acta 45,

    675–685.

    Diederichs, K., and Karplus, P.A. (1997). Improved R-factors for

    diffraction data analysis in macromolecular crystallography. Nat.

    Struct. Biol. 4, 269–275.

    Emanuelsson, O., Nielsen, H., and von Heijne, G. (1999). ChloroP, a

    neutral network-based method for predicting chloroplast transit pep-

    tides and their cleavage sites. Protein Sci. 8, 978–984.

    Emsley, P., and Cowtan, K. (2004). Coot: Model-building tools for

    molecular graphics. Acta Crystallogr. D60, 2126–2132.

    Farmer, E.E., Johnson, R.R., and Ryan, C.A. (1991). Regulation of

    expression of proteinase inhibitor genes by methyl jasmonate and

    jasmonic acid. Plant Physiol. 98, 995–1002.

    Farmer, E.E., and Ryan, C.A. (1990). Interplant communication: Air-

    borne methyl jasmonate induces synthesis of proteinase inhibitors in

    plant leaves. Proc. Natl. Acad. Sci. USA 87, 7713–7716.

    Farmer, E.E., and Ryan, C.A. (1992). Octadecanoid-derived signals in

    plants. Trends Cell Biol. 2, 236–241.

    Feussner, I., and Wasternack, C. (2002). The lipoxygenase pathway.

    Annu. Rev. Plant Biol. 53, 275–297.

    Flower, D.R. (1994). The lipocalin protein family: A role in cell regulation.

    FEBS Lett. 354, 7–11.

    Flower, D.R., North, A.C., and Sansom, C.E. (2000). The lipocalin

    protein family: Structural and sequence overview. Biochim. Biophys.

    Acta 1482, 9–24.

    Flower, D.R., Sansom, C.E., Beck, M.E., and Altwood, T.K. (1995).

    The first prokaryotic lipocalins. Trends Biochem. Sci. 20, 498–499.

    Ganfornina, M.D., Guitierrez, G., Bastiani, M., and Sanchez, D.

    (2000). A phylogenetic analysis of the lipocalin protein family. Mol.

    Biol. Evol. 17, 114–126.

    Garavito, R.M., Malkowski, M.G., and de Witt, D.L. (2002). The

    structures of prostaglandin endoperoxide H synthases-1 and -2.

    Prostaglandins Other Lipid Mediat. 68–69, 129–152.

    Gerhardt, B. (1983). Localization of b-oxidation enzymes in peroxi-

    somes isolated from nonfatty plant tissues. Planta 159, 238–246.

    Gouet, P., Courcelle, E., Stuart, D.I., and Metoz, F. (1999). ESPript: Mul-

    tiple sequence alignments in PostScript. Bioinformatics 15, 305–308.

    Grechkin, A.N. (1994). Cyclization of natural allene oxide fatty acids.

    The anchimeric assistance of beta, gamma-double bond beside the

    oxirane and the reaction mechanism. Biochim. Biophys. Acta 14,

    199–206.

    Grzyb, J., Latowski, D., and Strzalka, K. (2006). Lipocalins – A family

    portrait. J. Plant Physiol. 163, 895–915.

    Gundlach, H., Müller, M.J., Kutchan, T.M., and Zenk, M.H. (1992).

    Jasmonic acid as a signal transducer in elicitor-induced plant cell

    cultures. Proc. Natl. Acad. Sci. USA 89, 2389–2393.

    Gunstone, F.D. (1954). Fatty acids: Part II: The nature of the oxygenated

    acid present in Vernonia anthelmintica (Willd.) seed oil. J. Chem. Soc.

    (Lond.) May, 1611–1616.

    Hamberg, M., and Fahlstadius, P. (1990). Allene oxide cyclase: A new en-

    zyme in plant lipid metabolism. Arch. Biochem. Biophys. 276, 518–526.

    Hamberg, M., and Hughes, M.A. (1988). Fatty acid allene oxides. III.

    Albumin-induced cyclization of 12,13(S)-epoxy-9(Z),11-octadecadie-

    noic acid. Lipids 23, 469–475.

    He, Y., Fukushige, H., Hildebrand, D.F., and Gan, S. (2002). Evidence

    supporting a role of jasmonic acid in Arabidopsis leaf senescence.

    Plant Physiol. 128, 876–884.

    Holm, L., and Sander, C. (1994). Searching protein structure databases

    has come of age. Proteins 19, 165–173.

    Kabsch, W. (1993). Automatic processing of rotation diffraction data

    from crystals of initially unknown symmetry and cell constants. J.

    Appl. Crystallogr. 26, 795–800.

    Kleinman, R., Smith, C.R., Jr., and Yates, S.G. (1965). Search for new

    industrial oils: Fifty-eight Euphorbiaceae oils, including one rich in

    vernolic acid. J. Am. Oil Chem. Soc. 42, 169–172.

    Kramell, R., Miersch, O., Hause, B., Ortel, B., Parthier, B., and

    Wasternack, B. (1997). Amino acid conjugates of jasmonic acid

    induce jasmonate-responsive gene expression in barley (Hordeum

    vulgare L.) leaves. FEBS Lett. 414, 197–202.

    Laemmli, U.K. (1970). Cleavage of structural proteins during the as-

    sembly of the head of bacteriophage T4. Nature 227, 680–685.

    Laudert, D., Hennig, P., Stelmach, B.A., Müller, A., Andert, L., and

    Weiler, E.W. (1997). Analysis of 12-oxo-phytodienoic acid enantio-

    mers in biological samples by capillary gas chromatography-mass

    spectrometry using cyclodextrin stationary phases. Anal. Biochem.

    246, 211–217.

    Laudert, D., Pfannschmidt, U., Lottspeich, F., Holländer-Czytko, H.,

    and Weiler, E.W. (1996). Cloning, molecular and functional charac-

    terization of Arabidopsis thaliana allene oxide synthase (CYP74), the

    first enzyme of the octadecanoid pathway to jasmonates. Plant Mol.

    Biol. 31, 323–335.

    Li, C., Schilmiller, A.L., Liu, G., Lee, G.I., Jayanty, S., Sageman, C.,

    Vrebalov, J., Giovannoni, J.J., Yagi, K., Kobayashi, Y., and Howe,

    G.A. (2005). Role of beta-oxidation in jasmonate biosynthesis and

    systemic wound signaling in tomato. Plant Cell 17, 971–986.

    Crystal Structure of AOC2 3215

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • Maiti, M., van Domselaar, G.H., Zhang, H., and Wishart, D.S. (2004).

    SuperPose: A simple server for sophisticated structural superposition.

    Nucleic Acids Res. 32, W590–W594.

    Malkowski, M.G., Thuresson, E.D., Lakkides, K.M., Rieke, C.J., Micielli,

    R., Smith, W.L., and Garavito, R.M. (2001). Structure of eicosapen-

    taenoic and linoleic acids in the cyclooxygenase site of prostaglandin

    endoperoxide H synthase-1. J. Biol. Chem. 276, 37547–37555.

    Maucher, H., Stenzel, I., Miersch, O., Stein, N., Prasad, M., Zierold,

    U., Schweizer, P., Dorer, C., Hause, B., and Wasternack, C. (2004).

    The allene oxide cyclase of barley (Hordeum vulgare L.)–Cloning and

    organ-specific expression. Phytochemistry 65, 801–811.

    Meyer, A., Miersch, O., Büttner, C., Dathe, W., and Sembdner, G.

    (1984). Occurrence of the plant growth regulator jasmonic acid in

    plants. J. Plant Growth Regul. 3, 1–8.

    Mueller, M.J., Brodschelm, W., Spannagl, E., and Zenk, M.H. (1993).

    Signaling in the elicitation process is mediated through the octade-

    canoid pathway leading to jasmonic acid. Proc. Natl. Acad. Sci. USA

    90, 7490–7494.

    Murshudov, G.N., Vagin, A.A., and Dodson, E.J. (1997). Refinement of

    macromolecular structures by the Maximum-Likelihood Method. Acta

    Crystallogr. D53, 240–255.

    Murzin, A.G. (1994). Principles determining the structure of b-sheet

    barrels in proteins. J. Mol. Biol. 236, 1369–1381.

    Murzin, A.G., Brenner, S.E., Hubbard, T., and Chothia, C. (1995).

    SCOP: A structural classification of proteins database for the inves-

    tigation of sequences and structures. J. Mol. Biol. 247, 536–540.

    Nagata, A., Suzuki, Y., Igarashi, M., Eguchi, N., Toh, H., Urade, Y.,

    and Hayaishi, O. (1991). Human brain prostaglandin D synthase has

    been evolutionarily differentiated from lipophilic-ligand carrier pro-

    teins. Proc. Natl. Acad. Sci. USA 88, 4020–4024.

    Narváez-Vásquez, J., Florin-Christensen, J., and Ryan, C.A. (1999).

    Positional specificity of a phospholipase A2 activity induced by

    wounding, systemin, and oligosaccharide elicitors in tomato leaves.

    Plant Cell 11, 2249–2260.

    Oh, K., and Murofushi, N. (2002). Design and synthesis of novel

    imidazole derivatives as potent inhibitors of allene oxide synthase

    (CYP74). Bioorg. Med. Chem. 10, 3707–3711.

    Oldham, M.L., Brash, A.R., and Newcomer, M.E. (2005). The structure

    of coral allene oxide synthase reveals a catalase adapted for metab-

    olism of a fatty acid hydroperoxide. Proc. Natl. Acad. Sci. USA 102,

    297–302.

    Pan, Z., Durst, F., Werck-Reichhart, D., Gardner, H.W., Camara, B.,

    Cornish, K., and Backhaus, R.A. (1995). The major protein of

    guayule rubber particles is a cytochrome P450. Characterization

    based on cDNA cloning and spectroscopic analysis of the solubilized

    enzyme and its reaction products. J. Biol. Chem. 270, 8487–8494.

    Pape, T., and Schneider, T.R. (2004). HKL2MAP: A graphical user

    interface for phasing with SHELX programs. J. Appl. Crystallogr. 37,

    843–844.

    Parets-Soler, A., Pardo, J.M., and Serrano, R. (1990). Immunocyto-

    localization of plasmamembrane Hþ-ATPase. Plant Physiol. 93,1654–1658.

    Pervaiz, S., and Brew, K. (1985). Homology of b-lactoglobulin, serum

    retinol-binding protein, and protein HC. Science 228, 335–337.

    Pervaiz, S., and Brew, K. (1987). Homology and structure-function

    correlation between alpha 1-acid glycoprotein and serum retinol-

    binding protein and its relatives. FASEB J. 1, 209–214.

    Picot, D., Loll, P.J., and Garavito, R.M. (1994). The X-ray crystal

    structure of the membrane protein prostaglandin H2 synthase-1.

    Nature 367, 243–249.

    Reymond, P., and Farmer, E.E. (1998). Jasmonate and salicylate as

    global signals for defense gene expression. Curr. Opin. Plant Biol. 1,

    404–411.

    Rowlinson, S.W., Crews, B.C., Lanzo, C.A., and Marnett, L.J. (1999).

    The binding of arachidonic acid in the cyclooxygenase active site of

    mouse prostaglandin endoperoxides synthase-2 (COX-2). J. Biol.

    Chem. 274, 23305–23310.

    Salier, J.P. (2000). Chromosomal location, exon/intron organiza-

    tion and evolution of lipocalin genes. Biochim. Biophys. Acta 18,

    25–34.

    Sanchez, D., Ganfornia, M.D., Gutierrez, G., and Marin, A. (2003).

    Exon-intron structure and evolution of the lipocalin gene family. Mol.

    Biol. Evol. 20, 775–783.

    Schaller, F., Biesgen, C., Müssig, C., Altmann, T., and Weiler, E.W.

    (2000). 12-Oxophytodienoate reductase 3 (OPR3) is the isoenzyme

    involved in jasmonate biosynthesis. Planta 210, 979–984.

    Schaller, F., Hennig, P., and Weiler, E.W. (1998). 12-Oxophytodie-

    noate-10,11-reductase: Occurrence of two isoenzymes of different

    specificity against stereoisomers of 12-oxophytodienoic acid. Plant

    Physiol. 118, 1345–1351.

    Schaller, F., Schaller, A., and Stintzi, A. (2005). Biosynthesis and

    metabolism of jasmonates. J. Plant Growth Regul. 23, 179–199.

    Schaller, F., and Weiler, E.W. (1997a). Enzymes of octadecanoid

    biosynthesis in plants. 12-oxo-phytodienoate 10,11-reductase. Eur.

    J. Biochem. 245, 294–299.

    Schaller, F., and Weiler, E.W. (1997b). Molecular cloning and charac-

    terization of 12-oxophytodienoate reductase, an enzyme of the

    octadecanoid signaling pathway from Arabidopsis thaliana. Structural

    and functional relationship to yeast old yellow enzyme. J. Biol. Chem.

    272, 28066–28072.

    Schlehuber, S., and Skerra, A. (2005). Lipocalins in drug discovery:

    From natural ligand-binding proteins to ‘‘anticalins’’. Drug Discov.

    Today 10, 23–33.

    Schneider, C., Boeglin, W.E., Prusakiewicz, J.J., Rowlinson, S.W.,

    Marnett, L.J., Samel, N., and Brash, A.R. (2002). Control of pros-

    taglandin stereochemistry at the 15-carbon by cyclooxygenase-1 and

    -2: A critical role for Serine-530 and Valine-349. J. Biol. Chem. 277,

    478–485.

    Schuettelkopf, A.W., and van Aalten, D.M.F. (2004). PRODRG - A tool

    for high-throughput crystallography of protein-ligand complexes. Acta

    Crystallogr. D60, 1355–1363.

    Sheldrick, G.M., and Schneider, T.R. (1997). SHELXL: High-resolution

    refinement. Methods Enzymol. 277, 319–343.

    Siedow, J.N. (1991). Plant lipoxygenase: Structure and function. Annu.

    Rev. Plant Physiol. 132, 1065–1076.

    Song, W.C., and Brash, A.R. (1991). Purification of an allene oxide

    synthase and identification of the enzyme as a cytochrome P-450.

    Science 253, 781–784.

    Song, W.C., Funk, C.D., and Brash, A.R. (1993). Molecular cloning of

    an allene oxide synthase: A cytochrome P450 specialized for the

    metabolism of fatty acid hydroperoxides. Proc. Natl. Acad. Sci. USA

    90, 8519–8523.

    Spitzer, V., Aitzemuller, K., and Vosmann, K. (1996). The seed oil of

    Bernardia puchella (Euphorbiaceae) - A rich source of vernolic acid. J.

    Am. Oil Chem. Soc. 73, 1733–1735.

    Stelmach, B.A., Müller, A., Hennig, P., Laudert, D., Andert, L., and

    Weiler, E.W. (1998). Quantitation of the octadecanoid 12-oxo-phyto-

    dienoic acid, a signalling compound in plant mechanotransduction.

    Phytochemistry 47, 539–546.

    Stenzel, I., Hause, B., Maucher, H., Pitzschke, A., Miersch, O., Ziegler,

    J., Ryan, C.A., and Wasternack, C. (2003a). Allene oxide cyclase

    dependence of the wound response and vascular bundle-specific

    generation of jasmonates in tomato - Amplification in wound signalling.

    Plant J. 33, 577–589.

    Stenzel, I., Hause, B., Miersch, O., Kurz, T., Maucher, H., Weichert,

    H., Ziegler, J., Feussner, I., and Wasternack, C. (2003b). Jasmonate

    3216 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/18/11/3201/6115538 by guest on 10 June 2021

  • biosynthesis and the allene oxide cyclase family of Arabidopsis

    thaliana. Plant Mol. Biol. 51, 895–911.

    Stintzi, A., Weber, H., Reymond, P., Browse, J., and Farmer, E.E.

    (2001). Plant defense in the absence of jasmonic acid: The role of

    cyclopentenones. Proc. Natl. Acad. Sci. USA 98, 12837–12842.

    Strassner, J., Schaller, F., Frick, U., Howe, G.A., Weiler, E.W.,

    Amrhein, N., Macheroux, P., and Schaller, A. (2002). Characterization

    and cDNA-microarray expression analysis of 12-oxo-phyto-dienoate

    reductase reveals differential roles for octadecanoid biosynthe-

    sis in the local versus the systemic wound response. Plant J. 32,

    585–601.

    Terwilliger, T.C. (2000). Maximum likelihood density modification. Acta

    Crystallogr. D56, 965–972.

    Towbin, H., Staehelin, T., and Gordon, J. (1979). Electrophoretic

    transfer of proteins from polyacrylamide gels to nitrocellulose sheets:

    Procedure and some applications. Proc. Natl. Acad. Sci. USA 76,

    4350–4354.

    Ueda, J., and Kato, J. (1980). Isolation and identification of a senes-

    cence-promoting substance from wormwood (Artemisia absinthium

    L.). Plant Physiol. 66, 246–249.

    Urade, Y., and Hayaishi, O. (2000). Biochemical, structural, genetic,

    physiological, and pathophysiological features of lipocalin-type pros-

    taglandin D synthase. Biochim. Biophys. Acta 1482, 259–271.

    Van Duyne, G.D., Standaert, R.F., Karplus, P.A., Schreiber, S.L., and

    Clardy, J. (1993). Atomic structure of the human immunophilin FKBP-

    12 complexes with FK506 and rapamycin. J. Mol. Biol. 229, 105–124.

    Vick, B.A., and Zimmerman, D.C. (1981). Lipoxygenase, hydroperox-

    ide isomerase and hydroperoxide cyclase in young cotton seedlings.

    Plant Physiol. 67, 92–97.

    Vick, B.A., and Zimmerman, D.C. (1984). Biosynthesis of jasmonic acid

    by several plant species. Plant Physiol. 75, 458–461.

    Vick, B.A., and Zimmerman, D.C. (1987). Pathways of fatty acid

    hydroperoxide metabolism in spinach leaf chloroplasts. Plant Physiol.

    85, 1073–1078.

    Wasternack, C., and Hause, B. (2002). Jasmonates and octadeca-

    noids: Signals in plant stress responses and development. Prog.

    Nucleic Acid Res. Mol. Biol. 72, 165–221.

    Weiler, E.W., Laudert, D., Schaller, F.,