Monoterpene Synthases from Common Sage (Salvia officinalis)*
cDNA ISOLATION, CHARACTERIZATION, AND FUNCTIONAL EXPRESSION OF (+)-SABINENE SYNTHASE, 1,8-CINEOLE SYNTHASE, AND (+)-BORNYL DIPHOSPHATE SYNTHASE*

Mitchell L. WiseDagger , Thomas J. Savage§, Eva Katahira, and Rodney Croteau

From the Institute of Biological Chemistry, and the Department of Biochemistry and Biophysics, Washington State University, Pullman, Washington 99164-6340

    ABSTRACT
Top
Abstract
Introduction
Procedures
Results & Discussion
References

Common sage (Salvia officinalis) produces an extremely broad range of cyclic monoterpenes bearing diverse carbon skeletons, including members of the p-menthane (1,8-cineole), pinane (alpha - and beta -pinene), thujane (isothujone), camphane (camphene), and bornane (camphor) families. An homology-based polymerase chain reaction cloning strategy was developed and used to isolate the cDNAs encoding three multiproduct monoterpene synthases from this species that were functionally expressed in Escherichia coli. The heterologously expressed synthases produce (+)-bornyl diphosphate, 1,8-cineole, and (+)-sabinene, respectively, as their major products from geranyl diphosphate. The bornyl diphosphate synthase also produces significant amounts of (+)-alpha -pinene, (+)-camphene, and (±)-limonene. The 1,8-cineole synthase produces significant amounts of (+)- and (-)-alpha -pinene, (+)- and (-)-beta -pinene, myrcene and (+)-sabinene, and the (+)-sabinene synthase produces significant quantities of gamma -terpinene and terpinolene. All three enzymes appear to be translated as preproteins bearing an amino-terminal plastid targeting sequence, consistent with the plastidial origin of monoterpenes in plants. Deduced sequence analysis and size exclusion chromatography indicate that the recombinant bornyl diphosphate synthase is a homodimer, whereas the other two recombinant enzymes are monomeric, consistent with the size and subunit architecture of their native enzyme counterparts. The distribution and stereochemistry of the products generated by the recombinant (+)-bornyl diphosphate synthase suggest that this enzyme might represent both (+)-bornyl diphosphate synthase and (+)-pinene synthase which were previously assumed to be distinct enzymes.

    INTRODUCTION
Top
Abstract
Introduction
Procedures
Results & Discussion
References

The cyclization of the universal precursor geranyl diphosphate to form monocyclic and bicyclic monoterpenes is catalyzed by a group of enzymes termed monoterpene synthases (or cyclases). The biochemical transformation of geranyl diphosphate to cyclic products has been investigated using enzymes from a variety of plants, including both angiosperms (1) and gymnosperms (2-4), and a mechanistic paradigm for these transformations (Scheme 1) is well established (1, 5). Thus, geranyl diphosphate is initially ionized and isomerized to form either (3R)- or (3S)-linalyl diphosphate, depending on the particular enzyme. This step permits rotation about the C2-C3 single bond of the bound allylic isomer to the cisoid conformer which, upon subsequent ionization, promotes electrophilic attack by C1 on the C6-C7 double bond, resulting in the formation of the alpha -terpinyl cation as a central intermediate. Further transformations of this reactive intermediate may be effected by additional intramolecular electrophilic additions, hydride shifts, or other rearrangements before termination of the sequence by deprotonation of the final cation or capture by an external nucleophile, such as a hydroxyl ion or the diphosphate group. Although the fate of the substrate has been well characterized in numerous monoterpene cyclization reactions, the molecular mechanisms by which the enzymes effect these transformations is still poorly understood.


View larger version (24K):
[in this window]
[in a new window]
 
Scheme 1.   Conversion of geranyl diphosphate to the monoterpenes of sage. Formation of the monocyclic and bicyclic products requires preliminary isomerization of geranyl diphosphate to linalyl diphosphate. Acyclic products can be formed from either geranyl diphosphate or linalyl diphosphate. OPP denotes the diphosphate moiety.

Culinary sage (Salvia officinalis) produces a number of monoterpenes, including (+)- and (-)-alpha -pinene, (+)- and (-)-beta -pinene, (+)- and (-)-camphene, (+)-sabinene, (+)- and (-)-limonene, myrcene, 1,8-cineole, and (+)-bornyl diphosphate (Scheme 1) (1). Because sage produces this broad range of acyclic, monocyclic, and bicyclic monoterpenes, including several olefin isomers, a cyclic ether and a diphosphate ester, this plant has provided an ideal system for the study of a variety of synthases, all of which utilize the same substrate but produce different products by variations on a single reaction mechanism (1, 5). These include (+)-bornyl diphosphate synthase (the enzyme producing the precursor of (+)-camphor) (6, 7), 1,8-cineole synthase (8), (+)-sabinene synthase (the enzyme producing the precursor of (-)-3-isothujone) (9, 10), and several pinene synthases (11-14). As is typical of monoterpene cyclases (5, 15), many of these enzymes from sage appear to generate multiple products from geranyl diphosphate. Investigations with the partially purified native enzymes have suggested that a single enzyme, termed (+)-pinene synthase (cyclase I), is responsible for the synthesis of both (+)-alpha -pinene and (+)-camphene, with lesser amounts of (+)-limonene and myrcene, whereas a second enzyme, (-)-pinene synthase (cyclase II), has been shown to produce (-)-alpha -pinene, (-)-beta -pinene, and (-)-camphene, with minor amounts of (-)-limonene, terpinolene, and myrcene (11, 12). More recently, a third synthase from sage, termed cyclase III, has been described which produces a mixture of (+)-alpha -pinene and (+)-beta -pinene, along with minor amounts of myrcene (13, 14). Evidence that these reactions are catalyzed by individual multifunctional enzymes is provided by co-purification and differential inhibition studies (12), as well as by isotopically sensitive branching experiments (13, 15, 16). Despite considerable effort, the (+)-pinene synthase has never been chromatographically separated from the aforementioned (+)-bornyl diphosphate synthase (17), nor has the (-)-pinene synthase been fully resolved from 1,8-cineole synthase, although stereochemical considerations indicate that the latter two are probably distinct enzyme species (8, 18).

In this report, we describe the homology-based cloning, and subsequent sequencing and heterologous expression, of three monoterpene synthase cDNA genes from sage, the recombinant enzymes from which produce three different major types of cyclic monoterpene products, (+)-sabinene (a bicyclic olefin), 1,8-cineole (a bicyclic ether), and (+)-bornyl diphosphate (a bicyclic diphosphate ester), respectively (Scheme 1). Comparison of the sizes, subunit architectures, and product distributions of these multiple-product enzymes clarifies the assignment of specific catalytic functions to defined monoterpene synthases, and the deduced sequences provide information on the relatedness of these enzymes within the species and to other terpenoid synthases of plant origin. Additionally, comparison of the primary structures of these mechanistically different monoterpene synthases from the same plant, the first examples of this type thus far available, allows preliminary assessment of active site interactions and provides the foundation for more detailed study of structure-function relationships.

    EXPERIMENTAL PROCEDURES
Top
Abstract
Introduction
Procedures
Results & Discussion
References

Plant Materials, Substrate, and Reagents-- Sage plants (S. officinalis L.) were grown from seed as described previously (19). [1-3H]Geranyl diphosphate (250 Ci/mol) was prepared by an established method (8). Terpenoid standards were from our own collection. Unless otherwise stated, all reagents were obtained from Sigma or Aldrich Chemical Co. DNA sequences were assembled and analyzed using GCG software (20).

cDNA Library Construction, PCR1-based Probe Generation, and Library Screening-- Approximately 15 g of emerging sage leaves (shoot tips) from 3-week-old plants were ground to a fine powder in liquid nitrogen and extracted into buffer composed of 200 mM Tris-HCl (pH 8.5), 300 mM LiCl, 5 mM thiourea, 1 mM aurintricarboxylic acid, 10 mM dithiothreitol, and 10 mM EDTA, and containing 1% (w/v) polyvinylpyrrolidone (Mr ~40,000). Total RNA thus extracted was prepared by precipitation with isopropyl alcohol, followed by CsCl density gradient centrifugation, as described previously (21). Poly(A)+ mRNA was isolated by chromatography on oligo(dT)-cellulose (Qiagen) and 6.3 µg of the resulting mRNA was used to construct a lambda ZAPII cDNA library according to the manufacturer's instructions (Stratagene).

A general strategy for the homology-based PCR cloning of terpenoid synthases of higher plant origin has been suggested (22) based upon a comparison of the cDNA sequences of a monoterpene synthase (23), a sesquiterpene synthase (24), and a diterpene synthase (25) of angiosperm origin. Thus, three regions of deduced amino acid sequence (corresponding to residues 180-188, 197-203, and 380-387 of limonene synthase from spearmint (23)) were employed to design primers corresponding to 1F, 5'-A(G/A)(G/A)A(C/T)GA(G/A)(G/A)AIGGI(G/A)A(G/A)TA(C/T)AA(G/A)GA-3'; 2F, 5'-ATG(T/C)TICA(G/A)(C/T)TITA(T/C)GA(G/A)GC-3'; and 3R 5'-CTI(G/T)(C/T)I(G/A)AIGGICT(G/A)AT(G/A)TAC(G/T)T(C/T)-3'. Using purified sage leaf cDNA library phage as template (5 µl at 1.5 × 109 plaque-forming units/ml), PCR was performed under a wide range of amplification conditions (26, 27) in a total volume of 50 µl containing 10 mM Tris-HCl (pH 9.0), 50 mM KCl, 2.5 mM MgCl2, 200 µM of each dNTP, 0.5 µM of each primer, and 2.5 units of Taq polymerase (Life Technologies, Inc.). Analysis of the PCR reaction products by agarose gel electrophoresis (26) indicated that only the combination of primers 2F and 3R amplified a discrete product of approximately 600 bp, which was ligated into pT7Blue (Novagen), and transformed into Escherichia coli NovaBlue cells. Plasmid DNA was prepared from 32 individual transformants; seven of these had inserts of the predicted size (~600 bp). These inserts were partially sequenced (DyeDeoxy Terminator Cycle Sequencing; Applied Biosystems) to reveal two distinct "terpenoid synthase-like" sequences.

The relative ability of these two potential probes to hybridize with expressed genes was evaluated by DNA-RNA hybridization. Two samples of sage leaf mRNA isolated as above (3 µg each) were electrophoresed on 1% (w/v) agarose under denaturing conditions and blotted onto separate polyvinylidene difluoride membranes using standard techniques (26). Each membrane was evaluated with 32P-labeled probe, generated from one or the other of the 600-bp fragments using random hexamer priming (28), by standard hybridization and washing protocols (26). Autoradiography of the membrane revealed that both probes hybridized to a 2-kilobase transcript, although one probe generated a significantly stronger signal (~10-fold) than the other. This probe was subsequently employed to screen the cDNA library in an attempt to isolate full-length cDNA sequences encoding the corresponding presumptive terpene synthase. UV cross-linked nitrocellulose lifts containing 3-5 × 104 primary plaques (plated on E. coli XL1-Blue-MRF'), after pre-hybridization (in 1.25 × SSPE, 0.5 × Denhart's reagent, 9% formamide, 0.002% SDS, and 10 µg/ml denatured E. coli DNA, for 2 h at 42 °C), were hybridized in the same medium with approximately 8 µCi of the 32P-labeled probe for 48 h. Filters were washed, first at room temperature (in 2 × SSC with 0.1% SDS) then at 55 °C (in 1 × SSC with 0.1% SDS), and subsequently exposed to x-ray film at -70 °C (26). Plaques yielding positive signals were purified through two additional rounds of hybridization. A total of 77 purified lambda ZAP clones so isolated were excised in vivo to generate BluescriptII SK(-) phagemids and transformed into E. coli SOLR cells according to the Stratagene protocol. The size of each cDNA insert was determined by PCR using T3 and T7 promoter primers, and transformed clones containing an insert >1.6 kilobases were either expressed to assay for monoterpene synthase activity or sequenced at the 5' terminus using the T3 promoter primer. Bluescript plasmids expressing monoterpene synthase activity in cell-free extracts of transformed E. coli (see below) were fully sequenced on both DNA strands by primer walking or by the method of nested deletions using exonuclease III and mung bean nuclease (26).

To improve the efficiency of functional expression and facilitate subsequent enzyme purification, each of the apparent full-length pBluescript clones that expressed monoterpene synthase activity was subcloned, in-frame, into pGEX vectors (Pharmacia) using a convenient BamHI (SBS and SSS) or EcoRI (SCS) restriction site at the 5'-end, and the XhoI restriction site at the 3' terminus. Fidelity in subcloning was confirmed by complete sequencing, and these plasmid constructs were expressed in E. coli XL1-Blue-MFR' cells.

cDNA Expression in E. coli-- The Bluescript plasmids expressed in E. coli strain XL1-Blue were grown in 5 ml of LB medium (26), supplemented with 100 µg of ampicillin/ml, to an A600 = 0.5 at 37 °C with constant shaking, then induced with 1 to 3 mM isopropyl-1-thio-beta -D-galactopyranoside. The cells were allowed an additional 4 h growth at 37 °C before harvesting by centrifugation (2000 × g, 10 min) and lysis by sonication (Braun-Sonic 2000 with microprobe at maximum power for 15 s), on ice, in 50 mM Mopso buffer containing 10% glycerol, 10 mM MgCl2, and 5 mM dithiothreitol (either pH 6.5 or 7.1, as appropriate for the assays described below). The sonicates were cleared by centrifugation (18,000 × g, 10 min) and the resulting supernatant was used as the enzyme source. The pGEX constructs in E. coli XL1-Blue-MFR' cells were similarly grown at 37 °C to A600 = 1.0 to 1.5, then induced with 1 mM isopropyl-1-thio-beta -D-galactopyranoside and incubated overnight at 20 °C with constant shaking. The cells were then harvested and lysed, and the soluble supernatant prepared as before. Purification of the resulting fusion proteins was attempted using the glutathione-Sepharose affinity column according to the manufacturer's instructions (Pharmacia). Of the three expressed monoterpene synthases (see below), only one (SBS) bound to the matrix but, even in this case, affinity-based purification proved to be unreliable. Therefore, partial purification of the heterologously expressed synthases was achieved by ion-exchange chromatography on O-diethylaminoethyl-cellulose (Whatman DE-52) using a 0-400 mM NaCl gradient. The partially purified preparations were desalted by repeated ultrafiltration and dilution using an Amicon Centriprep 30 concentrator (30 kDa cutoff) and the appropriate assay buffer. The pGEX-expressed fusion proteins were also subjected to gel permeation chromatography (Pharmacia FPLC system) using a Pharmacia XY 16 × 70 column packed with Superdex S-200 and equilibrated with the appropriate 50 mM Mopso buffer system. The column was developed at a flow rate of 0.3 ml/min and calibrated using the Sigma MW-GF-200 molecular weight marker kit. Kav values of the recombinant enzymes were compared with the calibration standards to establish molecular weights (29), which were then corrected for the engineered fusion and transit peptide to estimate the molecular weight of the corresponding native form.

Enzyme Assay-- Monoterpene synthase activities were assayed by methods previously described (6, 8, 12, 30). Briefly, an aliquot of the bacterial cell lysate, appropriate column fractions, or partially purified and desalted enzyme preparation, in 0.5 or 1.0 ml of 50 mM Mopso buffer (pH 6.1 for SBS, and 7.1 for SCS and SSS) containing 10 mM MgCl2, 5 mM dithiothreitol, and 10% (v/v) glycerol, was transferred to a 7-ml glass Teflon sealed, screw-capped tube, and the mixture was overlaid with 1 ml of pentane to trap volatile products. The reaction was initiated by the addition of 4.5 µM [1-3H]geranyl diphosphate (1.3 µCi), with incubation at 31 °C with gentle shaking for 0.5 to 3.0 h. The pentane layer and an additional pentane extract (2 × 1 ml) were passed over a short column of silica gel surmounted by anhydrous MgSO4 (in a Pasteur pipette) to afford the monoterpene olefin fraction. Subsequent extraction of the remaining aqueous phase with diethyl either (2 × 1 ml), and passage of this extract through the same column, yielded the oxygenated monoterpene fraction. The residual aqueous phase was then treated with excess potato apyrase and wheat germ acid phosphatase to hydrolyze monoterpenol diphosphate esters (6, 30). The liberated alcohols were then extracted into diethyl ether (2 × 1 ml) and the combined extract dried over anhydrous MgSO4. Radioactivity in the various fractions was determined by liquid scintillation counting of aliquots (Packard 460 CD with external standard quench correction) and the remaining material was concentrated for radio-GC and GC-MS analysis.

Kinetic analyses were carried out with the partially purified, recombinant pGEX fusion proteins by determination of initial reaction rates at a minimum of 10 substrate concentrations ranging from 0.45 to 45 µM [1-3H]geranyl diphosphate, at saturating levels of the divalent metal ion cofactor. The results were analyzed by non-linear regression of the Michaelis-Menten equation using the curve-fitting capabilities of Sigma-Plot (Jandel Corp.).

Product Identification-- Radio-GC was performed on a Gow-Mac 550P gas chromatograph with thermal conductivity detector directly coupled to a Nuclear-Chicago 8731 gas proportional counter (31). An AT-1000 packed column (Alltech) was used with helium as carrier at 30 ml/min and with temperature programming from 70 to 200 °C (at 5 °C/min) for analysis of monoterpene olefins, and from 100 to 200 °C (at 5 °C/min) for analysis of oxygenated monoterpenes. Authentic standards (10-20 µg/component) were included with each injection in order to correlate the retention times determined by mass and radioactivity detectors.

GC-MS was performed on a Hewlett-Packard 6890 GC-quadrupole mass selective detector system interfaced with a Hewlett-Packard Chemstation for data analysis. An Alltech AT-1000 fused silica capillary column (30 m × 0.25 mm inner diameter) was employed. Inleting was done by cool, on-column injection at 40 °C, with oven programming from 40 °C (50 °C/min) to 50 °C (5 min hold) then to 160 °C (10 °C/min), under a constant flow of 0.7 ml of helium/min. Full spectra were recorded for major reaction products which were identified by comparison of retention times to authentic standards and by comparison of spectra to those of the NSB75K library using the G1033A NIST probability based matching algorithm. Chiral phase separations were performed on a Hewlett-Packard 5890 GC by split injection (80:1) on a 30-m cyclodex-B capillary column (J & W Scientific) using H2 as carrier at 0.6 ml/min and temperature programming from 70 to 200 °C at 10 °C/min with flame ionization detection. Compound identification was based on retention time identity with the authentic standard.

    RESULTS AND DISCUSSION
Top
Abstract
Introduction
Procedures
Results & Discussion
References

Similarity-based Cloning of Monoterpene Synthases from S. officinalis-- Because S. officinalis (culinary or common sage) produces such a broad structural variety of monoterpenes, this species has been utilized extensively for studies on the enzymology, stereochemistry, and mechanism of monoterpene cyclization reactions (1, 5). Structural analyses of the responsible monoterpene synthases, and more detailed study of the cyclization mechanisms, require the isolation of cDNA species encoding these target enzymes. Protein purification from sage, as the basis for cDNA isolation, has been of limited success (17) because of the number of synthases present and their similarity in physical properties (32), and thus far has not permitted a reverse genetic approach to cloning of any of the monoterpene synthases from this species.

As a possible alternative to protein-based cloning of terpene synthases, a homology-based PCR strategy was recently proposed (22) that was developed by comparison of the deduced amino acid sequences of cDNAs encoding a monoterpene synthase (23), a sesquiterpene synthase (24), and a diterpene synthase (25) of phylogenetically distant angiosperm species. Three conserved regions of sequence were identified that appeared to be useful for the design of degenerate PCR primers. Two of these primers ultimately amplified a 600-bp fragment using cDNA from a sage leaf library as template. Cloning and sequencing showed the amplified products to correspond to two distinct sequence groups, both of which showed similarity to sequences of cloned terpene synthases, but only one of which hybridized strongly to a 2-kilobase target upon Northern blot analysis of sage leaf mRNA. This more efficient probe was utilized to screen the sage leaf cDNA library, from which 77 positive phagemids were purified. Size selection of the purified and in vivo excised clones yielded a subset of 44 with insert size greater than 1.6 kilobases, and these were expressed in E. coli XL1-Blue cells and the resulting extracts were assayed for functional monoterpene synthase activity by monitoring the conversion of [1-3H]geranyl diphosphate to monoterpene olefins, oxygenated monoterpenes, and monoterpenyl diphosphate esters.

Nine functionally active clones were identified by this means, four types of which showed unique product profiles. Two cDNA clones, of which the clone designated 3C6 was most active in expression, yielded an enzyme in the corresponding bacterial extracts that produced principally bornyl diphosphate from geranyl diphosphate. This recombinant enzyme, designated SBS (sage bornyl diphosphate synthase), was presumed to represent the native (+)-bornyl diphosphate synthase of sage, one of the prominent enzymes of oil gland extracts of this species (6) that produces the first dedicated intermediate in (+)-camphor biosynthesis (7, 33, 34). Four clones, of which clone 3B5 yielded the highest activity, expressed a synthase in bacterial extracts that converted geranyl diphosphate to 1,8-cineole as the major product. This acquisition was designated SCS (sage 1,8-cineole synthase) and considered to represent the native 1,8-cineole synthase, an enzyme for which the mechanism of cyclization has been studied in detail (8, 35). Two additional clones, of which clone 3F25 was seemingly the most active in expression, yielded E. coli extracts capable of transforming geranyl diphosphate to sabinene as the dominant olefin product. This acquisition was named SSS (sage sabinene synthase), with correspondence assigned to the native (+)-sabinene synthase which catalyzes the cyclization to the bicyclic olefin precursor of (-)-isothujone (9, 10, 36). The last functional clone (3M13) yielded an expressed activity that catalyzed the conversion of geranyl diphosphate to the monocyclic olefin limonene as essentially the only product.

Clones 3C6, 3B5, 3F25, and 3M13 were fully sequenced (see Fig. 1 for deduced amino acid alignments), whereupon it was shown that the insert of 3M13 was a 5'-truncated version of 3F25 (data not shown), resulting in the translation of a shortened fusion protein starting at residue 106 of the sabinene synthase (SSS) that apparently leads to derailment of the normal bicyclization to instead yield the monocyclic product limonene. The remaining 35 positive clones, that were inactive in expressing a functional monoterpene synthase activity, were partially sequenced from the 5' terminus to search for additional truncated versions or inserts that were out of frame for proper expression. This approach revealed three truncated and/or out of frame versions of SBS, 20 additional versions of SCS, and 11 versions of SSS, while unveiling one new clone (number 3F5 and designated SUS). This clone showed extensive sequence homology to 3C6, 3B5, and 3F25 but encoded a premature stop codon (Fig. 1) that would result, because of this cloning artifact, in translational truncation toward the carboxyl terminus. A functional form of the SUS protein has not yet been obtained.


View larger version (141K):
[in this window]
[in a new window]
 
Fig. 1.   Alignment of deduced amino acid sequences of monoterpene synthases. The designations correspond to: A.g.myrc, A. grandis myrcene synthase (43); A.g.limo, A. grandis limonene synthase (43); A.g.pine, A. grandis (-)-alpha -pinene synthase (43); P.f.limo, P. frutescens limonene synthase (42); M.s.limo, M. spicata limonene synthase (23); S.o.born, S. officinalis bornyl diphosphate synthase (AF051900); S.o.cine, S. officinalis 1,8-cineole synthase (AF051899); S.o.sabi, S. officinalis sabinene synthase (AF051901); and SUS, S. officinalis unknown synthase. The solid bars underline the conserved regions for primer construction used for synthesis of the probes. Asterisks indicate the three unique serine residues of SBS. The lack of sequence similarity in the putative plastidial transit peptides (represented by the first 55-60 residues) is notable. The alignment was created with the PILEUP program (20).

Product Profiles of Recombinant Synthases-- Since the formation of multiple products from geranyl diphosphate is a common, if unusual, feature of the monoterpene synthases (1, 5), the product profiles of the recombinant enzymes were examined in detail by radio-GC and GC-MS. Recombinant sabinene synthase (SSS) produces exclusively monoterpene olefins, which by radio-GC analysis (Fig. 2) were identified as sabinene (63%), gamma -terpinene (21%), terpinolene (7.0%), limonene (6.5%), and myrcene (2.5%). The major products of this enzyme (sabinene and gamma -terpinene) are formed by a cyclization mechanism involving a 1,2-hydride shift in the alpha -terpinyl cation intermediate (Scheme 1). Identification of sabinene as the major component was confirmed by retention time identity with an authentic standard on GC-MS and by comparison of the mass spectrum to the library standard: m/z 136 (P+, 14%), 94 (P+-42, 14%), 93 (P+-43, base peak), 91 (P+-45, 40%), 80 (P+-56, 10%), 79 (P+-57, 26%), 77 (P+-59, 36%), and 69 (P+-67, 9%). Chiral phase capillary GC analysis demonstrated the biosynthetic sabinene to be coincident with authentic (+)-sabinene (data not shown); however, the (-)-enantiomer was not available for analysis to confirm the absolute configuration of this product. Previous studies have shown that cell-free extracts from sage produce only the (+)-antipode of sabinene from geranyl diphosphate (9, 10), supporting the assignment of the (+)-stereoisomer in this case; the other principal olefinic products of SSS are achiral.


View larger version (17K):
[in this window]
[in a new window]
 
Fig. 2.   Radio-GC analysis of the olefinic products of recombinant monoterpene synthases. The radio-GC profiles for the olefinic products generated from [1-3H]geranyl diphosphate are illustrated for recombinant sage sabinene synthase (SSS), 1,8-cineole synthase (SCS), and bornyl diphosphate synthase (SBS), and for a crude enzyme extract from isolated sage oil gland secretory cells. The bottom tracing is the response of the mass detector to authentic standards of alpha -pinene (1), camphene (2), beta -pinene (3), sabinene (4), myrcene (5), limonene (6), gamma -terpinene (7), and terpinolene (8).

Cineole synthase (SCS) was shown by aliquot counting and radio-GC of the various metabolite fractions (Figs. 2 and 3) to produce both oxygenated monoterpenes (1,8-cineole, 79%, with a few percent alpha -terpineol) and a mixture of olefins (~20%). Confirmation of 1,8-cineole as the major product was carried out by GC-MS to demonstrate identity of retention time and mass spectrum to the authentic standard: m/z 154 (P+, 39%), 111 (P+-43, 56%), 108 (P+-46, 65%), 93 (P+-61, 46%), 84 (P+-70, 71%), 81 (P+-73, base peak), 71 (P+-83, 81%) and 69 (P+-85, 65%). Chiral phase capillary GC analysis allowed resolution, confirmation, and quantification of the olefins (Fig. 4) as (+)-alpha -pinene (5.5% of total products), (-)-alpha -pinene (0.9%), myrcene (2.9%), sabinene (2.6%, presumably the (+)-enantiomer), (+)-beta -pinene (2.7%), (-)-beta -pinene (4.1%), (+)-limonene (1.1%), and (-)-limonene (0.4%). The stereochemistry of the enzymatic transformation leading to 1,8-cineole has been examined (8) and shown to involve the cyclization of the bound intermediate (3R)-linalyl diphosphate in anti,endo-conformation, i.e. the same overall stereochemistry required for the production of (+)-alpha -pinene, (+)-beta -pinene, and (+)-limonene (12). The formation of the (-)-series of antipodes must therefore occur via the extended anti,exo-conformation. This apparent loss of stereochemical fidelity in the production of minor amounts of the olefin by-products may be a consequence of the fact that the enzyme, expressed as the pGEX fusion of the preprotein of the native synthase, was proteolytically processed by the E. coli host to a form that could compromise substrate and intermediate binding conformations (see below).


View larger version (16K):
[in this window]
[in a new window]
 
Fig. 3.   Radio-GC analysis of the oxygenated products of recombinant monoterpene synthases. The radio-GC profiles for the oxygenated products generated from [1-3H]geranyl diphosphate are illustrated for recombinant sage 1,8-cineole synthase (SCS) and bornyl diphosphate synthase (SBS), and for a crude enzyme extract from isolated sage oil gland secretory cells. The bottom tracing is the response of the mass detector to authentic standards of limonene (1), 1,8-cineole (2), trans-sabinene hydrate (3), linalool (4), alpha -terpineol (5), borneol (6), nerol (7), and geraniol (8). Geraniol is derived from hydrolysis of the substrate by contaminating phosphatases, and nerol and linalool are derived from nonenzymatic solvolysis of the substrate during the course of the reaction.


View larger version (19K):
[in this window]
[in a new window]
 
Fig. 4.   Chiral phase capillary GC analysis of the olefinic products of recombinant monoterpene synthases. The GC profiles for the olefin enantiomers generated from geranyl diphosphate are illustrated for recombinant sage bornyl diphosphate synthase (SBS) and sage 1,8-cineole synthase (SCS). Analytical conditions using the cyclodex-B column are described under "Experimental Procedures." Identifications are based on comparison to authentic standards: (-)-alpha -pinene (1), (+)-alpha -pinene (2), myrcene (3), (-)-camphene (4), (+)-camphene (5), (-)-limonene (6), (+)-limonene (7), terpinolene (8), (+)-sabinene (9), (+)-beta -pinene (10), (-)-beta -pinene (11), unidentified (12).

Bornyl diphosphate synthase (SBS) was shown, by radio-GC evaluation of all metabolite fractions, to produce principally bornyl diphosphate (75%), as demonstrated by enzymatic hydrolysis of this product followed by separation of the derived borneol from the residual geraniol (liberated from the substrate) and from lesser amounts of nonenzymatic solvolysis products (also generated from geranyl diphosphate in the course of the analysis) (Fig. 3). The production of bornyl diphosphate by this recombinant enzyme was also demonstrated directly by radio-high performance liquid chromatographic analysis of the aqueous reaction mixture using an ion-paring, reversed-phase chromatography protocol previously established for the separation of prenyl diphosphate esters (data not shown) (37). GC-MS analysis of the derived borneol confirmed the identity of this product by comparison of retention time and mass spectrum to the authentic standard: m/z 154 (P+, <1%), 139 (P+-15, 8%), 121 (P+-33, 6%),110 (P+-44, 18%), 96 (P+-58, 8%), 95 (P+-59, base peak), 93 (P+-61, 7%), 67 (P+-87, 8%), and 55 (P+-99, 8%). Additionally, chiral phase capillary GC analysis of the derived borneol demonstrated the exclusive presence of the (+)-antipode (data not shown), as expected based on studies with the corresponding native enzyme (6, 38, 39). The recombinant (+)-bornyl diphosphate synthase was also shown, by radio-GC analysis of the olefin fraction (Fig. 2) and chiral phase GC analysis (Fig. 4), to produce a series of olefins (25% of total product) identified as (+)-alpha -pinene (3.4% of total product), (+)-camphene (9.5%), (-)-camphene (0.5%), (+)-limonene (3.9%), (-)-limonene (3.9%), terpinolene (2.1%), and myrcene (1.5%). Since formation of the (+)-olefin series is mechanistically related to the formation of (+)-bornyl diphosphate via the anti,endo-cyclization of the intermediate (3R)-linalyl diphosphate (18, 38-41) (Scheme 1), the generation of small amounts of the antipodal (-)-camphene and (-)-limonene by the recombinant cyclase again suggests some loss of stereochemical fidelity in the overall reaction sequence.

Sequence Analysis-- Alignment of the deduced amino acid sequences (Fig. 1) of SBS clone 3C6 (2025 bp, with an open reading frame of 1794 nucleotides encoding 598 amino acids for a protein of 69.3 kDa and calculated pI of 6.06), SCS clone 3B5 (1968 bp, with an open reading frame of 1773 nucleotides encoding 591 amino acids for a protein of 69.4 kDa and calculated pI of 5.79), and SSS clone 3F25 (1911 bp, with an open reading frame of 1767 nucleotides encoding 589 amino acids for a protein of 68.9 kDa and calculated pI of 5.22), and SUS truncated clone 3F5 (2022 bp), with the published sequences for (-)-limonene synthase from Mentha spicata (spearmint) (23) and Perilla frutescens (42), and three monoterpene olefin synthases from Abies grandis (grand fir) (43), illustrates that there are several regions of conservation between these eight monoterpene synthases of diverse origin. Comparison of these sequences using the GCG GAP program (20) (Table I) revealed the monoterpene synthases from sage to resemble each other and the limonene synthases from related members of the Lamiaceae (50-70% identity, 70-85% similarity) more closely than the monoterpene synthases of the gymnosperm grand fir (~32% identity), or the linalool synthase from Clarkia breweri (44)(~25% identity).

                              
View this table:
[in this window]
[in a new window]
 
Table I
Pairwise comparison of deduced amino acid sequence identity of cloned monoterpene synthases
The abbreviations used in this table are: A.g. myrc, A. grandis myrcene synthase (43); A.g. pine, A. grandis (-)-pinene synthase (43); A.g. limo, A. grandis limonene synthase (43); C.b. lino, C. breweri linalool synthase (44); P.f. limo, P. frutescens limonene synthase (42); M.s. limo, M. spicata limonene synthase (23); S.o. born, S. officinalis bornyl diphosphate synthase (AF015900); S.o. cine, S. officinalis 1,8-cincole synthase (AF051899); S.o. sabi, S. officinalis sabinene synthase (AF051901).

Monoterpene biosynthesis is compartmentalized in plastids (45-47). Thus, the monoterpene synthases are encoded as preproteins bearing an amino-terminal transit peptide for import of these nuclear gene products into plastids (leucoplasts of the oil gland cells in the present instance), where they are proteolytically processed to the mature forms. In vitro translation and plastidial processing of the gene product has been demonstrated directly with the limonene synthase cDNA from spearmint.2 In all of the monoterpene synthases (Fig. 1), the 50-60 amino-terminal residues are characterized by a low degree of similarity, typical of targeting sequences, yet they all share common features of transit peptides in being rich in serine, threonine, and small hydrophobic residues but with few acidic residues (48, 49). All native monoterpene synthases thus far examined (2, 17, 22, 23) appear to be NH2-terminally blocked, preventing direct determination (by sequencing) of the transit peptide-mature protein cleavage junction. Significantly, a tandem pair of arginine residues (e.g. Arg55-Arg56 of SBS) are strictly conserved in the deduced sequences of all of the monoterpene syntheses (Fig. 1) and they define the most NH2-terminal region of obvious homology, suggesting a possible cleavage site. Recent studies3 have shown that truncation of the recombinant spearmint limonene synthase preprotein immediately upstream of these tandem arginines yields a fully functional "pseudomature" form of the enzyme, whereas truncation downstream from this element severely impairs activity.

Downstream of the tandem arginines are several regions of homology, including the highly conserved (I, L, or V)DDXXD motif (e.g. residues Ile350-Asp355 of SBS) found in virtually all deduced sequences for enzymes that utilize prenyl diphosphate substrates (50, 51). This aspartate-rich element is now generally recognized as a binding site for the metal ion chelated diphosphate ester substrate (50, 52-55). Several other highly conserved regions are also apparent, including Arg299-Trp-Trp301, Arg372-Trp-Glu/Gln374, Tyr384-Met-Gln/Lys386, and Cys516-Tyr-Met-X-Glu/Asp520 (amino acid placement referred to SBS). The active site peptide LQLYEASFLL, previously isolated from the co-purified (+)-pinene synthase and (+)-bornyl diphosphate synthase of sage (17), was located at residues 195-204 of SBS and also at residues 187-196 of SSS. Very similar sequences in the same location were found in SCS and in the two limonene synthase sequences from M. spicata and P. frutescens (Fig. 1).

An intriguing aspect of both prenyltransferase and terpenoid cyclization catalysis (1, 56, 57) is the electrophilic nature of these reaction mechanisms, and the means by which these enzymes stabilize and guide highly reactive carbocationic intermediates without alkylation at the active site. Dougherty (58) has suggested pi -cation interactions as a means of stabilizing such reactive intermediates in enzyme-catalyzed reactions, and the concept has received some support from both sequence analysis and x-ray crystallographic studies (59-61). While far from conclusive, the number of conserved aromatic residues displayed among the monoterpene synthases does suggest the possibility of such a role for aromatic side chains in these electrophilic cyclizations.

The (+)-bornyl diphosphate synthase from sage has previously been shown to be inhibited by the "active serine"-directed reagent diisopropyl fluorophosphate, a characteristic not shared by other monoterpene synthases (6). Because of the unique utilization of the substrate diphosphate moiety as the terminating nucleophile by this enzyme (62, 63), it was hypothesized that a serine residue may be involved in binding and transfer of the diphosphate function in the course of the reaction. Sequence comparison of SBS with the other two monoterpene synthases of sage reveals five unique serine residues at positions 67, 255, 302, 454, and 469 (the serine at position 454 is conserved in monoterpene synthases from other species). Three of these serines (at positions 67, 255, and 302) are within otherwise highly conserved regions and are, therefore, obvious targets for selective covalent modification with radiolabeled diisopropyl fluorophosphate and directed mutagenesis studies.

Properties of Recombinant Synthases-- Calibrated gel permeation chromatography of the pGEX fusion form of SBS revealed a single peak of activity at an elution volume corresponding to an Mr ~200,000, indicating that the expressed fusion preprotein (corresponding to a molecular weight of about 2 × 96,300) was a functional dimer. Treatment of the SBS protein with thrombin to remove the glutathione S-transferase fusion tag, followed by re-chromatography, indicated a decrease in molecular weight to approximately 135,000, consistent with the loss of the 27-kDa transferase peptide from each subunit at a calculated molecular weigth of 69,300 for the preprotein. Further correction of the molecular mass to account for the transit peptide would yield a dimer of about 120 kDa which corresponds roughly to the native dimer molecular mass of both (+)-bornyl diphosphate synthase and (+)-pinene synthase from sage (6, 12), two enzymes which have never been satisfactorily resolved as distinct species. Although a dimeric quaternary structure is not unique to these two synthases, the vast majority of the monoterpene synthases characterized to date are monomeric (1, 5). The product profile of SBS clone 3C6 is qualitatively similar to the combination of both of these enzymes (i.e. (+)-bornyl diphosphate and the (+)-series of alpha -pinene and related olefins) (6, 12), although the quantitative distributions do not exactly match, and the stereochemistry of the olefin products is anomalous. Thus, as indicated previously, (+)-bornyl diphosphate, (+)-alpha -pinene, (+)-camphene, and (+)-limonene arise via the same overall cyclization stereochemistry, and these enantiomers are produced exclusively from geranyl diphosphate by the native (+)-bornyl diphosphate and (+)-pinene synthase activities (18, 38-41). The small amounts of (-)-limonene and (-)-camphene formed by the recombinant enzyme are attributed to antipodal cyclizations via abnormal, extended substrate conformations, as the phenomenon has been described previously, especially when using neryl diphosphate (the cis-analog of geranyl diphosphate) as an alternate substrate (40, 64). The geranyl substrate, however, was verified as >99% pure, thereby eliminating this possibility in the present instance and suggesting that loss of stereochemical fidelity (to the extent of 5% of the total product mixture) may be attributed to the presence of the glutathione S-transferase fusion peptide plus transit peptide which may alter substrate binding directly, or indirectly by compromising subunit assembly. To address this question, as well as the correct identity of the SBS protein, will require the detailed assessment of truncated enzymes that more closely resemble the native form.

Gel permeation chromatography of SCS revealed a single peak of activity at an elution volume corresponding to an Mr of 72,000, whereas SSS gave two peaks of activity, an aggregated form eluting in the void volume and a second corresponding to an Mr of 60,000. Both of these molecular weights are significantly lower than those predicted from pGEX expression-based fusion of the glutathione S-transferase (27 kDa) with the respective preproteins (both SCS and SSS ~96 kDa). Thrombin treatment was without influence on the gel permeation chromatographic behavior of these enzymes, indicating the absence of the glutathione S-transferase peptide tag and rationalizing the previously observed inability of the recombinant SCS and SSS enzymes to bind to the glutathione-affinity column. Inspection of the 5'-sequences of the corresponding pGEX constructs showed both to be free of in-frame stop codons that might have permitted polycistronic translation of the preprotein devoid of the glutathione S-transferase peptide. The apparent truncation was therefore attributed to proteolytic processing of the recombinant SCS and SSS in the E. coli host to proteins that more closely resemble the preprotein forms of the native, monomeric 1,8-cineole synthase (8, 35) and (+)-sabinene synthase (9, 10) of sage. Similar proteolytic processing of a recombinant limonene synthase preprotein from spearmint has been observed previously in this E. coli host (23).

1,8-Cineole synthase has never been satisfactorily separated from the aforementioned (-)-pinene synthase from sage but, in this instance, the product distribution of SCS does not match the product distribution of (-)-pinene synthase, either quantitatively, qualitatively, or in stereochemical terms, since the reactions catalyzed are of the opposite antipodal series (8, 11, 18). However, the product distribution of SCS shows some parallels with that of the recently described cyclase III which produces (+)-alpha -pinene and (+)-beta -pinene (13, 14). Even here, the match is not perfect and the production of anomalous products of the antipodal (-)-series (<6% of total) again suggests that substrate binding interactions may be compromised. To assess the latter possibility, the Km values for SCS (7.0 µM), SSS (7.4 µM), and SBS (3.0 µM) were determined. These values are likely somewhat high because the recombinant enzymes were not purified sufficiently to remove all contaminating phosphatases that result in some depletion of the substrate geranyl diphosphate. While the calculated Km values compare reasonably well with the literature values of 1.1 µM (8), 2.0 µM (9), and 2.0 µM (18, 41), respectively, for the corresponding native enzymes, they are sufficiently higher to suggest at least subtle alteration in binding capacity of the recombinant forms.

Although the combination of products generated by the three recombinant enzymes described in this paper represent many of the monoterpenes synthesized by extracts of sage oil gland cells, several lines of evidence suggest that more monoterpene synthases remain to be acquired from this species. Thus, clone 3F5 (SUS) has not yet been functionally expressed, one of the two probes generated by PCR has not yet located the corresponding cDNA (one probe matches SUS), and an active site peptide (17) has not yet been matched to a cDNA sequence. The lack of the full set of synthases, coupled to the production of multiple products, including abnormal enantiomers, by the extant recombinant forms, has obscured the assignment of cDNA sequences to their native monoterpene synthase counterparts. This is most vividly illustrated by clone 3C6 which is clearly the recombinant form of (+)-bornyl diphosphate synthase and may also represent the enzyme responsible for the activity previously assigned to (+)-pinene synthase. Nevertheless, it is the great mechanistic diversity of this group of enzymes, originating from a single plant species yet producing a cyclic ether, diphosphate ester, and numerous olefins, that makes this set of catalysts unique and so biochemically appealing. Although the enzymatic generation of multiple products is an unusual and complex phenomenon, it does provide a most convenient and very powerful reporting device for communicating the influence of directed mutagenic change on the cyclization reaction. Defining how the similarities and differences of these enzymes relate to the control of catalytic channeling to the different products can be expected to provide the foundation for a detailed, molecular level understanding of structure-function relationships in the monoterpene synthases.

    ACKNOWLEDGEMENTS

We thank Gerhard Munske of the Washington State University Laboratory for Bioanalysis and Biotechnology for primer synthesis and nucleotide sequencing, and Joyce Tamura-Brown for typing the manuscript.

    FOOTNOTES

* This work was supported in part by National Institutes of Health Grant GM-31354 and Project 0268 from the Agricultural Research Center, Washington State University.The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

The nucleotide sequences reported in this paper will appear in the GSDB, DDBJ, EMBL, and NCBI nucleotide sequence data banks under accession numbers AF051899 (1,8-cineole synthase), AF051900 (bornyl diphosphate synthase), and AF051901 (sabinene synthase).

Dagger Plant Biochemistry Research and Training Center Postdoctoral Fellow.

§ Present address: Mendel Biotechnology, Inc., 21375 Cabot Blvd., Hayward, CA 94543.

To whom correspondence should be addressed: Institute of Biological Chemistry, Washington State University, Pullman, WA 99164-6340. Tel.: 509-335--1790; Fax: 509-335-7643; E-mail: croteau{at}mail.wsu.edu.

1 The abbreviations used are: PCR, polymerase chain reaction; bp, base pair(s); GC, gas chromatography; Mopso, 3-(N-morpholino)-2-hydroxypropane sulfonic acid; MS, mass spectrum/spectrometry; SBS, sage bornyl diphosphate synthase; SCS, sage 1,8-cineole synthase; SSS, sage sabinene synthase; SUS, sage unknown synthase.

2 J. Gershenzon, G. W. Turner, E. Nielsen, and R. Croteau, manuscript in preparation.

3 D. C. Williams, E. J. Katahira, D. J. McGarvey, and R. Croteau, submitted for publication.

    REFERENCES
Top
Abstract
Introduction
Procedures
Results & Discussion
References

  1. Croteau, R. (1987) Chem. Rev. 87, 929-954
  2. Lewinsohn, E., Gijzen, M., and Croteau, R. (1992) Arch. Biochem. Biophys. 293, 167-173[Medline] [Order article via Infotrieve]
  3. Savage, T. J., Hatch, M. W., and Croteau, R. (1994) J. Biol. Chem. 269, 4012-4020[Abstract/Free Full Text]
  4. Savage, T. J., Ichii, H., Hume, S. D., Little, D. B., and Croteau, R. (1995) Arch. Biochem. Biophys. 320, 257-265[CrossRef][Medline] [Order article via Infotrieve]
  5. Wise, M. L., and Croteau, R. (1998) in Comprehensive Natural Products Chemistry: Isoprenoids (Cane, D. E., ed), Vol. 2, Elsevier Science, Oxford, UK, in press
  6. Croteau, R., and Karp, F. (1979) Arch. Biochem. Biophys. 198, 512-522[Medline] [Order article via Infotrieve]
  7. Croteau, R., and Karp, F. (1979) Arch. Biochem. Biophys. 198, 523-532[Medline] [Order article via Infotrieve]
  8. Croteau, R., Alonso, W. R., Koepp, A. E., and Johnson, M. A. (1994) Arch. Biochem. Biophys. 309, 184-192[CrossRef][Medline] [Order article via Infotrieve]
  9. Croteau, R. (1992) in Recent Developments in Flavor and Fragrance Chemistry (Hopp, R., and Mori, K., eds), pp. 263-273, VCH, Weinheim, Germany
  10. Croteau, R. (1992) Am. Chem. Soc. Symp. Ser. 490, 8-20
  11. Gambliel, H., and Croteau, R. (1982) J. Biol. Chem. 257, 2335-2342[Free Full Text]
  12. Gambliel, H., and Croteau, R. (1984) J. Biol. Chem. 259, 740-748[Abstract/Free Full Text]
  13. Wagschal, K. C., Pyun, H.-J., Coates, R. M., and Croteau, R. (1994) Arch. Biochem. Biophys. 308, 477-487[CrossRef][Medline] [Order article via Infotrieve]
  14. Pyun, H.-J., Wagschal, K. C., Jung, D.-I., Coates, R. M., and Croteau, R. (1994) Arch. Biochem. Biophys. 308, 488-496[CrossRef][Medline] [Order article via Infotrieve]
  15. Wagschal, K., Savage, T. J., and Croteau, R. (1991) Tetrahedron 47, 5933-5944[CrossRef]
  16. Croteau, R. B., Wheeler, C. J., Cane, D. E., Ebert, R., and Ha, H. J. (1987) Biochemistry 26, 5383-5389[Medline] [Order article via Infotrieve]
  17. McGeady, P., and Croteau, R. (1995) Arch. Biochem. Biophys. 317, 149-155[CrossRef][Medline] [Order article via Infotrieve]
  18. Croteau, R., Satterwhite, D. M., Wheeler, C. J., and Felton, N. M. (1989) J. Biol. Chem. 264, 2075-2080[Abstract/Free Full Text]
  19. Croteau, R., and Karp, F. (1976) Arch. Biochem. Biophys. 176, 734-746[Medline] [Order article via Infotrieve]
  20. Genetics Computer Group. (1996) Wisconsin Package version 9.0, Genetics Computer Group (GCG), Madison, WI
  21. Lewinsohn, E., Steele, C. L., and Croteau, R. (1994) Plant Mol. Biol. Rep. 12, 20-25
  22. Steele, C. L., Lewinsohn, E., and Croteau, R. (1995) Proc. Natl. Acad. Sci. U. S. A. 92, 4164-4168[Abstract]
  23. Colby, S. M., Alonso, W. R., Katahira, E. J., McGarvey, D. J., and Croteau, R. (1993) J. Biol. Chem. 268, 23016-23024[Abstract/Free Full Text]
  24. Facchini, P. J., and Chappell, J. (1992) Proc. Natl. Acad. Sci. U. S. A. 89, 11088-11092[Abstract]
  25. Mau, C. J. D., and West, C. A. (1994) Proc. Natl. Acad. Sci. U. S. A. 91, 8497-8501[Abstract]
  26. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory Manual, 2nd Ed., pp. 2.69-2.76, Cold Spring Harbor Laboratory, Cold Spring Harbor, NY
  27. Innis, M. A., and Gelfand, D. H. (1990) in PCR Protocols (Innis, M. A., Gelfand, D. H., Sninsky, J. J., and White, T. J., eds), pp. 3-12 and 253-258, Academic Press, San Diego, CA
  28. Tabor, S., Struhl, K., Scharf, S. J., and Gelfand, D. H. (1991) in Current Protocols in Molecular Biology (Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G., Smith, J. A., and Struhl, K., eds), pp. 3.5.9-3.5.10, John Wiley & Sons, New York
  29. Cooper, T. G. (1977) The Tools of Biochemistry, pp. 169-193, John Wiley & Sons, New York
  30. Croteau, R., and Cane, D. E. (1985) Methods Enzymol. 110, 383-405
  31. Satterwhite, D. M., and Croteau, R. (1988) J. Chromatogr. 454, 61-73
  32. Alonso, W. R., and Croteau, R. (1993) Methods Plant Biochem. 9, 239-260
  33. Croteau, R., and Karp, F. (1977) Arch. Biochem. Biophys. 184, 77-86[Medline] [Order article via Infotrieve]
  34. Croteau, R., Hooper, C. L., and Felton, M. (1978) Arch. Biochem. Biophys. 188, 182-193[Medline] [Order article via Infotrieve]
  35. Croteau, R., and Karp, F. (1977) Arch. Biochem. Biophys. 179, 257-265[Medline] [Order article via Infotrieve]
  36. Karp, F., and Croteau, R. (1982) Arch. Biochem. Biophys. 216, 616-624[Medline] [Order article via Infotrieve]
  37. McCaskill, D., and Croteau, R. (1993) Anal. Biochem. 215, 142-149[CrossRef][Medline] [Order article via Infotrieve]
  38. Croteau, R., Felton, N. M., and Wheeler, C. J. (1985) J. Biol. Chem. 260, 5956-5962[Abstract/Free Full Text]
  39. Croteau, R., Satterwhite, D. M., Cane, D. E., and Chang, C. C. (1986) J. Biol. Chem. 261, 13438-13445[Abstract/Free Full Text]
  40. Croteau, R., Satterwhite, D. M., Cane, D. E., and Chang, C. C. (1988) J. Biol. Chem. 263, 10063-10071[Abstract/Free Full Text]
  41. Croteau, R., Gershenzon, J., Wheeler, C. J., and Satterwhite, D. M. (1990) Arch. Biochem. Biophys. 277, 374-381[Medline] [Order article via Infotrieve]
  42. Yuba, A., Yazaki, K., Tabata, M., Honda, G., and Croteau, R. (1996) Arch. Biochem. Biophys. 332, 280-287[CrossRef][Medline] [Order article via Infotrieve]
  43. Bohlmann, J., Steele, C. L., and Croteau, R. (1997) J. Biol. Chem. 272, 21784-21792[Abstract/Free Full Text]
  44. Dudareva, N., Cseke, L., Blanc, V. M., and Pichersky, E. (1996) Plant Cell 8, 1137-1148[Abstract/Free Full Text]
  45. Gleizes, M., Pauly, G., Carde, J.-P., Marpeau, A., and Bernard-Dagan, C. (1983) Planta 159, 373-381
  46. Mettal, U., Boland, W., Beyer, P., and Kleinig, H. (1988) Eur. J. Biochem. 170, 613-616[Abstract]
  47. Perez, L. M., Pauly, G., Carde, J.-P., Belingheri, L., and Gleizes, M. (1990) Plant Physiol. Biochem. 28, 221-229
  48. Keegstra, K., Olsen, J. J., and Theg, S. M. (1989) Annu. Rev. Plant Physiol. Plant Mol. Biol. 40, 471-501[CrossRef]
  49. von Heijne, G., Steppuhn, J., and Herrmann, R. G. (1989) Eur. J. Biochem. 180, 535-545[Abstract]
  50. Chen, A., Kroon, P. A., and Poulter, C. D. (1994) Protein Sci. 3, 600-607[Abstract/Free Full Text]
  51. Chen, X.-Y., Chen, Y., Heinstein, P., and Davisson, V. J. (1995) Arch. Biochem. Biophys. 324, 255-266[CrossRef][Medline] [Order article via Infotrieve]
  52. Ashby, M. N., and Edwards, P. A. (1990) J. Biol. Chem. 265, 13157-13164[Abstract/Free Full Text]
  53. Tarshis, L. C., Yan, M., Poulter, C. D., and Sacchettini, J. C. (1994) Biochemistry 33, 10871-10877[Medline] [Order article via Infotrieve]
  54. Cane, D. E., Sohng, J. K., Lamberson, C. R., Rudnicki, S. M., Wu, Z., Lloyd, M. D., Oliver, J. S., and Hubbard, B. R. (1994) Biochemistry 33, 5846-5857[Medline] [Order article via Infotrieve]
  55. Tarshis, L. C., Proteau, P. J., Kellogg, B. A., Sacchettini, J. C., and Poulter, C. D. (1996) Proc. Natl. Acad. Sci. U. S. A. 93, 15018-15023[Abstract/Free Full Text]
  56. Poulter, C. D., and Rilling, H. C. (1981) in Biosynthesis of Isoprenoid Compounds (Porter, J. W., and Spurgeon, S. L., eds), Vol. 1, pp. 161-224, John Wiley & Sons, New York
  57. Cane, D. E. (1990) Chem. Rev. 90, 1089-1103
  58. Dougherty, D. A. (1996) Science 271, 163-168[Abstract]
  59. Sussman, J. L., Harel, M., Frolow, F., Oefner, C., Goldman, A., Toker, L., and Silman, I. (1991) Science 253, 872-879[Medline] [Order article via Infotrieve]
  60. Shi, Z., Buntel, C. J., and Griffin, J. H. (1994) Proc. Natl. Acad. Sci. U. S. A. 91, 7370-7374[Abstract]
  61. Poralla, K., Hewelt, A., Prestwich, G. D., Abe, I., Reipen, I., and Sprenger, G. (1994) Trends Biochem. Sci. 19, 157-158[CrossRef][Medline] [Order article via Infotrieve]
  62. Cane, D. E., Saito, A., Croteau, R., Shaskus, J., and Felton, M. (1982) J. Am. Chem. Soc. 104, 5831-5833
  63. Croteau, R., Shaskus, J. J., Renstrøm, B., Felton, N. M., Cane, D. E., Saito, A., and Chang, C. (1985) Biochemistry 24, 7077-7085[Medline] [Order article via Infotrieve]
  64. Croteau, R., and Satterwhite, D. M. (1989) J. Biol. Chem. 264, 15309-15315[Abstract/Free Full Text]


Copyright © 1998 by The American Society for Biochemistry and Molecular Biology, Inc.