You are on page 1of 11

pubs.acs.

org/biochemistry Article

Structurally Guided Reprogramming of Valerenadiene Synthase


Garrett E. Zinck and Joe Chappell*
Cite This: Biochemistry 2021, 60, 3868−3878 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Valerena-1,10-diene synthase (VDS) catalyzes the con-


version of the universal precursor farnesyl diphosphate into the unusual
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

sesquiterpene valerena-1,10-diene (VLD), which possesses a unique


isobutenyl substituent group. In planta, one of VLD’s isobutenyl terminal
Downloaded via UNIV ESTADUAL PAULISTA on January 22, 2022 at 03:11:26 (UTC).

methyl groups becomes oxidized to a carboxylic acid forming valerenic


acid (VA), an allosteric modulator of the GABAA receptor. Because a
structure−activity relationship study of VA for its modulatory activity is
desired, we sought to manipulate the VDS enzyme for the biosynthesis of
structurally diverse scaffolds that could ultimately lead to the generation of
VA analogues. Using three-dimensional structural homology models,
phylogenetic sequence comparisons to well-characterized sesquiterpene
synthases, and a substrate−active site contact mapping approach, the
contributions of specific amino acid residues within or near the VDS
active site to possible catalytic cascades for VLD and other sesquiterpene
products were assessed. An essential role of Tyr535 in a germacrenyl route
to VLD was demonstrated, while its contribution to a family of other
sesquiterpenes derived from a humulyl route was not. No role for Cys415
or Cys452 serving as a proton donor to reaction intermediates in VLD biosynthesis was observed. However, a gatekeeper role for
Asn455 in directing farnesyl carbocations down all-trans catalytic cascades (humulyl and germacrenyl routes) versus a cisoid cascade
(nerolidyl route) was demonstrated. Altogether, these results have mapped residues that establish a context for the catalytic cascades
operating in VDS and future manipulations for generating more structurally constrained scaffolds.

T erpenes and terpenoids arise from the condensation of


ubiquitous five-carbon isoprenoid building blocks iso-
pentyl pyrophosphate and dimethylallyl pyrophosphate into
bond,6 or by a nucleophilic addition (alkylation) from a side
chain of an amino acid within the active site.7 Interestingly,
TPSs utilize one or various combinations of these reaction
linear precursors that undergo complex regio- and stereo- mechanisms to produce some of the world’s most renowned
chemical cyclization reactions catalyzed by terpene synthases therapeutics, agrochemicals, fragrances, and nutraceuticals.8
(TPSs). The elegance of these cyclized terpenes is most easily From a biotechnological point of view, to access the diversity
captured by noting the multiple ring systems, multiple of terpenoid scaffolds9 it is imperative to gain a fundamental
stereocenters, and variety of substituent groups captured in comprehension of the primary catalytic cascades the TPS uses
small molecules of ≤500 Da.1−4 Estimates of the number of to cyclize the ionized isoprenoid substrate into highly desired
naturally occurring terpenes identified to date exceed 100,000, terpene product(s).1 This can be achieved through structure−
but this accounts for only 1−10% of the theoretical function mapping of key amino acid residues responsible for
possibilities.1−4 To achieve such chemodiversity, TPSs begin the catalytic plasticity found within the TPS.10−15 Previously,
with the generation of a highly reactive carbocation the Ro and Chappell groups independently characterized
intermediate species through the ionization of their allylic valerena-1,10-diene synthase (VDS) from Valeriana off icina-
diphosphate substrates (Scheme 1). Then, through a series of lis.16,17 VDS is a class I sesquiterpene synthase that converts
well-defined intramolecular bond formations; stereospecific the acyclic, all-trans farnesyl diphosphate (FPP) into a five- and
hydride, methyl, and methylene migrations; and various ring six-fused member ring system with an unusual isobutenyl
openings and closings, TPSs form penultimate terpene and
terpenoid scaffolds. The catalytic cascade is finally halted and Received: August 3, 2021
the hydrocarbon product released from the active site when Revised: November 29, 2021
the reactive carbocation intermediate is either quenched by the Published: December 13, 2021
capture of a water molecule within the active site or supplied
by the bulk solvent (generating hydroxylated terpene species5),
when a stereospecific proton is abstracted to form a double

© 2021 American Chemical Society https://doi.org/10.1021/acs.biochem.1c00523


3868 Biochemistry 2021, 60, 3868−3878
Biochemistry pubs.acs.org/biochemistry Article

Scheme 1. Proposed Catalytic Cascades for the Sesquiterpene Reaction Products Generated by VDS and VDS Mutantsa

a
Valerena-1,10-diene (17) is the predominate reaction product of the wild type VDS and can arise from two possible routes, initiated by either a
1,10 (germacrenyl route, red arrows and highlighting) or a 1,11 (humulyl route, blue arrows and highlighting) ring closure of 1.16,17,48 The other
major products, 9 and 10, generated by wild type VDS, could arise from a germacrenyl cation route as suggested previously.62,67 An earlier study
also used specifically labeled FPP to suggest the biosynthetic origin of α-gurjunene (12) via a germacrenyl route.68 Mutant VDS enzymes lose their
ability to biosynthesize the dominant products (9, 10, and 17) and instead generate a diverse set of sesquiterpenes: germacrene D (8) (germacrenyl
cation route);69 α-humulene (5) and β-caryophyllene (11) (humulyl cation route);51 and α-copaene (14), δ-cadinene (15), and cubebol (16),
possibly arising from a helmintogermacrenyl carbocation (green arrows and highlighting) (6).67 DCCR, the cyclopropylcarbinyl cation−
cyclopropylcarbinyl cation rearrangement,2 was suggested by Paknikar et al.48 Details of paths a−c are provided in ref 17.

substituent group known as valerena-1,10-diene (VLD). key amino acids within VDS that mapped to reaction products
Therapeutically, valerenic acid (VA), VLD’s carboxylated that offered structural constraints at VLD’s isobutenyl side
derivative that arises from the action of a cytochrome chain (Scheme 1). Using the information and experimental
P450(s) localized to the roots of V. off icinalis,18,19 was strategies from previous efforts to alter TPS specificity,12,27 we
shown to allosterically modulate the mammalian GABAA discovered amino acid substitutions that altered the product
receptor in vitro and mitigate the physical signs of stress in specificity toward native, structurally constrained terpene
murine animal models.20−22 In silico models predict VA’s reaction products or completely derailed the catalytic
flexible isobutenyl acidic tail forms hydrophobic interactions trajectory toward abortive, structurally constrained reaction
with specific amino acid residues within the β subunits of the products not previously identified for VDS.
GABAA receptor.23 While the Khom group had moderate
success in generating valerenic acid analogues with improved
selectivity and potency for the β subunit in vitro, none of these
■ MATERIALS AND METHODS
Molecular Modeling. Homology modeling of enzymes
analogues physically constrained the flexible isobutenyl side was carried out with MODELER 9.21 (University of
chain.24 The literature supports the theory that if one California, San Francisco, CA)28 using the Nicotiana tabacum
constrains a ligand (for example, a drug), one lowers the (tobacco) 5-epi-aristolochene synthase (EAS) coordinates
entropic barrier of reaching its “active” binding conformation, [Protein Data Bank (PDB) entry 5EAT]29 as the template.
naturally enhancing its ability to bind to its target receptor.25,26 The model was evaluated for accuracy with freely available
One such way to introduce conformational constraint is SaliLab Model Evaluation version 2.1 (modbase.compbio.ucsf.
through the stabilization of the ligand’s side chains via edu/evaluation).30−32 The model with the smallest predicted
cyclization, which, by default, reduces the ligand’s number of root-mean-square deviation and DOPE score was subsequently
rotatable bonds.25 Hence, the aim of this work was to identify used for docking studies. Proposed reaction intermediates were
3869 https://doi.org/10.1021/acs.biochem.1c00523
Biochemistry 2021, 60, 3868−3878
Biochemistry pubs.acs.org/biochemistry Article

constructed in CHEMDRAW (PerkinElmer) and energy- The cultures were then induced with 0.4 mM IPTG and
minimized using MOPAC2016 (Stewart Computational incubated at 23 °C for 20 h before the cells were collected at
Chemistry, Colorado Springs, CO).33 Proposed reaction 6000g for 10 min. The cell pellets were stored at −80 °C until
intermediates were docked into the VDS model using they were processed.
Autodock Vina version 1.1.2 (The Scripps Research Institute, Cell pellets were lysed in 50 mM NaH2PO4 (pH 7.4), 300
La Jolla, CA).34 Images were made using PYMOL version 1.8 mM NaCl, 5 mM imidazole, and 5% (v/v) glycerol by
(DeLano Scientific, South San Francisco, CA).35 sonification (30% power, five bursts, 20 s each, 1 min rest
Site-Directed Mutagenesis of VDS. Site-directed muta- between bursts) using a QSonica 500 instrument equipped
genesis was carried out using a modified version of the with a 1.6 mm probe. The cell debris was removed by
QuikChange Lightning Site-Directed Mutagenesis Kit (Agi- centrifugation at 21000g for 30 min at 4 °C. Total protein
lent) according to the manufacturer’s instructions. In brief, concentrations were estimated by the Bradford assay (Bio-
PrimeSTAR HS DNA Polymerase (Takara) was used in a Rad) using bovine serum albumin as the standard.
“round-the-horn” in vitro polymerase chain reaction muta- Steady-State Kinetic Determinations. Michaelis−Ment-
genesis reaction (98 °C for 2 min, 55 °C for 15 s, and 72 °C en kinetics were determined for each mutant enzyme and
for 7 min for 18 cycles; 72 °C for 7 min final extension for one compared with that of the wild type using radiolabeled
cycle). DpnI digestion of the template plasmid and trans- [1-3H]FPP (20 Ci mmol−1, American Radiolabeled Chem-
formation into Escherichia coli DH5α competent cells were icals) under standard assay conditions [10 μL (1−3 μg/μL
carried out per the manufacturer’s protocols. All mutations crude protein)] of the cleared lysate incubated with 1−60 μM
were verified by single-pass DNA sequencing (UK HealthCare [1-3H]FPP in a 50 μL total reaction volume for 15 min at 37
Genomics Laboratory). The template construct for all site- °C. Reactions were stopped with 50 μL of 2× stop buffer (0.4
directed mutagenesis reactions was pET-28a(+)-VDS M NaOH and 0.2 M EDTA), and the 3H-labeled hydrocarbon
(AGB05610) as described previously.17 Mutagenic primers products extracted with 200 μL of n-hexanes. After a silica
are listed in Table S1. scrub of the hexane extract to remove unreacted prenyl
In Vivo Product Analysis. Reaction products of mutant diphosphates and prenyl alcohols, aliquots were counted in a
proteins were profiled via heterologous expression in BL21- liquid scintillation counter. A crude lysate from bacteria
(DE3) cells. Plasmids harboring the correct coding sequence transformed with the empty pET28(a)+ vector (Novagen) was
for mutant enzymes were co-transformed into E. coli used for background subtraction. Michaelis−Menten kinetic
BL21(DE3) with pBbA5c-MevT(CO)-MBIS(CO,ispA) (ob- analysis was performed using GraphPad enzyme kinetic
tained from Addgene, plasmid 35151).36 Recombinant software version 9 (GraphPad Software, Inc., La Jolla, CA).
bacterial colonies were selected with kanamycin (25 μg/mL) In Vitro Product Analysis. For GC-MS profiling of wild
and chloramphenicol (17.5 μg/mL), and three independent type and mutant enzymes, aliquots of the cleared lysate (1−3
colonies from each plate were inoculated into 1 mL of 2×YT/ μg/μL total crude protein) were incubated with 230 μM cold
2% glycerol liquid medium with antibiotics and incubated at 37 FPP in 250 mM Tris-HCl (pH 7.0) and 50 mM MgCl2 (1 mL
°C for 16−20 h while being shaken. The next morning, cells total reaction volume), overlaid with 1 mL of n-hexanes at 37
from overnight seed culture were diluted into 10 mL of 2×YT/ °C in 4 mL glass vials. After 16 h, 600 μL aliquots of the
2% glycerol liquid medium at an OD600 of 0.05 supplemented organic overlay were carefully dried under a gentle nitrogen
with antibiotics and overlaid with 10% n-dodecane. Shake flask stream while being kept on ice. Concentrated samples were
fermentations were incubated at 37 °C for 2 h before the flasks resuspended in n-hexanes, and small volumes injected onto the
were cooled at 4 °C for 10 min. After addition of IPTG GC-MS instrument for profiling as described above.
(isopropyl β-D-1-thiogalactopyranoside) to a final concen- Immunodetection of VDS Enzymes. Expression levels of
tration of 0.4 mM, the flasks were incubated at 28 °C for an wild type and mutant enzymes were determined by separating
additional 20 h. The cultures were then centrifuged at 920g for aliquots of soluble cleared E. coli lysates by sodium dodecyl
10 min, and the dodecane overlay was removed for GC sulfate−polyacrylamide gel electrophoresis under reducing
analysis. conditions. Proteins were transferred to a nitrocellulose
Aliquots of the dodecane overlay were analyzed on an membrane, blocked with 5% nonfat dry milk, and probed
Agilent 7890 GC instrument [1:9 split injection; HP-5MS with a mouse monoclonal anti-polyhistidine antibody con-
column, 30 m × 0.25 mm, 0.25 μm film, 250 °C inlet jugated to alkaline phosphatase (Sigma-Aldrich) according to
temperature; oven temperature of 100 °C initially, increased to the manufacturer’s recommended protocol. Alkaline phospha-
140 °C (10 °C/min), then to 200 °C (5 °C/min), and finally tase-detected VDS immune complexes were estimated using a
to 300 °C (120 °C/min) and held for 5 min; He flow rate of purified wild type enzyme standard curve and ImageJ (Figure
0.9 mL/min] connected to a 5975C Agilent mass spectrometer S3).37


(run in positive ionization mode, 70 eV, scanning from 40 to
500 amu). Products were identified by comparison of mass
spectra to those of authentic standards, matches to the NIST
RESULTS AND DISCUSSION
2.0 library, and those reported by Yeo et al.17 (Figure S1). Building an Accurate Three-Dimensional (3D) Struc-
Caryophyllene (Sigma-Aldrich) was used as the standard for tural Representation of VDS. To target amino acid residues
quantification. within VDS that might govern product specificity, an energy-
In Vitro Mutant Expression for Enzyme Assays. minimized model of VDS was rendered upon the molecular
Plasmids harboring wild type or mutant genes encoding VDS coordinates of the EAS−FHP−Mg2+3 complex (PDB entry
were expressed in BL21(DE3) cells as described above. 5eat)29 using MODELLAR28 (Figure 1A). The VDS model
Overnight, saturated cultures were inoculated into 2×YT/2% exhibited the classical “terpene fold” consisting of α-helices
glycerol liquid medium with antibiotics at an OD600 of 0.05 connected by loops and turns adopting the αβ domain
and grown at 37 °C until an OD600 of 0.5−1.0 was reached. architecture (Figure 1A).1,29,38
3870 https://doi.org/10.1021/acs.biochem.1c00523
Biochemistry 2021, 60, 3868−3878
Biochemistry pubs.acs.org/biochemistry Article

spatial orientation of FPP, we aligned the EAS−FHP−Mg2+3


complex with each of the VDS−FPP complex models for visual
inspection (Figure S4). Of the nine theoretical binding
conformations, one binding conformation (6) geometrically
docked FPP in the characteristic “U” shape with the terminal
pyrophosphate group anchored between the trinuclear
magnesium cluster, angled toward the G1/G2 helix break.
With our substrate docked in a feasible orientation and an
accurate 3D molecular model of VDS, we set out to identify
the amino acid residues forming the active site topography.
Through a cavity detection radius constrained to 7 Å and a
cavity detection cutoff at three solvent radii, we identified 24
residues that formed the solvent inaccessible active site shell,
structurally poised to influence the catalytic trajectory of VDS
(Figure 1B).
Selection Criteria for Candidate Residues. Recognizing
that terpene product specificity is predominantly determined
by those residues lining the active site, we chose to restrict our
candidate selection to primarily first-tier residues identified by
our contact map strategy (Figure 1B).12,15 Several residues
classified as first-tier residues with well-documented catalytic
roles in sesquiterpene synthases were excluded from
consideration because of their essential contributions to the
Figure 1. (A) In silico 3D structural superimposition of the energy-
general catalytic process (Figure 1B): Arg277, conserved
minimized VDS tertiary model (NH2 terminus, orange rods; COOH
terminus, green rods; GenBank entry AGB0561017) and the template among all terpene synthases and necessary for offsetting the
crystal structure TEAS (NH2 terminal residues, yellow rods; COOH negative charge from the ionized diphosphate group of FPP;29
terminal residues, blue rods; PDB entry 5EAT).29 The red circle Trp286, responsible for the active site contour and stabilization
highlights VDS’ proposed catalytic pocket. (B) VDS’ active site cavity of carbocation intermediates through π−cation interactions;1
complexed with Mg2+3−FPP (Mg2+3 atoms shown as red spheres, FPP Asp314, part of the class I terpene synthase conserved metal-
carbon atoms colored gray, phosphorus atoms colored orange, and binding domain (DDXXD motif) that is necessary for Mg2+
oxygen atoms colored red; “DDXXD” and “NSE/DTE” metal-binding coordination and catalysis;7 Asp459, part of the class I terpene
domains are colored yellow and purple, respectively; J-K and A-C synthase second metal-binding domain triad (NSE/DTE
loops are colored orange and red, respectively). (C) Snapshot of motif) necessary for Mg2+ coordination and catalysis40 (Figure
proposed H+ donor residues within VDS interacting with 9 (blue
sticks) and 13 (red sticks) and other active site residues. Y535’s OH
1B).
group (J-K loop) is 3.8 Å (blue dotted line) and 4.3 Å (red dotted Instead, we focused on residues inferred to be associated
line) from C7 on 9 and 13, respectively. C452’s SH group (H2/3 with two other biochemical and structural criteria: (1) residues
helix) is 10.5 Å (blue dotted line) and 11.9 Å (red dotted line) from with the capability of initiating the second half-reaction of
C7 on 9 and 13, respectively. C415’s carbonyl group is approximately VDS’s catalytic cascade via donation of a proton across a
5.0 Å (blue dotted line) and 5.8 Å (red dotted line) from C7 on 9 and double bond and (2) residues surrounding the metal-binding
13, respectively. (D) Back side view of VDS displaying the influence domains that could restrict substrate-binding conformations.
on active site topology (gray meshwork) by select residues flanking Previously, our group postulated cysteine and tyrosine
the two metal-binding domains resulting in the 1,10 or 1,11 initial residues, structurally aligned to interact with the substrate,
ring closure of 1 (1,10 forms 3, blue; 1,11 forms 4, red). W286 (C
are capable of acting as catalytic acids through the donation of
helix), I307 (D helix), Y417 (G2 helix), and L420 (G2 helix) form the
bottom of the active site cavity. N455 is structurally aimed at C1 on 4, a proton from their hydroxyl or thiol side chains to reinitiate
while T310 and S311 (D helix) are poised at C12 and C13 on 3. the catalytic cascade at a stable reaction intermediate.41 This
was further supported by site-directed mutagenesis studies,
which demonstrated that a loss of the residue’s proton donor
The active site cavity (shape analogous to a cylinder) functional group halted the catalytic cascade at a neutral
measured approximately 14.1 Å tall (W286−D318) by 10.9 Å reaction intermediate limiting continuation of the cyclization
wide (S311−D539) (total volume of 1316 Å3) surrounded by cascade to the final hydrocarbon product.41,42 With this
predominantly nonpolar amino acid residues, akin to rationale, we sought to map the potential proton donor amino
previously crystallized sesquiterpene synthases (Figure acid residues within VDS. From our contact map, we identified
1B).29,38,39 The model’s Cα’s deviated by 2.154 Å with 92% residues Cys415 (G1/G2 helix break), Cys452 (H helix), and
of its atoms within 3.5 Å of their structural positions in EAS Tyr535 (J/K loop) as potential proton donor candidates that
(termed native overlap30). Finally, we identified the evolutio- warrant further investigation (Figure 1B).
narily conserved α-helices in the COOH terminus through Several research groups have found that while mutating
superimposition of the EAS crystal structure (Figure 1A). residues within the metal-binding domains of terpene
Using the default scoring function of Autodock Vina,34 we synthases often leads to loss of enzyme activity, targeting the
anchored FPP (MOPAC energy-minimized33) into the VDS residues just outside these domains diversifies the product
active site using the restricted search space with 1 Å spacing specificity.10,13,43−45 When we compared the locations of the
with a 15 Å grid box. We generated nine theoretical binding first-tier residues with respect to VDS’ two metal-binding
conformations of FPP, all with similar binding energies (−7.7 domains (D314 DTYD318 and D459 GE467), we selected Ile307,
to −7.9 kcal/mol). To visually determine the most reasonable Thr310, Ser311, and Asn455 as potential candidates that look
3871 https://doi.org/10.1021/acs.biochem.1c00523
Biochemistry 2021, 60, 3868−3878
Biochemistry pubs.acs.org/biochemistry Article

Table 1. Comparison of the In Vivo Reaction Products from WT and Mutant VDS Enzymesh

a
No detected sesquiterpene products; below the GC-MS limit of detection. bTrace amounts (tr.) represent <1% of the detectable reaction
products. cProduct identifications based upon comparisons to authentic standards. dProduct identifications based upon comparisons to mass
spectral data found in the NIST library 2.0. eProduct identifications based upon comparisons to those reported by Yeo et al.17 fThe mass spectrum
of the unknown sesquiterpene alcohol (Unk) can be found in Figure S1. gThe total sesquiterpene productivity (Total) was calculated by averaging
the individual sesquiterpene productivities (milligrams per liter per OD600 unit) from three independent bacterial cultures measured after aerobic
fermentation at 28 °C for 24 h and is reported for only select mutants. The gold and light blue shadings signify codon mutations to neutral and
proton donor amino acid residues, respectively. Colored product numbers correspond to their structural assignment and catalytic cascade cation
route (red for germacrenyl, blue for humulyl, and green for helmintogermacrenyl) in Scheme 1. hProduct distributions were calculated as
percentages of individual peak areas in relation to the total peak areas and quantified relative to a caryophyllene standard.

to restrict FPP’s binding conformation through hydrogen Briefly, expression plasmids with the mutant VDS genes were
bonding, van der Waals interactions, and substrate steric transformed into E. coli cells and incubated for 24 h with an n-
hindrance, thus poised to directly influence the catalytic dodecane overlay designed to “capture” the hydrophobic,
outcome. From our 3D homology model, we noticed these volatile reaction products. At 24 h, small aliquots (1 μL) of the
residues appear to exert the greatest regio- and stereochemical dodecane layer were injected onto a GC/MS instrument for
control over FPP-binding conformations, making them prime product detection, identification (Figure S1), and quantifica-
candidates for our site-directed mutagenesis study (Figure 1B). tion (Table 1). We evaluated each mutant’s ability to alter
Evaluation Procedure for Candidate Residues. We VDS’ product specificity by comparing the percent change in
utilized a well-documented in vivo terpene screening plat- the sesquiterpene product distribution compared with wild
form27,46 to detect and quantify the various sesquiterpene type VDS’ percent reaction product distribution, providing a
reaction products generated by the first-tier residue mutants. relative measure of the mutant’s activity compared with that of
3872 https://doi.org/10.1021/acs.biochem.1c00523
Biochemistry 2021, 60, 3868−3878
Biochemistry pubs.acs.org/biochemistry Article

Table 2. Steady-State Kinetic Parameters and In Vitro Product Distributions (percent) of WT and Mutant VDS Enzymesf

a
Trace amounts (tr.) represent <1% of the detectable reaction products. bProduct identifications are based upon comparisons to authentic
standards. cProduct identifications are based upon comparisons to mass spectral data found in the NIST library 2.0. dProduct identifications are
based upon comparisons to those reported by Yeo et al.17 eThe mass spectrum of the unknown sesquiterpene alcohol (Unk) can be found in Figure
S1. Error bars ± standard error of the mean. One-way analysis of variance using Dunnett’s multiple comparisons was performed to determine the
statistical differences between the observed WT’s KM, kcat, and kcat/KM and those of the mutants. *P < 0.05, **P < 0.005. fColored product numbers
correspond to their structural assignment and catalytic cascade cation route (red for germacrenyl, blue for humulyl, and green for
helmintogermacrenyl) in Scheme 1.

the wild type. Mutants that were active and demonstrated at [9 or 13 (Scheme 1)].42 From our in vivo assay, we detected
least a 10% shift from VLD (17) (66% of the wild type’s bicyclogermacrene (9), identified by its mass fragmentation
sesquiterpene products) were considered for further enzyme pattern compared to an authentic standard provided by
characterization. Results from all of the mutants evaluated in Bohlmann47 (Figure S2), as 100% of the reaction products
this work are shown in Table 1. For the mutants that met our (Table 1). This is a significant shift from wild type VDS’
criteria in the in vivo screen, we confirmed their reaction product distribution percentages that naturally produce 66%
products by an in vitro product profile assay, using the cleared 17, 5% 10, and 29% 9 (Table 1 and Figure 2). In vitro, Y535F’s
crude lysate from freshly transformed E. coli cells as our source substrate affinity constant, KM (9.35 μM), improved to half of
of enzyme. Results from the in vitro reactions are summarized that for wild type VDS (20.1 μM), while its turnover rate, kcat,
in Table 2. Finally, we characterized each mutant’s steady-state decreased ∼3-fold (0.005 s−1 vs 0.018 s−1), and its overall
kinetic parameters using a radiolabeled [1-3H]FPP assay (1− catalytic efficiency (kcat/KM) decreased to ∼60% of that of wild
60 μM) (Figure S5). We chose to use the crude cleared lysate type VDS (Table 2).
as our enzyme source for the kinetic assays due to the potential Could 9 be the neutral stable intermediate in VLD’s catalytic
loss of catalytic activity during the purification process.27 For cascade? According to Paknikar et al.,48 it is conceivable to
validation of our experimental accuracy, we compared the propose Tyr535 reprotonates C7 of 9, thus allowing the
kinetic parameters of the crude lysate of wild type VDS to concomitant formation of the aromadenrenyl-like cation and
those of the purified wild type VDS and observed no significant permitting the catalytic cascade to continue onto 17. However,
difference in their KM values (20.1 μM vs 29.8 μM clarified such a conclusion is tempered by the conservation of Tyr535
lysate and column purified) but an approximately 50% among other sesquiterpene synthases, including a bicycloger-
reduction in the purified enzyme’s turnover rate, kcat [0.018 macrene synthase (PdBCGS) in Figure 3.29,49−55 If Tyr535
vs 0.009 s−1 clarified lysate and column purified (Figure S6)]. might be the Brønsted acid in VDS’ active site that reinitiates
For k cat determinations, we calculated each mutant’s the catalytic cascade, then it might be a unique feature of VDS
concentration in picomoles within the crude lysate with and not serve such a role in other bicyclogermacrene synthases.
reference to a standard curve constructed with purified VDS Additionally, our previous efforts to elucidate putative VDS’s
detected by immunodetection against their N-terminal reaction intermediates using C13 labeling revealed the
histidine tag (Figure S3). Although the in vitro assays were biogenic origin of VLD’s unique isobutenyl tail likely arose
incubated for 16 h, shorter incubation periods gave similar from formation of a C1−C11 bond, which could proceed
product profiles, only with less total product yield (Figure S7). through a caryophyllenyl carbocation intermediate or the
Tyr535 Is a Likely Proton Donor in VDS. Following the germacrenyl cation intermediate as suggested by Paknikar et
same biochemical rationale as that described by Rising et al.,42 al.48 (Scheme 1 and Figure S8).17 The involvement of a
we mutated Tyr535 (J/K loop, 4.3 or 3.8 Å from C7 of 9 or caryophyllenyl intermediate in the biogenesis of sesquiterpenes
13, respectively) to a phenylalanine residue (Figure 1C). This containing this enigmatic isobutenyl side chain was first
afforded us the ability to maintain the spatial geometry proposed by Connolly et al. who studied unique sesquiterpene
previously provided by the tyrosine residue in the active site, all carbon structures isolated from liverworts.56 Conversely, 9’s
the while removing any potential contribution of proton biogenic carbon skeleton arises from the initial formation of
donation to the active site matrix. From our modeling and the C1−C10 bond from the all-trans-farnesyl carbocation (1).
previous terpene mechanistic studies, we hypothesized this This is subsequently followed by concomitant deprotonation at
mutation would halt the cyclization cascade at one of the C1 of the germacrenyl cation intermediate (3) leading to the
proposed stable intermediates within VLD’s catalytic cascade formation of the C1−C11 bond and yielding the bicyclo-
3873 https://doi.org/10.1021/acs.biochem.1c00523
Biochemistry 2021, 60, 3868−3878
Biochemistry pubs.acs.org/biochemistry Article

promiscuous enzyme that generates multiple reaction products,


a precedent reported for many other terpene synthases.15,57,62
Also, the results from the Y535F mutation suggest that Y535
might indeed be responsible for initiating donation of a proton
to the bicyclogermacrene intermediate, leading to the cyclo-
propylcarbinyl cation−cyclopropylcarbinyl cation rearrange-
ment2 and formation of 17 as the final reaction product. Also
arguing for the germacrenyl cation pathway for the biosyn-
thesis of 17 is the fact that no mutants yielding significant
amounts of α-humulene (5) or β-caryophyllene (11) were
uncovered. While such a negative result cannot be taken as
conclusive evidence that the humulyl cation pathway for the
biosynthesis of 17 does not exist in VDS, the Y535F mutant is
fully consistent with the importance of the germacrenyl cation
in the biosynthesis of valerena-1,10-diene (17).
Probing the Role of Cys415 and Cys452. From our
previous efforts to alter the product specificity of terpene
synthases,12 we hypothesized that if a TPS reaction cascade is
truncated at a stable intermediate, then its catalytic turnover
rate and efficiency should be at least comparable to (if not
greater than) those of the wild type enzyme such as we
observed with the C440A mutant from EAS.41,63 We combed
the active site residues of VDS from our constructed contact
map, searching for a cysteinyl sulfhydryl positioned to interact
with the β-isoprene unit of FPP and discovered Cys415,
located within the G1/G2 helix break region (Figure 1B). Baer
et al. identified and characterized the G1/G2 helix break
“effector triad”, a conserved structural motif comprised of three
amino acids among all bacteria, fungi, and plant class I terpene
Figure 2. Comparison of the metabolites of E. coli cells expressing synthases responsible for the induced fit between the synthase
either wild type VDS, mutant, or empty vector plasmids. Aliquots of a and substrate aiding in the transition from the open to the
dodecane overlay are compared by GC. Gray dashed lines indicate
products that accumulate in a variety of bacterial lines. Peak numbers closed conformation.64 In the VDS homology model, Cys415’s
correspond to the assigned reaction products in Scheme 1. Asterisks carbonyl group is structurally positioned in the “effector”
denote the unknown sesquiterpene alcohol that co-elutes with 9. position (3.7 Å from C3 on FPP) and could be responsible for
triggering the cleavage and release of the diphosphate of FPP
germacrene (9) intermediate (Scheme 1).57 According to the to form the delocalized allylic carbocation (1), while residues
mechanistic model of Paknikar et al.,48 17 can form from the Ser414 (linker) and Arg456 (pyrophosphate sensor) comprise
bicyclogermacrene intermediate via an energetically unfavor- the rest of the effector triad (Figure 1C). To explore the
able cyclopropylcarbinyl cation rearrangement (CCR), a catalytic role of Cys415, we substituted the thiol-containing
mechanism utilized by eukaryotes in the biosynthesis of residue for an alanine and serine residue and profiled the
triterpenoids that is useful in medicinal and industrial reaction products in vivo (Table 1). Both mutations exhibited a
applications.58−61 This model is also consistent with the 13C dramatic decrease in total productivity with an approximate
labeling studies of Yeo et al.17 Altogether, VDS appears to be a 50% diversion of the catalytic specificity from 17 to 9. Because

Figure 3. Alignment of amino acid residues mutated within VDS and representative class I sesquiterpene synthases. The “DDXXD” and “NSE/
DTE” metal-binding domains are highlighted in gray. The numbering corresponds to the amino acid positions within VDS. Alignments were
performed with the Multiple Sequence Comparison by Log-Expectation (MUSCLE) alignment package (version 3.8.31) within the MacVector
software program (version 18.1.5) and the C-terminal domains of AaADS [Artemisia annua amorpha-4,11-diene synthase (GenBank entry
AAF61439)], AaBFS [A. annua β-farnesene synthase (GenBank entry AAX39387)], AaBCS [A. annua β-caryophyllene synthase (GenBank entry
AAL79181)], GaDCS [Gossypium arboretum δ-cadinene synthase (GenBank entry AAA93065)], NtEAS [Nicotiana tabacum 5-epi-aristolochene
synthase (GenBank entry AAA19216)], ObGDS [Ocimum basilicum germacrene-D synthase (GenBank entry AAV63786)], PdBCGS [Phyla dulcis
bicyclogermacrene synthase (GenBank entry AFR23369)], ScGAS [Solidago canadensis germacrene-A synthase (GenBank entry CAC36896)],
VoVDS [Valeriana of f icinalis valerena-1,10-diene synthase (GenBank entry AGB05610)], and ZmACS [Zea mays α-copaene synthase (GenBank
entry AAX99148)].

3874 https://doi.org/10.1021/acs.biochem.1c00523
Biochemistry 2021, 60, 3868−3878
Biochemistry pubs.acs.org/biochemistry Article

neither C415A nor C415S eliminated the production of the synthase (Figure 3), no further mutations at this site were
primary reaction products found with the wild type enzyme, it investigated.
is unlikely Cys415 is directly involved in the active site acid− Thr310 and Ser311 are situated above the first aspartate
base chemistry. residue of the DDXXD motif oriented toward the active site
From the homology model, Cys283 (C helix) and Cys452 pocket and juxtaposed relative to the γ-isoprene unit of FPP,
(H helix) are positioned 11.9 and 10.5 Å from C7 on 9 and 13, potentially poised to influence the orientation of the initial
respectively (Figure 1C). However, we narrowed our focus to farnesyl carbocation and control of the ensuing catalytic
Cys452 instead of Cys283, because it is positioned like Cys270 cascade (Figure 1B). Consistent with the essential nature of
in EAS (PDB entry 5EAT) and appears to be involved in the these particular residues (Figure 3), all attempts to mutate
formation of an active site hydrogen bond network with Ser311 with hydrophobic (alanine and isoleucine), polar
Tyr541 and the creation of an essential structural feature in (threonine), or aromatic (phenylalanine) mutations lead to
VDS.29 We removed the proton donor capabilities of Cys452 inactive VDS mutants (Table 1). However, substitutions of
with an alanine mutation and characterized the reaction Thr310 with alanine, isoleucine, or methionine yielded
products both in vivo and in vitro. The in vivo analysis detected catalytically active enzymes with alterations in the proportion
a product distribution for C452A of 31% valerenadiene (17), of valerenadiene (17), bicyclogermacrene (9), and allo-
65% bicyclogermacrene (9), and 4% allo-aromadendrene (10), aromadendrene (10) (Tables 1 and 2), except that the
similar to the product profile shift of the C415A and C415S methionine mutant (T310M) also generated a new product, α-
mutants (Table 1). However, the KM for the C452A mutant gurjunene (12), in 27% abundance. While this particular
(20.7 μM) is identical to that of VDS (20.1 μM), but its kcat of mutant maintained a KM (22.8 μM) for FPP comparable to
0.038 s−1 is double that for VDS (0.018 s−1), which translates that of the wild type enzyme, its catalytic turnover rate (kcat)
to a doubling of the catalytic efficiency (kcat/KM) of C452A was only ∼1/10th of that of wild type VDS (Table 2).
(Table 2). The T310 mutations are consistent with the germacrenyl
The failure of C452A to completely shift the product cation pathway being the dominant pathway for the biosyn-
specificity of VDS toward 9 is not consistent with its role as the thesis of VLD (17). Other than the T310S mutant, which was
Lewis acid in donating a proton to intermediate 13 in the VDS largely inactive, mutations to nonpolar residues like alanine,
catalytic cascade (Scheme 1). An alternative is that a water isoleucine, or methionine shifted the product profile to greater
molecule may be trapped in the active site serving as the proportions of bicylcogermacrene (9), consistent with T310
requisite proton donor, a hypothesis that cannot be ruled out being a gateway residue for the conversion of bicyclogerma-
for VDS.1,5,6,44,45,57 Perhaps Cys452 is contributing more to crene (9) to VLD (17). The extra bulk contributed by the
active site topology guiding catalytic outcomes. The analogous T310M mutant would also be consistent with this mutant
position within the previously characterized bicyclogermacrene allowing the bicyclogermacrene (9) intermediate to initially
synthase49 (Figure 3) is occupied by a leucine residue, located proceed down the catalytic cascade but to be diverted to a
just outside the active site on helix H (Figure 1B,D), which novel product, α-gurjunene (12), because the active site was
may have sufficient influence on the active site contour to steer constrained from performing the cyclopropylcarbinyl cation−
the catalytic activity more specifically toward 9. A logical cyclopropylcarbinyl cation rearrangement. Schrepfer et al. used
inference from this interpretation would be to evaluate a a similar logic in mapping functionally plastic residues within
C452L mutation in VDS for its product specificity and catalytic taxadiene synthase, a type I diterpene synthase.14
efficiency. Asn455 Suppresses a Dormant Cyclization Cascade
Contributions of Residues Proximal to a Metal- in VDS. On the basis of the report of Salmon et al.66 that a
Binding Domain. Several research groups have demonstrated valine to glycine mutation within a β-farnesene synthase was
that residues surrounding the metal-binding domains found sufficient to convert product specificity from linear to
within amorpha-4,11-diene synthase (ADS) can contribute to completely cyclized, we examined the possible role of the
the initial and subsequent ring closures of ionized FPP in the analogous position within VDS, 455 (Figure 3), to contribute
formation of the amorphenyl cation, directly impacting to possible intermediate formation in the VLD cascade. When
product specificity and catalytic activity.10,43,65 Hence, we Asn455 was substituted with alanine, cysteine, serine, or
investigated Ile307, Thr310, and Ser311, the analogous threonine, the resulting mutant enzymes possessed only 12−
residues proximal to the DDXXD metal-binding domain in 25% of the wild type enzyme activity and produced >40% of
VDS (Figure 3). germacradien-4-ol (7) as the majority product. This is in
To begin to evaluate the contribution of the hydrophobic contrast to VDS, or the other mutants mentioned above that
surface comprised of these residues to the catalytic activity, yield only a small percentage of 7. However, more surprising
Ile307 associated with the D helix was substituted with a was the fact that much of the remaining products resulting
leucine. This position resides within the hydrophobic core from these mutants consisted of α-copaene (14), δ-cadinene
deep within the VDS’ active site along with Trp286, Tyr417, (15), and cubebol (16), products that can arise only from an
and Leu420 (Figure 1D). Interestingly, an I307L mutation, a all-trans to cisoid conformation change of FPP to nerolidyl
mutation of a conserved residue, resulted in an almost 60% diphosphate, NPP (Scheme 1, green arrow route). Moreover,
decline in relative activity without any change in overall these mutants also lost their ability to produce 9 and 17. While
product specificity other than a shift in the ratio of products, a none of the Asn455 mutants accumulated products that could
>35% increase in the amount of bicyclogermacrene accumulat- be interpreted as intermediates for either of the possible
ing, and the concomitant decrease in valerenadiene (Table 1). catalytic pathways to 17, these mutants appear to be
Due to the buried nature of this residue within the active site, confirming an important evolutionary role for this position in
the loss of activity with a highly conserved mutation, and the the cyclization pathways of type 1 sesquiterpene synthases as
conserved nature of the Ile at this position among terpene suggested by Salmon et al.13
3875 https://doi.org/10.1021/acs.biochem.1c00523
Biochemistry 2021, 60, 3868−3878
Biochemistry

■ ■
pubs.acs.org/biochemistry Article

CONCLUSIONS ACKNOWLEDGMENTS
While the experimental aim of this work was to find mutants of The authors thank Eric Nybo, Stephen Bell, Chase Kempinski,
VDS that produced cyclic derivatives with alternative Kristin Linscott, and Scott Kinison for their helpful discussions
configurations of the isobutenyl side chain or more structurally over the course of this work.


constrained analogues of the valerenadiene scaffold, the work
has established a map of residues impacting catalytic outcomes, ABBREVIATIONS
including a mutant producing exclusively bicyclogermacrene
VDS, valerenadiene synthase; BCG, bicyclogermacrene; FPP,
(Y535F), and identified Asn455 as a gatekeeper for
farnesyl diphosphate; VLD, valerena-1,10-diene; VA, valerenic
sesquiterpene products requiring a configurational isomer-
acid; TPS, terpene synthase.


ization of FPP to the cisoid precursor NPP. These results
should help focus future efforts in creating more constrained
scaffolds for building analogues to test for their modulatory REFERENCES
activity of the GABAA receptor and anxiolytic activities in (1) Christianson, D. W. Structural and Chemical Biology of
animal models. Terpenoid Cyclases. Chem. Rev. 2017, 117 (17), 11570−11648.


(2) Poulter, C. D.; Rilling, H. C. The prenyl transfer reaction.
Enzymic and mechanistic studies of the 1′-4 coupling reaction in the
ASSOCIATED CONTENT
terpene biosynthetic pathway. Acc. Chem. Res. 1978, 11 (8), 307−313.
*
sı Supporting Information (3) Ruzicka, L. The isoprene rule and the biogenesis of terpenic
The Supporting Information is available free of charge at compounds. Experientia 1953, 9 (10), 357−67.
https://pubs.acs.org/doi/10.1021/acs.biochem.1c00523. (4) Tantillo, D. J. Biosynthesis via carbocations: theoretical studies
on terpene formation. Nat. Prod. Rep. 2011, 28 (6), 1035−53.
VDS mutagenesis primers (Table S1), mass spectra of (5) Grundy, D. J.; Chen, M.; Gonzalez, V.; Leoni, S.; Miller, D. J.;
sesquiterpenes (Figure S1), structural conformation of Christianson, D. W.; Allemann, R. K. Mechanism of Germacradien-4-
bicyclogermacrene from E. coli (Figure S2), immunode- ol Synthase-Controlled Water Capture. Biochemistry 2016, 55 (14),
tection of WT VDS and mutant enzymes (Figure S3), 2112−21.
VDS-TEAS homology model (Figure S4), WT VDS and (6) Felicetti, B.; Cane, D. E. Aristolochene synthase: mechanistic
mutant enzyme kinetic plots (Figure S5), WT VDS analysis of active site residues by site-directed mutagenesis. J. Am.
column purified versus from clarified bacterial lysate Chem. Soc. 2004, 126 (23), 7212−21.
(7) Kersten, R. D.; Diedrich, J. K.; Yates, J. R., 3rd; Noel, J. P.
(Figure S6), GC chromatogram of hydrocarbon
Mechanism-Based Post-Translational Modification and Inactivation in
products after a 2 h assay (Figure S7), 13C labeling Terpene Synthases. ACS Chem. Biol. 2015, 10 (11), 2501−11.
pattern for Scheme 1 (Figure S8), and additional (8) Singh, B.; Sharma, R. A. Plant terpenes: defense responses,
references (PDF) phylogenetic analysis, regulation and clinical applications. 3 Biotech
2015, 5 (2), 129−151.
Accession Codes (9) Paddon, C. J.; Westfall, P. J.; Pitera, D. J.; Benjamin, K.; Fisher,
The NCBI Accession ID for the valerenadiene synthase K.; McPhee, D.; Leavell, M. D.; Tai, A.; Main, A.; Eng, D.; Polichuk,
described here is AGB05610. D. R.; Teoh, K. H.; Reed, D. W.; Treynor, T.; Lenihan, J.; Fleck, M.;

■ AUTHOR INFORMATION
Corresponding Author
Bajad, S.; Dang, G.; Dengrove, D.; Diola, D.; Dorin, G.; Ellens, K. W.;
Fickes, S.; Galazzo, J.; Gaucher, S. P.; Geistlinger, T.; Henry, R.;
Hepp, M.; Horning, T.; Iqbal, T.; Jiang, H.; Kizer, L.; Lieu, B.; Melis,
D.; Moss, N.; Regentin, R.; Secrest, S.; Tsuruta, H.; Vazquez, R.;
Joe Chappell − Pharmaceutical Sciences, College of Pharmacy, Westblade, L. F.; Xu, L.; Yu, M.; Zhang, Y.; Zhao, L.; Lievense, J.;
University of Kentucky, Lexington, Kentucky 40536-0596, Covello, P. S.; Keasling, J. D.; Reiling, K. K.; Renninger, N. S.;
United States; orcid.org/0000-0001-8469-8976; Newman, J. D. High-level semi-synthetic production of the potent
Phone: (859) 218-0775; Email: chappell@uky.edu antimalarial artemisinin. Nature 2013, 496 (7446), 528−32.
(10) Abdallah, I. I.; van Merkerk, R.; Klumpenaar, E.; Quax, W. J.
Author Catalysis of amorpha-4,11-diene synthase unraveled and improved by
Garrett E. Zinck − Pharmaceutical Sciences, College of mutability landscape guided engineering. Sci. Rep. 2018, 8 (1), 9961.
Pharmacy, University of Kentucky, Lexington, Kentucky (11) Dixit, M.; Weitman, M.; Gao, J.; Major, D. T. Chemical
40536-0596, United States; orcid.org/0000-0001-9738- Control in the Battle against Fidelity in Promiscuous Natural Product
6576 Biosynthesis: The Case of Trichodiene Synthase. ACS Catal. 2017, 7
(1), 812−818.
Complete contact information is available at: (12) Greenhagen, B. T.; O’Maille, P. E.; Noel, J. P.; Chappell, J.
https://pubs.acs.org/10.1021/acs.biochem.1c00523 Identifying and manipulating structural determinates linking catalytic
specificities in terpene synthases. Proc. Natl. Acad. Sci. U. S. A. 2006,
Author Contributions 103 (26), 9826−31.
J.C. and G.E.Z. provided the conceptual context for the work. (13) Salmon, M.; Laurendon, C.; Vardakou, M.; Cheema, J.;
Defernez, M.; Green, S.; Faraldos, J. A.; O’Maille, P. E. Emergence of
J.C. and G.E.Z. designed the experiments and experimental
terpene cyclization in Artemisia annua. Nat. Commun. 2015, 6, 6143.
approaches. G.E.Z. performed all of the experiments. G.E.Z. (14) Schrepfer, P.; Buettner, A.; Goerner, C.; Hertel, M.; van Rijn, J.;
and J.C. wrote the manuscript. Wallrapp, F.; Eisenreich, W.; Sieber, V.; Kourist, R.; Bruck, T.
Funding Identification of amino acid networks governing catalysis in the closed
This work was supported by the Digenis Endowment (to J.C.) complex of class I terpene synthases. Proc. Natl. Acad. Sci. U. S. A.
and the University of Kentucky College of Pharmacy (to 2016, 113 (8), E958−67.
G.E.Z.). (15) Yoshikuni, Y.; Ferrin, T. E.; Keasling, J. D. Designed divergent
evolution of enzyme function. Nature 2006, 440 (7087), 1078−82.
Notes (16) Pyle, B. W.; Tran, H. T.; Pickel, B.; Haslam, T. M.; Gao, Z.;
The authors declare no competing financial interest. MacNevin, G.; Vederas, J. C.; Kim, S. U.; Ro, D. K. Enzymatic

3876 https://doi.org/10.1021/acs.biochem.1c00523
Biochemistry 2021, 60, 3868−3878
Biochemistry pubs.acs.org/biochemistry Article

synthesis of valerena-4,7(11)-diene by a unique sesquiterpene (34) Trott, O.; Olson, A. J. AutoDock Vina: improving the speed
synthase from the valerian plant (Valeriana officinalis). FEBS J. and accuracy of docking with a new scoring function, efficient
2012, 279 (17), 3136−46. optimization, and multithreading. J. Comput. Chem. 2010, 31 (2),
(17) Yeo, Y. S.; Nybo, S. E.; Chittiboyina, A. G.; Weerasooriya, A. 455−61.
D.; Wang, Y. H.; Gongora-Castillo, E.; Vaillancourt, B.; Buell, C. R.; (35) The PyMOL Molecular Graphics System, ver. 1.8; Schrodinger,
DellaPenna, D.; Celiz, M. D.; Jones, A. D.; Wurtele, E. S.; Ransom, LLC, 2015.
N.; Dudareva, N.; Shaaban, K. A.; Tibrewal, N.; Chandra, S.; Smillie, (36) Peralta-Yahya, P. P.; Ouellet, M.; Chan, R.; Mukhopadhyay, A.;
T.; Khan, I. A.; Coates, R. M.; Watt, D. S.; Chappell, J. Functional Keasling, J. D.; Lee, T. S. Identification and microbial production of a
identification of valerena-1,10-diene synthase, a terpene synthase terpene-based advanced biofuel. Nat. Commun. 2011, 2, 483.
catalyzing a unique chemical cascade in the biosynthesis of (37) Schneider, C. A.; Rasband, W. S.; Eliceiri, K. W. NIH Image to
biologically active sesquiterpenes in Valeriana officinalis. J. Biol. ImageJ: 25 years of image analysis. Nat. Methods 2012, 9 (7), 671−5.
Chem. 2013, 288 (5), 3163−73. (38) Lesburg, C. A.; Zhai, G.; Cane, D. E.; Christianson, D. W.
(18) Bos, R.; Woerdenbag, H. J.; Hendriks, H.; Zwaving, J. H.; De Crystal structure of pentalenene synthase: mechanistic insights on
Smet, P. A. G. M.; Tittel, G.; Wikström, H. V.; Scheffer, J. J. C. terpenoid cyclization reactions in biology. Science 1997, 277 (5333),
Analytical Aspects of Phytotherapeutic Valerian Preparations. 1820−4.
Phytochem. Anal. 1996, 7 (3), 143−151. (39) Tarshis, L. C.; Yan, M.; Poulter, C. D.; Sacchettini, J. C. Crystal
(19) Ricigliano, V.; Kumar, S.; Kinison, S.; Brooks, C.; Nybo, S. E.; structure of recombinant farnesyl diphosphate synthase at 2.6-A
Chappell, J.; Howarth, D. G. Regulation of sesquiterpenoid resolution. Biochemistry 1994, 33 (36), 10871−7.
metabolism in recombinant and elicited Valeriana officinalis hairy (40) Aaron, J. A.; Christianson, D. W. Trinuclear Metal Clusters in
roots. Phytochemistry 2016, 125, 43−53. Catalysis by Terpenoid Synthases. Pure Appl. Chem. 2010, 82 (8),
(20) Benke, D.; Barberis, A.; Kopp, S.; Altmann, K. H.; Schubiger, 1585−1597.
M.; Vogt, K. E.; Rudolph, U.; Mohler, H. GABA A receptors as in vivo (41) Greenhagen, B. T. Origins of Isoprenoid Diversity: A Study of
substrate for the anxiolytic action of valerenic acid, a major Structure-Function Relationships in Sequiterpene Synthases. Uni-
constituent of valerian root extracts. Neuropharmacology 2009, 56 versity of Kentucky, Lexington, KY, 2003.
(1), 174−81. (42) Rising, K. A.; Starks, C. M.; Noel, J. P.; Chappell, J.
(21) Khom, S.; Baburin, I.; Timin, E.; Hohaus, A.; Trauner, G.; Demonstration of Germacrene A as an Intermediate in 5-Epi-
aristolochene Synthase Catalysis. J. Am. Chem. Soc. 2000, 122 (9),
Kopp, B.; Hering, S. Valerenic acid potentiates and inhibits GABA(A)
1861−1866.
receptors: molecular mechanism and subunit specificity. Neuro-
(43) Fang, X.; Li, J. X.; Huang, J. Q.; Xiao, Y. L.; Zhang, P.; Chen, X.
pharmacology 2007, 53 (1), 178−87.
Y. Systematic identification of functional residues of Artemisia annua
(22) Khom, S.; Strommer, B.; Ramharter, J.; Schwarz, T.; Schwarzer,
amorpha-4,11-diene synthase. Biochem. J. 2017, 474 (13), 2191−
C.; Erker, T.; Ecker, G. F.; Mulzer, J.; Hering, S. Valerenic acid
2202.
derivatives as novel subunit-selective GABAA receptor ligands - in
(44) Rynkiewicz, M. J.; Cane, D. E.; Christianson, D. W. X-ray
vitro and in vivo characterization. Br. J. Pharmacol. 2010, 161 (1), crystal structures of D100E trichodiene synthase and its pyrophos-
65−78. phate complex reveal the basis for terpene product diversity.
(23) Luger, D.; Poli, G.; Wieder, M.; Stadler, M.; Ke, S.; Ernst, M.; Biochemistry 2002, 41 (6), 1732−41.
Hohaus, A.; Linder, T.; Seidel, T.; Langer, T.; Khom, S.; Hering, S. (45) Yoshikuni, Y.; Martin, V. J.; Ferrin, T. E.; Keasling, J. D.
Identification of the putative binding pocket of valerenic acid on Engineering cotton (+)-delta-cadinene synthase to an altered
GABAA receptors using docking studies and site-directed muta- function: germacrene D-4-ol synthase. Chem. Biol. 2006, 13 (1),
genesis. Br. J. Pharmacol. 2015, 172 (22), 5403−13. 91−8.
(24) Khom, S.; Hintersteiner, J.; Luger, D.; Haider, M.; Pototschnig, (46) Nybo, S. E.; Saunders, J.; McCormick, S. P. Metabolic
G.; Mihovilovic, M. D.; Schwarzer, C.; Hering, S. Analysis of beta- engineering of Escherichia coli for production of valerenadiene. J.
Subunit-dependent GABAA Receptor Modulation and Behavioral Biotechnol. 2017, 262, 60−66.
Effects of Valerenic Acid Derivatives. J. Pharmacol. Exp. Ther. 2016, (47) Jakupovic, J.; Pathak, V. P.; Grenz, M.; Banerjee, S.; Wolfrum,
357 (3), 580−90. C.; Baruah, R. N.; Bohlmann, F. Sesquiterpene derivatives from
(25) Allen, S. E.; Dokholyan, N. V.; Bowers, A. A. Dynamic Docking othonna and related species. Phytochemistry 1987, 26 (4), 1049−
of Conformationally Constrained Macrocycles: Methods and 1052.
Applications. ACS Chem. Biol. 2016, 11 (1), 10−24. (48) Paknikar, S. K.; Kadam, S. H.; Ehrlich, A. L.; Bates, R. B.
(26) Pan, A. C.; Borhani, D. W.; Dror, R. O.; Shaw, D. E. Molecular Alternate biosynthesis of valerenadiene and related sesquiterpenes.
determinants of drug-receptor binding kinetics. Drug Discovery Today Nat. Prod. Commun. 2013, 8 (9), 1195−6.
2013, 18 (13−14), 667−73. (49) Attia, M.; Kim, S. U.; Ro, D. K. Molecular cloning and
(27) Bell, S. A.; Niehaus, T. D.; Nybo, S. E.; Chappell, J. Structure- characterization of (+)-epi-alpha-bisabolol synthase, catalyzing the
function mapping of key determinants for hydrocarbon biosynthesis first step in the biosynthesis of the natural sweetener, hernandulcin, in
by squalene and squalene synthase-like enzymes from the green alga Lippia dulcis. Arch. Biochem. Biophys. 2012, 527 (1), 37−44.
Botryococcus braunii race B. Biochemistry 2014, 53 (48), 7570−81. (50) Iijima, Y.; Davidovich-Rikanati, R.; Fridman, E.; Gang, D. R.;
(28) Sali, A.; Blundell, T. L. Comparative protein modelling by Bar, E.; Lewinsohn, E.; Pichersky, E. The biochemical and molecular
satisfaction of spatial restraints. J. Mol. Biol. 1993, 234 (3), 779−815. basis for the divergent patterns in the biosynthesis of terpenes and
(29) Starks, C. M.; Back, K.; Chappell, J.; Noel, J. P. Structural basis phenylpropenes in the peltate glands of three cultivars of basil. Plant
for cyclic terpene biosynthesis by tobacco 5-epi-aristolochene Physiol. 2004, 136 (3), 3724−36.
synthase. Science 1997, 277 (5333), 1815−20. (51) Cai, Y.; Jia, J. W.; Crock, J.; Lin, Z. X.; Chen, X. Y.; Croteau, R.
(30) Eramian, D.; Eswar, N.; Shen, M. Y.; Sali, A. How well can the A cDNA clone for beta-caryophyllene synthase from Artemisia annua.
accuracy of comparative protein structure models be predicted? Phytochemistry 2002, 61 (5), 523−9.
Protein Sci. 2008, 17 (11), 1881−93. (52) Bennett, M. H.; Mansfield, J. W.; Lewis, M. J.; Beale, M. H.
(31) Melo, F.; Sanchez, R.; Sali, A. Statistical potentials for fold Cloning and expression of sesquiterpene synthase genes from lettuce
assessment. Protein Sci. 2002, 11 (2), 430−48. (Lactuca sativa L.). Phytochemistry 2002, 60 (3), 255−61.
(32) Shen, M. Y.; Sali, A. Statistical potential for assessment and (53) Chen, X. Y.; Wang, M.; Chen, Y.; Davisson, V. J.; Heinstein, P.
prediction of protein structures. Protein Sci. 2006, 15 (11), 2507−24. Cloning and heterologous expression of a second (+)-delta-cadinene
(33) Stewart, J. J. P. MOPAC2016; Stewart Computational synthase from Gossypium arboreum. J. Nat. Prod. 1996, 59 (10),
Chemistry: Colorado Springs, CO, 2016. 944−51.

3877 https://doi.org/10.1021/acs.biochem.1c00523
Biochemistry 2021, 60, 3868−3878
Biochemistry pubs.acs.org/biochemistry Article

(54) van Der Hoeven, R. S.; Monforte, A. J.; Breeden, D.; Tanksley,
S. D.; Steffens, J. C. Genetic control and evolution of sesquiterpene
biosynthesis in Lycopersicon esculentum and L. hirsutum. Plant Cell
2000, 12 (11), 2283−94.
(55) Bouwmeester, H. J.; Kodde, J.; Verstappen, F. W.; Altug, I. G.;
de Kraker, J. W.; Wallaart, T. E. Isolation and characterization of two
germacrene A synthase cDNA clones from chicory. Plant Physiol.
2002, 129 (1), 134−44.
(56) Connolly, J. D.; Harrison, L. J.; Rycroft, D. S. The structure of
tamariscol, a new pacifigorgiane sesquiterpenoid alcohol from the
liverwort frullania tamarisci. Tetrahedron Lett. 1984, 25 (13), 1401−
1402.
(57) Garms, S.; Kollner, T. G.; Boland, W. A multiproduct terpene
synthase from Medicago truncatula generates cadalane sesquiterpenes
via two different mechanisms. J. Org. Chem. 2010, 75 (16), 5590−600.
(58) Blagg, B. S.; Jarstfer, M. B.; Rogers, D. H.; Poulter, C. D.
Recombinant squalene synthase. A mechanism for the rearrangement
of presqualene diphosphate to squalene. J. Am. Chem. Soc. 2002, 124
(30), 8846−53.
(59) Niehaus, T. D.; Okada, S.; Devarenne, T. P.; Watt, D. S.;
Sviripa, V.; Chappell, J. Identification of unique mechanisms for
triterpene biosynthesis in Botryococcus braunii. Proc. Natl. Acad. Sci.
U. S. A. 2011, 108 (30), 12260−5.
(60) Kempinski, C.; Chappell, J. Engineering triterpene metabolism
in the oilseed of Arabidopsis thaliana. Plant Biotechnol. J. 2019, 17 (2),
386−396.
(61) Kempinski, C.; Jiang, Z.; Zinck, G.; Sato, S. J.; Ge, Z.;
Clemente, T. E.; Chappell, J. Engineering linear, branched-chain
triterpene metabolism in monocots. Plant Biotechnol. J. 2019, 17 (2),
373−385.
(62) Vattekkatte, A.; Garms, S.; Boland, W. Alternate Cyclization
Cascade Initiated by Substrate Isomer in Multiproduct Terpene
Synthase from Medicago truncatula. J. Org. Chem. 2017, 82 (6),
2855−2861.
(63) Greenhagen, B. T.; Griggs, P.; Takahashi, S.; Ralston, L.;
Chappell, J. Probing sesquiterpene hydroxylase activities in a coupled
assay with terpene synthases. Arch. Biochem. Biophys. 2003, 409 (2),
385−94.
(64) Baer, P.; Rabe, P.; Fischer, K.; Citron, C. A.; Klapschinski, T.
A.; Groll, M.; Dickschat, J. S. Induced-fit mechanism in class I terpene
cyclases. Angew. Chem., Int. Ed. 2014, 53 (29), 7652−6.
(65) Li, Z.; Gao, R.; Hao, Q.; Zhao, H.; Cheng, L.; He, F.; Liu, L.;
Liu, X.; Chou, W. K.; Zhu, H.; Cane, D. E. The T296V Mutant of
Amorpha-4,11-diene Synthase Is Defective in Allylic Diphosphate
Isomerization but Retains the Ability To Cyclize the Intermediate
(3R)-Nerolidyl Diphosphate to Amorpha-4,11-diene. Biochemistry
2016, 55 (48), 6599−6604.
(66) Nguyen, T. D.; Faraldos, J. A.; Vardakou, M.; Salmon, M.;
O’Maille, P. E.; Ro, D. K. Discovery of germacrene A synthases in
Barnadesia spinosa: The first committed step in sesquiterpene lactone
biosynthesis in the basal member of the Asteraceae. Biochem. Biophys.
Res. Commun. 2016, 479 (4), 622−627.
(67) Garms, S.; Köllner, T. G.; Boland, W. A Multiproduct Terpene
Synthase from Medicago truncatula Generates Cadalane Sesquiter-
penes via Two Different Mechanisms. J. Org. Chem. 2010, 75 (16),
5590−5600.
(68) Schmidt, C. O.; Bouwmeester, H. J.; Bulow, N.; Konig, W. A.
Isolation, characterization, and mechanistic studies of (−)-alpha-
gurjunene synthase from Solidago canadensis. Arch. Biochem. Biophys.
1999, 364 (2), 167−77.
(69) Picaud, S.; Olsson, M. E.; Brodelius, M.; Brodelius, P. E.
Cloning, expression, purification and characterization of recombinant
(+)-germacrene D synthase from Zingiber officinale. Arch. Biochem.
Biophys. 2006, 452 (1), 17−28.

■ NOTE ADDED AFTER ASAP PUBLICATION


Table 2 was corrected shortly after asap publication on
December 13, 2021.

3878 https://doi.org/10.1021/acs.biochem.1c00523
Biochemistry 2021, 60, 3868−3878

You might also like