Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Single-step fermentative production of the cholesterol-

lowering drug pravastatin via reprogramming of


Penicillium chrysogenum
Kirsty J. McLeana,1, Marcus Hansb,1, Ben Meijrinkb, Wibo B. van Scheppingenb, Aad Vollebregtb, Kang Lan Teea,
Jan-Metske van der Laanb, David Leysa, Andrew W. Munroa,2, and Marco A. van den Bergb,2
a
Manchester Institute of Biotechnology, Faculty of Life Sciences, The University of Manchester, Manchester M1 7DN, United Kingdom; and bDSM
Biotechnology Center, 2613 AX, Delft, The Netherlands

Edited by Arnold L. Demain, Drew University, Madison, NJ, and approved January 28, 2015 (received for review October 3, 2014)

The cholesterol-lowering blockbuster drug pravastatin can be process. First, fermentation of P. citrinum produces compactin.
produced by stereoselective hydroxylation of the natural product Next, the statins are purified and the lactone ring is opened by
compactin. We report here the metabolic reprogramming of the addition of sodium hydroxide. After neutralization, the open
antibiotics producer Penicillium chrysogenum toward an industrial lactone form of compactin is converted to pravastatin in a bio-
pravastatin production process. Following the successful introduction transformation step using the bacterium Streptomyces carbophilus.
of the compactin pathway into the β-lactam–negative P. chrysoge- This organism contains a cytochrome P450 enzyme (P450 or
num DS50662, a new cytochrome P450 (P450 or CYP) from Amy- CYP) that is capable of hydroxylating compactin stereospecifi-
colatopsis orientalis (CYP105AS1) was isolated to catalyze the final cally at C-6 with an overall yield of 70% (P450sca-2) (9). CYPs
compactin hydroxylation step. Structural and biochemical charac- are hemoproteins of biotechnological interest due to their ability
terization of the WT CYP105AS1 reveals that this CYP is an effi- to perform regio- and stereoselective oxygen insertion chemistry.
cient compactin hydroxylase, but that predominant compactin
They are best known for their pivotal roles in human steroid
binding modes lead mainly to the ineffective epimer 6-epi-prava-
synthesis and in drug/xenobiotic detoxification and metabolism
statin. To avoid costly fractionation of the epimer, the enzyme
(10). In addition, nonhuman CYPs have increasing importance
was evolved to invert stereoselectivity, producing the pharma-
in production of antibiotics and other agents beneficial to
cologically active pravastatin form. Crystal structures of the opti-
mized mutant P450Prava bound to compactin demonstrate how the
human health (11). Among the statins, pravastatin has the
selected combination of mutations enhance compactin binding lowest potential for drug interactions (e.g., with antibiotics) as
and enable positioning of the substrate for stereo-specific oxida- it is not extensively metabolized by human CYPs (12). However,
tion. Expression of P450Prava fused to a redox partner in compac- low P450sca-2 activity during biotransformation, compactin tox-
tin-producing P. chrysogenum yielded more than 6 g/L pravastatin icity, and low pravastatin yield hamper large-scale manufacturing
at a pilot production scale, providing an effective new route to
industrial scale production of an important drug. Significance

pravastatin | P450 engineering | fermentation | statin | cholesterol Statins are successful widely used drugs that decrease the risk
of coronary heart disease and strokes by lowering cholesterol

S tatin drugs inhibit 3β-hydroxy-3-methylglutaryl CoA (HMG-


CoA) reductase, the enzyme catalyzing the rate-limiting step
in cholesterol biosynthesis. Statins reduce “bad” plasma (LDL)
levels. They selectively inhibit the key regulatory enzyme of
the cholesterol synthesis pathway, thus lowering levels of
plasma LDL (bad) cholesterol. Pravastatin is one of the leading
cholesterol levels and are thus effective against hypercholestero- and most effective statins, derived from the natural product
lemia. Several statins are on the market; the most prominent compactin. However, pravastatin production involves a costly
being the completely synthetic atorvastatin (Lipitor), the semi- dual-step fermentation and biotransformation process. Here
synthetic simvastatin (Zocor), and pravastatin (Pravachol) (Fig. 1) we present a single-step fermentative method for production

MICROBIOLOGY
(1, 2). Atorvastatin was identified by a Pfizer R&D pharma- of the active drug pravastatin. Reprogramming of the antibiotics-
ceutical screening program using entirely unnatural lead struc- producing fungus Penicillium chrysogenum, with discovery and
tures (3); simvastatin and pravastatin, discovered by Merck and engineering of an enzyme involved in the hydroxylation of com-
Sankyo, respectively, are examples of synthetic variants of the pactin, enables high level fermentation of the correct form of
naturally occurring statins lovastatin and compactin, and have pravastatin to facilitate efficient industrial-scale statin drug
superior pharmacokinetic properties (4, 5). Lovastatin and production.
compactin are produced by filamentous fungi such as Aspergillus
terreus and Penicillium citrinum, respectively (Fig. 1A). In recent Author contributions: K.J.M., M.H., W.B.v.S., A.W.M., and M.A.v.d.B. designed research;
K.J.M., M.H., B.M., W.B.v.S., A.V., K.L.T., and J.-M.v.d.L. performed research; K.J.M., M.H.,
CHEMISTRY

years, their biosynthetic gene clusters were identified and com- B.M., W.B.v.S., A.V., K.L.T., J.-M.v.d.L., D.L., A.W.M., and M.A.v.d.B. analyzed data; and
prise a set of nine proteins, including a HMG-CoA reductase, a K.J.M., M.H., D.L., A.W.M., and M.A.v.d.B. wrote the paper.
transporter, and a transcriptional activator. Two large polyketide The authors declare no conflict of interest.
synthases produce the nonaketide precursor and methylbutyrate This article is a PNAS Direct Submission.
side chain. Additional oxidoreductase and hydroxylase enzymes Freely available online through the PNAS open access option.
lead to the production of lovastatin and compactin (6, 7). Simva-
Data deposition: The atomic coordinates and structure factors have been deposited in the
statin (Fig. 1B) is a semisynthetic variant of lovastatin wherein Protein Data Bank, www.pdb.org (PDB ID code 4OQS and 4OQR). The sequence reported
the methylbutyrate side chain is converted into a dimethylbutyrate in this paper has been deposited in the GenBank database (accession no. KF751385).
moiety. This outcome can be achieved by direct methylation or by 1
K.J.M. and M.H. contributed equally to this work.
deacylation and subsequent addition of a de novo chemically 2
To whom correspondence may be addressed. Email: andrew.munro@manchester.ac.uk
synthesized dimethylbutyrate group (8). Pravastatin (Fig. 1B) is or marco.berg-van-den@dsm.com.
obtained after stereoselective hydroxylation of compactin at the This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
C-6 position. Industrially, this is achieved by a two-step production 1073/pnas.1419028112/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1419028112 PNAS | March 3, 2015 | vol. 112 | no. 9 | 2847–2852


Fig. 1. Structures of relevant statins and deriva-
tives. (A) Naturally produced statins (i, compactin;
ii, lovastatin). (B) Semisynthetic pharmaceutical
statins [i, pravastatin (Pravachol); ii, simvastatin (Zocor)].
(C) Alternative compactin oxidation products (i, 6-epi-
pravastatin; ii, 3-α-iso-pravastatin). (D) Deacylated
derivatives [i, deacylated compactin (ML-236A);
ii, deacylated pravastatin).

(13). Here, we present a radically different approach: development methylbutyrate side chain of compactin (SI Results), most likely
of a one-step, fully fermentative pravastatin production process. by an inducible esterase or lipase. Due to numerous candidate
The industrial high producing P. chrysogenum strain DS17690 was genes in the P. chrysogenum genome (>200) (16), a biochemical
genetically reprogrammed by introducing the full compactin bio- approach was used to identify the responsible enzyme. Com-
synthetic pathway and an evolved compactin hydroxylase CYP pactin producing P. chrysogenum strains were cultivated under
from Amycolatopsis orientalis, leading to an efficient pravastatin statin deacylating (induced by using urea as nitrogen source; SI
producing strain. Results) and control (i.e., noninducing, by using yeast extract as
nitrogen source) conditions (Fig. 2B). Proteins from cell lysates
Results were fractionated by hydrophobic interaction chromatography
P. chrysogenum as a Platform for Compactin Production. The β-lactam and fractions with high compactin deacylation activity analyzed
antibiotics producer P. chrysogenum is a well-developed industrial by SDS/PAGE. Relevant bands were characterized by MS
microorganism that underwent numerous rounds of classical strain analysis after tryptic digestion. A protein encoded by Pc15g00720
improvement, leading to current penicillin production titers of (GenBank accession no. CAP82958.1), a putative esterase, was
more than 50 g/L (14). Because these accumulated mutations only detectable in samples with increased statin deacylation. De-
contributed to the industrial robustness and the high flux from leting this gene reduced statin deacylation to an absolute minimum,
glucose to secondary metabolites, we speculated that this highly even after cultivation with urea as sole N source (Fig. S2).
adapted organism would be suitable for producing other sec-
ondary metabolites. To produce other pharmaceuticals in this Identification of an A. orientalis Compactin Hydroxylase. The S. carbo-
species, an isolate completely devoid of β-lactam antibiotics was philus P450sca-2 enzyme (9) functions using NADPH-ferredoxin
needed. Deleting all penicillin biosynthetic genes in a high pro- oxidoreductase and ferredoxin redox partners. Attempts to express
ducing strain led to such a host (15). To test whether the dele- P450sca-2 directly in our P. chrysogenum compactin-producing
tions had any adverse effect on secondary metabolite production strain did not lead to detectable pravastatin (Table S2). Several
potential, we reintroduced the biosynthetic genes, resulting in eukaryotic and prokaryotic systems capable of hydroxylating
the recovery of penicillin production (Table S1). Next, the compactin have been reported (17–22), but none give 100%
complete compactin (ML-236B) gene cluster from P. citrinum conversion nor show real improvements over S. carbophilus.
(7) was subcloned on three plasmids (Fig. S1). All nine mlc genes Moreover, as the natural P450sca-2 redox partners are unknown
were kept under control of their natural P. citrinum promoter and we required compactin hydroxylase functionality in a fungal
and terminator regions, assuming that these would function simi- host, we set out to isolate a new hydroxylating enzyme that is
larly in P. chrysogenum. Transformants were selected on acetamide functional in eukaryotes. Several microorganisms were screened
due to cotransformation of an acetamidase expression cassette. for compactin to pravastatin conversion, and the actinomycete
As DNA integration in P. chrysogenum occurs almost exclusively A. orientalis showed the greatest hydroxylation activity (up to 100%;
via the nonhomologous end joining (NHEJ) pathway, all frag- SI Materials and Methods). To clone the relevant gene, a genome
ments randomly integrate in the genome. Successful integration library was constructed and screened for compactin hydroxylation.
was confirmed by PCR and Southern blot analyses (SI Materials
and Methods). Stable transformants harboring all DNA frag-
ments were cultivated in synthetic medium and analyzed for
statin production. As expected, all three fragments must be
cotransformed in one strain to enable statin formation (Table
S2). Several strains produced relatively high amounts of statins:
up to 624 mg/L in shake flasks (Fig. 2A). This titer is remarkable,
because the natural compactin producer P. citrinum produces only
19 mg/L (Fig. 2A). These results confirm that the P. chrysogenum
β-lactam free platform strain retains its features for high sec-
ondary metabolite productivities, even for heterologous prod-
ucts. Moreover, the P. citrinum promoters are very efficient in
P. chrysogenum and do not underperform compared with known
strong homologous promoters such as pcbC (Table S2).
Fig. 2. Statin production levels in Penicillium strains. Red, ML-236A; blue,
Deletion of Esterase Activity Improves Compactin Yield in P. chrysogenum. compactin. (A) Left bar, P. citrinum NRRL 8082; other bars, P. chrysogenum
A surprisingly high ratio of deacylated (ML-236A) vs. acylated strains. (B) Influence of nitrogen source on ratio of deacylated (ML-236A) to
(compactin) statins (Fig. 1) in P. chrysogenum (up to 60%; Fig. 2A) acylated (compactin) statin. (a) Sodium nitrite; (b) yeast extract; (c) arginine;
suggested an imbalance in the introduced pathway or a competing (d) lysine; (e) urea; (f) acetamide; (g) ammonium sulfate; (h) hydroxylamine;
activity. Initial experiments confirmed in vivo cleavage of the (i) peptone; (j) ammonium nitrite.

2848 | www.pnas.org/cgi/doi/10.1073/pnas.1419028112 McLean et al.


Two positive clones were isolated, and sequence analysis revealed
a complete ORF with 40–60% amino acid identity to known CYP
enzymes in one of the clones (Fig. S3), containing a conserved heme
binding motif around the cysteine ligand to the CYP heme iron
(FGGGIHQCLG) (5). This ORF was cloned into the Escherichia
coli expression vector pBAD-DEST49 under control of the minimal
arabinose (araBAD) promoter, and activity assays confirmed the
CYP enzyme as a compactin hydroxylase (cmpH; accession no.
KF751385; CYP105AS1; Fig. S3).

Expression of CYP105AS1 in P. chrysogenum Leads to epi-Pravastatin.


One-step fermentative production of pravastatin requires the
A. orientalis CYP105AS1 to be functional in compactin-producing
P. chrysogenum. Various publications describe the use of re-
ductase domains from self-sufficient CYPs, such as the RhF from
Rhodococcus sp. and BM3 from Bacillus megaterium, for acti-
vation of type I CYP enzymes (23–25). Such hybrid fusion Fig. 3. Expression and evolution of the A. orientalis CYP105AS1. (A) Sche-
enzymes may catalyze oxidation reactions in absence of other matic representation of the self-sufficient CYP105AS1 expression construct.
oxidoreductase/ferredoxin partners. A codon-optimized version (B) Statin production in P. chrysogenum transformants.
of CYP105AS1 was fused in frame to the RhF reductase (FMN-
and 2Fe-2S cluster-binding) domain, and the resulting hybrid gene
was expressed under the strong pcbC promoter of P. chrysogenum highest pravastatin:6-epi-pravastatin ratios revealed that, in
(Fig. 3A). On introduction to P. chrysogenum compactin-producing most cases, a single amino acid change altered stereoselectivity.
strains, 19 of 192 transformants were identified that produced sig- Interestingly, in addition to active site mutations, outer shell
nificant amounts of pravastatin (up to 550 mg/L; Fig. 3B). mutations were also identified. Furthermore, expression mutants
The conversion activity of CYP105AS1 expressed in P. chrysogenum were isolated (#1; I233T) that did not alter the ratio but im-
was good, but not complete: ∼90% of the secreted statin was proved activity fourfold. The 10 best mutants were taken as
pravastatin (Fig. 3B; WT and WT-CpO). However, detailed templates and site-saturated fusion PCR used to achieve all
analyses with NMR and liquid chromatography (LC)-MS/MS possible combinations. Screening these allowed isolation of
revealed that, although the correct compactin C-6 was hydrox- mutants with corresponding pravastatin:6-epi-pravastatin ratios
ylated, the wrong stereoisomer 6-epi-pravastatin (Fig. 1C) was up to 96:4 (Table S3). Thus, with only two rounds of mutagenesis
produced almost exclusively. This isomer was also reported for S. and three to five amino acid changes, the ratio of the compactin
carbophilus bioconversion, but at around 10% (9). Here, E. coli hydroxylation products pravastatin, 6-epi-pravastatin was com-
and P. chrysogenum strains harboring CYP105AS1 produced 98– pletely inverted, and CYP105AS1 was engineered into a pravastatin
100% 6-epi-pravastatin and maximally only ∼2% pravastatin synthase (P450Prava).
(Table S3 and Fig. 3B). Although producing the nondesired
epimer, the transformants demonstrated that functional expres- Characterization of P450Prava Explains Enantiomeric Specificity. To
sion of the self-sufficient bacterial CYP in P. chrysogenum is characterize in vitro the influence of selected mutations on
sufficient to produce high amounts of hydroxylated statin. compactin affinity and CYP structure, CYP105AS1, five in-
dividual point mutants and combinations thereof (Table S4)
Evolution of CYP105AS1 Toward an Efficient Pravastatin Synthase. were expressed and purified. UV-visible (UV-Vis) spectros-
The CYP105AS1 structure was solved to 2.05 Å using molecular copy revealed a typical CYP heme spectrum for WT CYP105AS1,
replacement with the vitamin D hydroxylase CYP105A1 (26) with a Soret maximum at 419 nm, and α/β (Q) bands at 570/537 nm,
[Protein Data Bank (PDB) ID code 2ZBY; 44% sequence respectively. Subtle variations were observed for mutant forms
identity]. The final structure contains residues Glu10 to Asp400 (Table S4). Both compactin and pravastatin induced Soret shifts
with the exception of the polypeptide stretch from Ser77 to (toward 393 nm) through displacement of the heme iron axial
Ala85, with an RMSD of 1.38 Å for 355 Ca atoms (Fig. 4), and is water ligand, causing a low spin (LS) to high spin (HS) shift of
most similar to the filipin I:CYP105P1 complex (27) (PDB ID the ferric heme iron. CYP105AS1 did not fully convert to HS on
code 3ABA; 42% sequence identity). The disordered nature of saturation with compactin, displaying an ∼47% conversion and a

MICROBIOLOGY
the BC-loop stretch and the open position of the FG-loop ren- Kd value of 29.3 ± 1.1 μM (Table S4). The compactin-saturated
ders the heme pocket and adjacent I-helix solvent accessible in WT CYP105AS1 spectrum has a prominent HS shoulder at
the ligand-free state, as is observed for other CYP105 members 395 nm, and a LS Soret peak shifted from 419 to 417 nm.
(including CYP105P1). Extensive attempts to obtain a cocrystal All individual mutations improve both compactin affinity and
structure of CYP105AS1 with compactin/pravastatin revealed the extent of HS shift and had additive effects. P450Prava dis-
uninterpretable electron density in the active site (Fig. S4), plays the best Kd (1.4 ± 0.4 μM), which is a 21-fold improve-
suggesting there is ample space for the substrate to occupy ment over WT CYP105AS1, with near full conversion to the
multiple configurations. We hypothesized that minor changes in substrate-bound HS form and a Soret maximum at 393 nm
protein structure could influence and/or restrict the conforma- (Fig. S5).
CHEMISTRY

tional landscape of the substrate to achieve different stereo- The cocrystal structure of P450Prava with compactin at 1.8 Å
chemistry, resulting in increased pravastatin production at the led to a readily interpretable electron density for the ligand (Fig. 4).
expense of 6-epi-pravastatin. To this end, the CYP105AS1 gene Comparison with the CYP105AS1 open structure shows that
was mutated by error-prone PCR so that each clone contained compactin binding to P450Prava reorientates the BC-loop region
two to three mutations on average. PCR products were cloned in (with the Pro82-Ala88 region now disordered) and reorganizes
pBAD-DEST under control of arabinose induction. Per round, the FG helices. This closing motion shields the ligand and active
5,000 clones were analyzed for compactin hydroxylation, focus- site from bulk solvent, as previously observed in other CYP105
ing on the ratio of pravastatin to 6-epi-pravastatin. After a single enzymes (26, 27). Compactin displaces the aqua-sixth ligand
round of error-prone PCR, mutants with significantly improved from the heme iron and is bound in an active conformation with
pravastatin:6-epi-pravastatin ratios were identified (Table S3). the C6 carbon placed within 4.7 Å of the heme iron (Fig. 4 B and
Although CYP105AS1 (WT) hydroxylates compactin with a D). Direct protein-ligand interactions are limited to hydrophobic
pravastatin:6-epi-pravastatin ratio of 3:97, the best mutant (#6) contacts with residues from the BC-loop (Phe76, Pro80, Trp93,
had a ratio of 48:52. Sequence analysis of mutants with the Thr95), the FG-loop (Met179, Val180, Val182), the I-helix

McLean et al. PNAS | March 3, 2015 | vol. 112 | no. 9 | 2849


Discussion
Process improvement of existing pharmaceuticals is mostly in-
cremental, involving classical strain and process improvement
for fermentation processes and optimizing yields for chemical
processes. For the semisynthetic molecule pravastatin, process
improvement was also limited to classical strain improvement
and to the cloning of new hydroxylases. Here, we applied a
radical approach and completely redesigned the pravastatin
production process. We hypothesized that using existing in-
dustrial fermentation strains would shorten lead times and in-
crease opportunities to obtain desired strains and product levels.
We used an antibiotic-producing strain of P. chrysogenum as
a chassis. P. chrysogenum is well known for β-lactam production,
but this is an undesirable feature when producing other phar-
maceuticals (29). Therefore, we used a variant in which the
β-lactam biosynthetic genes are removed (15), essentially putting
into practice the suggestion made by Jami et al. (30) to use
P. chrysogenum for other biotechnological purposes. Reintroduction
Fig. 4. Structure of the WT CYP105AS1 and P450Prava. (A) An overlay of the
of the penicillin biosynthetic genes proved that all capabilities were
WT CYP105AS1 (in blue) and the P450Prava mutant structure (in green). The
retained in this mutant, and introducing the compactin cluster
positions of the mutations are indicated by spheres. (B) Stereo-and regio-
selectivity of the P450 prava mutant. The compactin substrate (yellow) is
demonstrated that beneficial mutations are not restricted to a single
orientated such that the C6 Hpro-S (β-configuration) is positioned directly
compound of interest but have general advantages for production of
adjacent to the heme iron. The heme is shown in red spheres with the heme secondary metabolites. These results demonstrated that regulatory
iron in orange. The compactin C6 carbon is in cyan. (C) The WT CYP105AS1 mechanisms from the donor species, P. citrinum, were fully
open active site, with key residues shown in atom colored sticks. (D) The functional in the recipient P. chrysogenum. Although not often
corresponding view of the P450Prava:compactin complex (compare with C). mentioned in the literature, degradation of statins via cleavage of
Compactin is shown in atom colored sticks (yellow carbons) with the corre- the methylbutyrate side chain was already reported (31, 32).
sponding 2Fo-Fc electron density contoured at 1σ as a blue mesh. However, in the P. chrysogenum background, this reaction was
much more pronounced, and efficient pravastatin production
requires strict control via the nitrogen source or by deleting the
(lle236, Val239, Ala240, Thr244), the K-β1 loop (Val282), and relevant esterase, encoded by Pc15g00720. It is tempting to
the C-terminal region (Ala388, Ala389). Indirect hydrogen bonding speculate that non–statin-producing fungi like P. chrysogenum
contacts, mediated via water molecules, occur between the com- have evolved this esterase in nature to protect themselves from
pactin ester C8-substituent and residues Trp93 and Thr95, in ad- the antifungal properties of statins (33), whereas statin-producing
dition to indirect contact made between the carbonyl lactone species solved this by an additional gene copy of the target enzyme
moiety and Thr286. Three of the five mutations are located HMG-CoA reductase and a specific statin transporter (34); how-
ever, evidence supporting such hypotheses is lacking.
within van der Waals contact of compactin (I95T, A180V, and Initial attempts to use the available P450sca-2 to hydroxylate
L236I), each leading to improved surface complementarity be- compactin directly in fungi were unsuccessful, and we attempted
tween ligand and protein (Fig. 4). In addition, the I95T mu- several strategies to improve its function. The S. carbophilus
tation strengthens the polar interaction network made with genome was sequenced and identified a ferredoxin and several
compactin. In contrast, Q127R and A265N are distant from the ferredoxin oxidoreductases, which were cointroduced with both
active site and unlikely to have direct effects on ligand affinity/ native and codon-optimized P450sca-2 gene variants (under
conformation. In the WT structure, Gln127 is part of an exten- control of strong P. chrysogenum promoters) into compactin
sive network of polar contacts made between the EG loop and producing strains, but these approaches did not lead to pravastatin
the D-helix (Fig. S6). This network changes on active site clo- production (Table S2). Several studies have engineered CYPs by
sure, due to motion of the G-helix and EG-loop with respect to fusing them to reductase modules in efforts to enhance their cata-
the D-helix. We suggest that the Q127R mutation disrupts the lytic rates (25, 35). However, fusion constructs between P450sca-2
and the P450BM3 reductase or the reductase module of the
open conformation network in particular, thus lowering the en-
Rhodococcus RhF P450-reductase fusion enzyme also did not
ergetic barrier to the formation of the closed conformation and result in compactin to pravastatin conversion. Recently, P450sca-2
favoring ligand binding (similar to gatekeeper mutations recently was successfully mutated and expressed in E. coli (36, 37). Inter-
described for P450BM3) (28). In contrast, the contribution of the estingly, the authors did not select a fusion to the Rhf reductase as
A265N mutation, located at the protein surface and distant from the reductase crystal structure is unknown (38), but selected the
regions involved in the open to closed transition, cannot be ra- Pseudomonas putida reductase/putidaredoxin system (38, 39),
tionalized from the structural data available. leading to 377.5 mg/L pravastatin from this three-component
redox system.
Expression of P450Prava in a Compactin Production Strain. P450Prava In contrast, our search for novel CYPs led to the isolation of
with a pravastatin:6-epi-pravastatin ratio of 96:4 was fused to the A. orientalis CYP105AS1, which readily hydroxylates compactin
Rhf reductase and randomly integrated in the best P. chrysogenum in E. coli. A. orientalis was included in our screen as it was
compactin strain. The resulting transformants produced large reported to contain broad substrate-range CYPs (40). Although
amounts of pravastatin; the highest isolate produced 688 mg/L stand-alone CYP105AS1 showed low activity in P. chrysogenum,
it was successfully expressed fused to RhF reductase, producing
(Fig. 3B). Most importantly, the improved stereoselectivity ratio
a fully functional enzyme both in E. coli and P. chrysogenum.
pravastatin:epi-pravastatin of the mutant P450Prava as observed in The CYP105AS1 structure reveals an open conformation and
E. coli was also confirmed in all P. chrysogenum transformants. suggests multiple compactin binding modes rather than a distinct
To assess industrial performance of the P450Prava-P. chrys- enzyme-substrate complex that could guide rational engineering
ogenum strain, 10-L fed-batch fermentations were carried out (SI for improved stereoselectivity. CYP mutations often result in
Materials and Methods). After a 200-h fermentation time, the novel product mixes and therefore new specificities (41), and
strain produced more than 6 g/L pravastatin (Fig. S7), with only we successfully realigned CYP105AS1 stereospecificity in two
minor amounts (<0.5 g/L) of deacylated statins and compactin. mutation rounds, without compromising expression levels. Most

2850 | www.pnas.org/cgi/doi/10.1073/pnas.1419028112 McLean et al.


beneficial mutations were identified in the active site above the Table 1. HPLC time table gradient for the separation of statins
heme (42). The P450Prava/compactin structure reveals how con- Time (min) A% B% Flow (mL/min) Curve
formational freedom of the substrate is restricted to essentially
a single conformation, where substrate is positioned with the 0.00 67 33 0.50 1
pro-S hydrogen from the C-6 carbon close to the heme iron, 0.65 67 33 0.50 6
explaining the enantiomeric excess of the mutant. Three muta- 1.00 10 90 0.50 6
tions, in particular, appear to reshape the active site, enhancing 1.45 10 90 0.50 6
ligand affinity and redefining the conformational landscape. This 1.50 67 33 0.50 6
structural change explains the significant difference in regio- and 1.80 67 33 0.50 6
stereoselectivity compared with CYP105AS1, also outperforming
P450sca-2 (9). P450sca-2 produces 10% 6-epi-pravastatin (9),
whereas the best CYP105AS1 mutants produce only 4%; this is
0.1 min for the Load Ahead option. As needle wash, 400 μL acetonitrile/
a decisive difference when considering industrial production at water (50:50) was used.
a several-ton scale. A second unwanted byproduct of hydroxylation,
The MS was operated with an Electrospray source in the positive mode
3α-iso-pravastatin (Fig. 1C; previously reported as 6α-iso-ML-236B) (ES+) with the capillary at 3.00 kV and a cone voltage of 35 V. The source
(9, 31), was hardly observed in our experiments. temperature was 120 °C, and the desolvation temperature was 360 °C. The
The production titers (>6 g/L) obtained by our engineered desolvation gas flow was 600 L/h, and the cone gas flow was 50 L/h. The MS
strain (Fig. S7) are much better than the classical two-step prava- settings were as follows: extractor, 3 V; RF lens, 0.3 V; LM resolution, 15.0;
statin production, which yields 2–3 g/L pravastatin (20, 43). HM resolution, 15.0; ion energy, 0.3 V; multiplier, 650 V; dwell time, 10 ms;
Pravastatin secretion was fully operational in P. chrysogenum interscan delay, 50 ms; interchannel delay, 10 ms; span, 0. Using this
when all compactin biosynthetic genes and the P450Prava mutant protocol, statins were measured in the selected ion recording mode as the
were randomly integrated. These findings suggest that the sodium adducts m/z 447.25 for the pravastatin isomers and at m/z 431.25 for
P. citrinum protein MlcE is also heterologously functional as compactin. The compounds elute at the following retention times: prava-
a transporter for all statins studied, including pravastatin. statin, 0.61 min; 6-epi-pravastatin, 0.66 min; 3α-iso-pravastatin, 0.78 min;
To our knowledge, our results are the first example of har- and compactin, 1.25 min.
nessing the potential of a previously improved industrial pro-
duction strain in a more generic way: the generation of a Expression of CYP105AS1 and Mutants in P. chrysogenum. Expression of
platform strain, harboring beneficial mutations, which can be functional CYPs was done using a synthetic version of pDONR-P4-P1r carrying
used for rapid development of novel production strains for an expression cassette consisting of the P. chrysogenum pcbC promoter, the
unrelated chemicals. cmpH gene fused in-frame to the Rhodococcus RhF reductase (23) (codon
optimized for expression in P. chrysogenum), and the P. chrysogenum penDE
Materials and Methods terminator (accession no. KF754797). NdeI (CATATG including the CYP ATG
start codon) and SpeI (located in the flexible linker region between the CYP
Additional procedures can be found in SI Materials and Methods. and reductase domains) sites were engineered to allow introduction of any
CYP variant. Linearized plasmids were transformed to P. chrysogenum.
Strains. E. coli DH10B and TOP10 (Invitrogen) were used for gene cloning.
P. chrysogenum strains used were laboratory strain Wisconsin 54-1255 (44), Evolution of CYP105AS1: Construction of an Error-Prone Library. Mutants of
the high penicillin producer DS17690 (16), the single penicillin gene cluster CYP105AS1 were generated using the Diversify PCR Random Mutagenesis Kit
derivative DS47274 (15), and the β-lactam free platform strain DS50662 (15). (Clontech) followed by Gateway cloning. Four different concentrations of
For compactin reference fermentations and genomic DNA isolation, P. citrinum manganese salt were evaluated. Libraries were constructed in a 15-μL BP
NRRL8082 was used (NRRL). reaction using 100 ng PCR-amplified CYP, 200 ng pDONR221, 3 μL BP Clonase II
enzyme mix, and Tris-EDTA buffer (pH 8.0) and incubated for 4 h at 25 °C.
Media. P. chrysogenum was cultivated for 168 h at 25 °C in synthetic medium Subsequent transfer to the expression vector [10 μL BP mix, 2 μL (150 ng/μL)
(SI Materials and Methods) at 250 (shake flask) or 400–550 rpm (microtiter pBAD-DEST49, 3 μL LR clonase II] was done for 2 h at 25 °C. The reaction was
plate; MTP). To produce penicillin V, 2 g/L phenoxyacetic acid was added. stopped by addition of 4 μg proteinase K (10 min, 37 °C). One microliter was
P. citrinum spores were harvested from potato dextrose agar (PDA) slants in transformed to E. coli TOP10, and 48 individual clones of each library were
0.5% Tween 80 and cultivated as described previously (45). picked to determine mutation frequency.

Construction of a Compactin Producing P. chrysogenum Strain. The P. citrinum Screening for Improved Pravastatin:6-epi-Pravastatin Conversion Ratio. Five
compactin gene cluster was cloned in three parts. The central and 3′-end (14 thousand individual clones from the error-prone library, with an average of

MICROBIOLOGY
and 6 kb) were readily PCR amplified and cloned using Gateway technology two to three nucleotide mutations per insert, were used to inoculate 96-MTP
(Invitrogen) into entry vectors pDONRP4-P1R and pDONR221, respectively containing 200 μL/well Luria-Bertani (LB) medium + 100 μg/mL ampicillin and
(Fig. S1). The 18-kb fragment was cloned in two steps. First, the 10- and 8-kb grown overnight at 37 °C. Subsequently, 40 μL of the overnight culture was
fragments were cloned separately in pCR2.1 TOPO T/A (Invitrogen) and transferred to deep well 96-MTP filled with 500 μL LB + 100 μg/mL ampicillin
combined afterward via compatible restriction ends. The small piece of and 1 mM δ-aminolevulinate. Plates were grown for 4 h at 30 °C, 300 rpm.
remaining and interfering oligonucleotide sequence in mlcA was removed Subsequently, 0.01% arabinose was added, followed by overnight in-
via exchange of an internal genomic PCR amplified mlcA fragment (4.1-kb cubation at 22 °C, 300 rpm. Cells were spun down and resuspended in 300 μL
Acc65I fragment). Ligation of all fragments produced the full-length 18-kb LB medium containing 100 mM phosphate buffer (pH 6.8), 0.2% glucose,
in pDONR41Zeo. All PCR amplified fragments were verified via sequencing. 0.2 g/L compactin, 1 mM δ-aminolevulinate, and 0.01% arabinose. After over-
CHEMISTRY

Linearized plasmids were transformed to P. chrysogenum as described pre- night incubation at 30 °C, 300 rpm, cells were spun down, and supernatant was
viously (15). extracted with an equal volume of methanol for LC-MS analysis.

LC-MS Analysis of Statins. A Waters Acquity ultraperformance liquid chro- Second Evolution Round: Site Saturation and Error-Prone PCR. DNA sequences
matography (UPLC) coupled to a Micromass ZQ 2000 mass spectrometer and encoding improved enzymes were PCR amplified, purified, and pooled in
controlled by Waters MassLynx software was used for LC-MS analysis. LC equimolar concentrations (10 ng/μL each). Using this as template DNA, site-
separation was performed on a Waters Acquity UPLC ethylene bridged hybrid saturation PCR was done for six positions (Table S5). DNA fragments were
Phenyl column with dimensions 50 × 2.1-mm ID (particle size, 1.7 μm) and isolated, mixed in equimolar amounts, and used in a fusion PCR under error-
with a flow of 0.5 mL/min and temperature of 60 °C. The binary mobile prone PCR conditions. No additional PCR primers were added. The resulting
phase consisted of water with 0.1% formic acid (A) and acetonitrile with library was cloned in pBAD-DEST49 as above.
0.1% formic acid (B) (Table 1).
The samples were kept at 10 °C and injected with a 2-μL sample loop, Crystallization and Structural Determination of CYP105AS1 (WT) and the
a needle volume of 30 μL, and at a syringe draw rate of 30 μL/min. The full P450Prava Mutant. Pure WT CYP105AS1 and P450prava mutant proteins (10–
loop injection was used with an overfill factor of 5.6 and loop offline after 13 mg/mL) were crystallized in 10 mM Tris and 50 mM NaCl, pH 7.5.

McLean et al. PNAS | March 3, 2015 | vol. 112 | no. 9 | 2851


Compactin (in DMSO) was added to the protein to a final concentration of pH 6.5, 200 mM sodium bromide, and 22.5% (wt/vol) PEG 3350. Single
1 mM with ≤2% (vol/vol) DMSO for cocrystallogenesis. Protein was centri- crystals were soaked in mother liquor supplemented with 10% (vol/vol) PEG
fuged (18,800 × g, 10 min, 4 °C) immediately before crystallization to remove 200 as cryoprotectant and flash-cooled by plunge freezing in liquid nitrogen.
any precipitate. Initial screens were performed using a Mosquito liquid Diffraction data were collected at Diamond synchrotron beamlines. Data
handling robot (TTP LabTech) and soluble protein 96-well screens (Molecular were reduced/scaled with the XDS suite (47) and processed using the CCP4 suite
Dimensions). Sitting drops (400 nL) were made by mixing equal volumes of (48). The WT CYP105AS1 structure was solved using molecular replacement with
protein solution and mother liquor and incubating at 4 °C. Crystallization the CYP105A1 structure as a search model (PDB ID code 2ZBY) (26), and the
conditions were identified using the Clear Screen Strategy I (CSS-1), Mor- P450Prava mutant was solved using the WT CYP105AS1 structure. Structures were
pheus, and PACT premier screens (Molecular Dimensions). Crystal microseeding refined using Refmac5 (48) and COOT (49). Data and final refinement statistics
procedures were used to improve crystal forms (46). Microseed stocks were are shown in Table S6.
prepared by using the Seed Bead kit (Hampton Research) with crystals in 50 μL
of mother liquor from the same condition. Sitting drops were set up as before, ACKNOWLEDGMENTS. We acknowledge access to Diamond beamlines and
including the microseed at a ratio of 4:1 protein to microseed before mixing Colin Levy for assistance with synchrotron X-ray data collection. We thank
with mother liquor. Diffraction quality crystals of the CYP105AS1 WT protein Marina Golovanova, Rolf Poldermans, Erwin Talens, Paul Klaassen, and
Adriana Carvalho-de Souza for technical support. These studies were funded
were obtained with 0.2 M ammonium chloride, 0.1 M Mes, pH 6.0, and 20%
by UK Biotechnology and Biological Sciences Research Council Industrial
(wt/vol) PEG 6000, or 0.3 M sodium acetate, 0.1 M Tris, pH 8.5, 10% (wt/vol) PEG Partnership Award Research Grant BB/G014329/1 with DSM (to A.W.M., D.L.,
8000, and 10% (wt/vol) PEG 1000. Diffraction quality crystals of the P450Prava and K.J.M.). This work was sponsored by DSM Anti-Infectives (presently DSM
compactin-bound mutant were obtained using 100 mM Bis-Tris propane at Sinochem Pharmaceuticals).

1. Weng TC, Yang YHK, Lin SJ, Tai SH (2010) A systematic review and meta-analysis on 24. Fuziwara S, et al. (2002) Catalytically functional flavocytochrome chimeras of P450
the therapeutic equivalence of statins. J Clin Pharm Ther 35(2):139–151. BM3 and nitric oxide synthase. J Inorg Biochem 91(4):515–526.
2. Lamon-Fava S (2013) Statins and lipid metabolism: An update. Curr Opin Lipidol 24(3): 25. Sieber V, Martinez CA, Arnold FH (2001) Libraries of hybrid proteins from distantly
221–226. related sequences. Nat Biotechnol 19(5):456–460.
3. Roth BD (2002) The discovery and development of atorvastatin, a potent novel hy- 26. Sugimoto H, et al. (2008) Crystal structure of CYP105A1 (P450SU-1) in complex with
polipidemic agent. Progress in Medicinal Chemistry, eds King FD, Oxford AW (Elsevier 1alpha,25-dihydroxyvitamin D3. Biochemistry 47(13):4017–4027.
Science BV, Amsterdam), Vol 40, pp 1–22. 27. Xu L-H, et al. (2010) Regio- and stereospecificity of filipin hydroxylation sites revealed
4. Yoshino G, et al. (1986) Effect of CS-514, an inhibitor of 3-hydroxy-3-methylglutaryl by crystal structures of cytochrome P450 105P1 and 105D6 from Streptomyces aver-
coenzyme A reductase, on lipoprotein and apolipoprotein in plasma of hypercholestero- mitilis. J Biol Chem 285(22):16844–16853.
lemic diabetics. Diabetes Res Clin Pract 2(3):179–181. 28. Butler CF, et al. (2013) Key mutations alter the cytochrome P450 BM3 conformational
5. Serizawa N (1996) Biochemical and molecular approaches for production of pravastatin, landscape and remove inherent substrate bias. J Biol Chem 288(35):25387–25399.
a potent cholesterol-lowering drug. Biotechnol Annu Rev 2:373–389. 29. Smith RG (2008) Penicillin and cephalosporin drug allergies: A paradigm shift. J Am
6. Kennedy J, et al. (1999) Modulation of polyketide synthase activity by accessory Podiatr Med Assoc 98(6):479–488.
proteins during lovastatin biosynthesis. Science 284(5418):1368–1372. 30. Jami MS, et al. (2010) The Penicillium chrysogenum extracellular proteome. Conver-
7. Abe Y, et al. (2002) Molecular cloning and characterization of an ML-236B (compactin) sion from a food-rotting strain to a versatile cell factory for white biotechnology. Mol
biosynthetic gene cluster in Penicillium citrinum. Mol Genet Genomics 267(5):636–646. Cell Proteomics 9(12):2729–2744.
8. Thaper RK, Kumar Y, Kumar SMD, Misra S, Khanna JM (1999) A cost-efficient synthesis 31. Serizawa N, et al. (1983) 6 alpha-Hydroxy-iso-ML-236B (6 alpha-hydroxy-iso-compactin)
of simvastatin via high-conversion methylation of an alkoxide ester enolate. Org and ML-236A, microbial transformation products of ML-236B. J Antibiot (Tokyo)
Process Res Dev 3(6):476–479. 36(7):918–920.
9. Matsuoka T, et al. (1989) Purification and characterization of cytochrome P-450sca 32. Komagata D, Yamashita H, Endo A (1986) Microbial conversion of compactin (ML-236B)
from Streptomyces carbophilus. ML-236B (compactin) induces a cytochrome P-450sca to ML-236A. J Antibiot (Tokyo) 39(11):1574–1577.
in Streptomyces carbophilus that hydroxylates ML-236B to pravastatin sodium 33. Galgóczy L, Nyilasi I, Papp T, Vágvölgyi C (2011) Statins as antifungal agents. World J
(CS-514), a tissue-selective inhibitor of 3-hydroxy-3-methylglutaryl-coenzyme-A Clin Infect Dis 1(1):4–10.
reductase. Eur J Biochem 184(3):707–713. 34. Abe Y, et al. (2002) Effect of increased dosage of the ML-236B (compactin) bio-
10. Guengerich FP (2005) Human cytochrome P450 enzymes. Cytochrome P450: Structure, synthetic gene cluster on ML-236B production in Penicillium citrinum. Mol Genet
Mechanism and Biochemstry, 3rd Ed), ed Ortiz de Montellano PR (Kluwer Academic/ Genomics 268(1):130–137.
Plenum Publishers, New York), pp 377–530. 35. Dodhia VR, Fantuzzi A, Gilardi G (2006) Engineering human cytochrome P450 en-
11. Podust LM, Sherman DH (2012) Diversity of P450 enzymes in the biosynthesis of zymes into catalytically self-sufficient chimeras using molecular Lego. J Biol Inorg
natural products. Nat Prod Rep 29(10):1251–1266. Chem 11(7):903–916.
12. Neuvonen PJ (2010) Drug interactions with HMG-CoA reductase inhibitors (statins): 36. Li P, Guan H, Li J, Lin Z (2009) Heterologous expression, purification, and character-
The importance of CYP enzymes, transporters and pharmacogenetics. Curr Opin In- ization of cytochrome P450sca-2 and mutants with improved solubility in Escherichia
vestig Drugs 11(3):323–332. coli. Protein Expr Purif 65(2):196–203.
13. Hosobuchi M, Kurosawa K, Yoshikawa H (1993) Application of computer to moni- 37. Fujii T, Fujii Y, Machida K, Ochiai A, Ito M (2009) Efficient biotransformations using
toring and control of fermentation process: Microbial conversion of ML-236B Na to Escherichia coli with tolC acrAB mutations expressing cytochrome P450 genes. Biosci
pravastatin. Biotechnol Bioeng 42(7):815–820. Biotechnol Biochem 73(4):805–810.
14. van den Berg MA (2011) Impact of the Penicillium chrysogenum genome on industrial 38. Ba L, Li P, Zhang H, Duan Y, Lin Z (2013) Semi-rational engineering of cytochrome
production of metabolites. Appl Microbiol Biotechnol 92(1):45–53. P450sca-2 in a hybrid system for enhanced catalytic activity: Insights into the impor-
15. Harris DM, et al. (2009) Exploring and dissecting genome-wide gene expression re- tant role of electron transfer. Biotechnol Bioeng 110(11):2815–2825.
sponses of Penicillium chrysogenum to phenylacetic acid consumption and penicillinG 39. Ba L, Li P, Zhang H, Duan Y, Lin Z (2013) Engineering of a hybrid biotransformation
production. BMC Genomics 10:75. system for cytochrome P450sca-2 in Escherichia coli. Biotechnol J 8(7):785–793.
16. van den Berg MA, et al. (2008) Genome sequencing and analysis of the filamentous 40. Basch J, Chiang S-J (2007) Cloning and expression of a cytochrome P450 hydroxylase
fungus Penicillium chrysogenum. Nat Biotechnol 26(10):1161–1168. gene from Amycolatopsis orientalis: Hydroxylation of epothilone B for the pro-
17. Chen C-H, Hu H-Y, Cho Y-C, Hsu W-H (2006) Screening of compactin-resistant mi- duction of epothilone F. J Ind Microbiol Biotechnol 34(2):171–176.
croorganisms capable of converting compactin to pravastatin. Curr Microbiol 53(2): 41. Jung ST, Lauchli R, Arnold FH (2011) Cytochrome P450: Taming a wild type enzyme.
108–112. Curr Opin Biotechnol 22(6):809–817.
18. Fujii Y, et al. (2011) Construction of a novel expression vector in Pseudonocardia 42. Trefzer A, et al. (2007) Biocatalytic conversion of avermectin to 4″-oxo-avermectin:
autotrophica and its application to efficient biotransformation of compactin to prava- Improvement of cytochrome p450 monooxygenase specificity by directed evolution.
statin, a specific HMG-CoA reductase inhibitor. Biochem Biophys Res Commun Appl Environ Microbiol 73(13):4317–4325.
404(1):511–516. 43. Manzoni M, Rollini M (2002) Biosynthesis and biotechnological production of statins
19. Lin C-L, Tang Y-L, Lin S-M (2011) Efficient bioconversion of compactin to pravastatin by filamentous fungi and application of these cholesterol-lowering drugs. Appl
by the quinoline-degrading microorganism Pseudonocardia carboxydivorans isolated Microbiol Biotechnol 58(5):555–564.
from petroleum-contaminated soil. Bioresour Technol 102(22):10187–10193. 44. Elander RP (2002) University of Wisconsin contributions to the early development of
20. Park JW, et al. (2003) Bioconversion of compactin into pravastatin by Streptomyces sp. penicillin and cephalosporin antibiotics. SIM News 52:270–278.
Biotechnol Lett 25(21):1827–1831. 45. Benedetti A, Manzoni M, Nichele M, Rollini M (2002) Process for the production of
21. Peng Y, Yashphe J, Demain AL (1997) Biotransformation of compactin to pravastatin pravastatin and lovastatin. Patent EP1266967.
by Actinomadura sp. 2966. J Antibiot (Tokyo) 50(12):1032–1035. 46. D’Arcy A, Villard F, Marsh M (2007) An automated microseed matrix-screening
22. Serizawa N, et al. (1983) Microbial hydroxylation of ML-236B (compactin). Studies on method for protein crystallization. Acta Crystallogr D Biol Crystallogr 63(Pt 4):550–554.
microorganisms capable of 3 beta-hydroxylation of ML-236B. J Antibiot (Tokyo) 36(7): 47. Kabsch W (2010) XDS. Acta Crystallogr D Biol Crystallogr 66(Pt 2):125–132.
887–891. 48. Winn MD, et al. (2011) Overview of the CCP4 suite and current developments. Acta
23. Nodate M, Kubota M, Misawa N (2006) Functional expression system for cytochrome Crystallogr D Biol Crystallogr 67(Pt 4):235–242.
P450 genes using the reductase domain of self-sufficient P450RhF from Rhodococcus 49. Emsley P, Cowtan K (2004) Coot: Model-building tools for molecular graphics. Acta
sp. NCIMB 9784. Appl Microbiol Biotechnol 71(4):455–462. Crystallogr D Biol Crystallogr 60(Pt 12 Pt 1):2126–2132.

2852 | www.pnas.org/cgi/doi/10.1073/pnas.1419028112 McLean et al.

You might also like