ArticlePDF Available

Diverse Combinatorial Biosynthesis Strategies for C-H Functionalization of Anthracyclinones

Authors:

Abstract and Figures

Streptomyces spp. are “nature’s antibiotic factories” that produce valuable bioactive metabolites, such as the cytotoxic anthracycline polyketides. While the anthracyclines have hundreds of natural and chemically synthesized analogues, much of the chemical diversity stems from enzymatic modifications to the saccharide chains and, to a lesser extent, from alterations to the core scaffold. Previous work has resulted in the generation of a BioBricks synthetic biology toolbox in Streptomyces coelicolor M1152ΔmatAB that could produce aklavinone, 9-epi-aklavinone, auramycinone, and nogalamycinone. In this work, we extended the platform to generate oxidatively modified analogues via two crucial strategies. (i) We swapped the ketoreductase and first-ring cyclase enzymes for the aromatase cyclase from the mithramycin biosynthetic pathway in our polyketide synthase (PKS) cassettes to generate 2-hydroxylated analogues. (ii) Next, we engineered several multioxygenase cassettes to catalyze 11-hydroxylation, 1-hydroxylation, 10-hydroxylation, 10-decarboxylation, and 4-hydroxyl regioisomerization. We also developed improved plasmid vectors and S. coelicolor M1152ΔmatAB expression hosts to produce anthracyclinones. This work sets the stage for the combinatorial biosynthesis of bespoke anthracyclines using recombinant Streptomyces spp. hosts.
Content may be subject to copyright.
Diverse Combinatorial Biosynthesis Strategies for CH
Functionalization of Anthracyclinones
Rongbin Wang,
Benjamin Nji Wandi,
Nora Schwartz,
Jacob Hecht,
Larissa Ponomareva,
Kendall Paige, Alexis West, Kathryn Desanti, Jennifer Nguyen, Jarmo Niemi, Jon S. Thorson,
Khaled A. Shaaban,*Mikko Metsä-Ketelä,*and S. Eric Nybo*
Cite This: https://doi.org/10.1021/acssynbio.4c00043
Read Online
ACCESS Metrics & More Article Recommendations *
Supporting Information
ABSTRACT: Streptomyces spp. are “nature’s antibiotic factories” that
produce valuable bioactive metabolites, such as the cytotoxic
anthracycline polyketides. While the anthracyclines have hundreds of
natural and chemically synthesized analogues, much of the chemical
diversity stems from enzymatic modifications to the saccharide chains
and, to a lesser extent, from alterations to the core scaold. Previous
work has resulted in the generation of a BioBricks synthetic biology
toolbox in Streptomyces coelicolor M1152ΔmatAB that could produce
aklavinone, 9-epi-aklavinone, auramycinone, and nogalamycinone. In
this work, we extended the platform to generate oxidatively modified
analogues via two crucial strategies. (i) We swapped the ketoreductase
and first-ring cyclase enzymes for the aromatase cyclase from the
mithramycin biosynthetic pathway in our polyketide synthase (PKS)
cassettes to generate 2-hydroxylated analogues. (ii) Next, we engineered several multioxygenase cassettes to catalyze 11-
hydroxylation, 1-hydroxylation, 10-hydroxylation, 10-decarboxylation, and 4-hydroxyl regioisomerization. We also developed
improved plasmid vectors and S. coelicolor M1152ΔmatAB expression hosts to produce anthracyclinones. This work sets the stage for
the combinatorial biosynthesis of bespoke anthracyclines using recombinant Streptomyces spp. hosts.
KEYWORDS: BioBricks, synthetic biology, natural product biosynthesis, anthracyclinones, Streptomyces coelicolor, oxygenase, anticancer
INTRODUCTION
Anthracyclines are glycosylated aromatic polyketides produced
by various soil bacteria in the actinomycete family.
Doxorubicin and aclarubicin are utilized as anticancer agents
for treating various human cancers, making them some of the
broadest spectrum antineoplastic agents used in the clinic.
1
Anthracyclines inhibit the proliferation of cancer cells through
two distinctive mechanisms: histone eviction and inhibition of
topoisomerase II, leading to the scission of DNA strands.
1,2
However, anthracyclines have limitations that diminish their
clinical utility. Cancer cells can develop drug resistance to the
anthracyclines, for example, by overexpressing the p-
glycoprotein ATP binding cassette.
3
Additionally, the long-
term use of anthracyclines is associated with cardiotoxicity.
4
These observations have motivated the systematic biosynthetic
modification of anthracyclines to achieve new analogues with
advantageous properties over currently used medications,
including increased potency, decreased drug resistance, and
reduced cardiotoxicity.
5
Anthracyclines are biosynthesized by polyketide synthase
(PKS) complexes, which are composed of a minimal PKS
(minPKS) consisting of a ketoacyl synthase (KSα), chain
length factor (CLF or KSβ), and acyl carrier protein (ACP)
that catalyzes the Claisen condensation of one molecule of
acetyl-CoA or propionyl-CoA to nine molecules of malonyl-
CoA (Figure 1A). The resulting poly β-keto thioester
decaketide undergoes controlled folding by 9-ketoreductase
(9-KR), aromatase (ARO), second-/third-ring cyclase (2/3-
CYC), and oxygenase (OXY) enzymes to generate the first
stable intermediates aklanonic acid and nogalonic acid. Further
reactions by methyltransferase (MET), fourth-ring cyclase (4-
CYC), and ketoreductase (7-KR) enzymes furnish the core
tetracyclic aromatic carbon skeletons (Figure 1A).
3
Previously,
we developed a BioBricks platform for the improved
biosynthesis of anthracyclinones.
6
We developed vectors to
produce four anthracyclinone scaolds: aklavinone (1), 9-epi-
aklavinone (2), auramycinone (3), and nogalamycinone (4)
(Figure 1A). In addition, endogenous activity from the host
Received: January 22, 2024
Revised: April 11, 2024
Accepted: April 16, 2024
Research Articlepubs.acs.org/synthbio
© XXXX The Authors. Published by
American Chemical Society A
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
This article is licensed under CC-BY 4.0
Downloaded via UNIV OF TURKU on April 29, 2024 at 13:48:59 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Figure 1. Metabolic engineering strategies for CH functionalization of anthracyclinones. Biosynthesis of (A) the four anthracyclinone aglycones
14, (B) degradation products 58through the action of endogenous enzymatic activities of the host strain S. coelicolor and (C) 2-hydroxylated
target compounds 914 accessible via PKS cassette engineering utilizing stemycin B biosynthetic logic. Depiction of the diversity of post-PKS
tailoring steps on (D) doxorubicin, (E) rhodomycin, (F) komodoquinone, and (G) kosinostatin pathways.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
B
strain Streptomyces coelicolor M1152ΔmatAB led to the
production of 7-deoxygenated versions of these compounds,
including 7-deoxy-aklavinone (5), 7-deoxy-9-epi-aklavinone
(6), 7-deoxy-auramycinone (7), and 7-deoxy-nogalamycinone
(8) (Figure 1B).
Anthracycline biosynthetic gene clusters harbor extensive
gene sets to modify the core anthracyclinone structures further.
Particularly, gene products catalyzing redox chemistry for C
H functionalization, which is an important component in the
chemodiversification of anthracyclines, are abundant. The
daunorubicin pathway harbors the FAD-dependent 11-
monooxygenase DnrF (Figure 1C).
79
RdmE also catalyzes
the same reaction on the rhodomycin pathway, containing the
15-methylesterase RdmC and the methyltransferase-like RdmB
for C-10 hydroxylation (Figure 1D).
7,9,10
The komodoquinone
pathway includes enzymes for 10-decarboxylation by the C-15
esterase EamC and the methyltransferase-like EamK (Figure
1E).
11,12
Finally, the kosinostatin biosynthetic gene cluster
encodes the short-chain aldol reductase KstA16 and the
cyclase-like KstA15 that jointly catalyze C-1 hydroxylation
(Figure 1F).
13
The reaction cascade is further extended to 4-
hydroxyl regioisomerization by the NmrA-like short-chain
dehydrogenase/reductase enzymes KstA11 and KstA10
(Figure 1F).
13
In this work, we were interested in combinatorial biosyn-
thesis and CH functionalization of anthracyclinones using
two distinct approaches. The first strategy included reprogram-
ming the PKS cassettes to aord 2-hydroxylated analogues,
similar to stemycin biosynthesis.
14
This could plausibly be
achieved by excluding the KR gene and exchanging ARO/CYC
genes with those residing on aureolic acid biosynthetic
pathways (Figure 1C).
1416
The second strategy included
diverse post-PKS tailoring genes from the kosinostatin,
13
rhodomycin,
9,10
doxorubicin,
8
and komodoquinone B path-
ways.
11
One significant goal of these studies was to
systematically evaluate the substrate promiscuity of post-PKS
tailoring enzymes toward alternative substrates 14.
The metabolic engineering presented here resulted in the
generation of 26 anthracyclinones, including nine novel
analogues with regiospecific CH oxygenation. Chemical
characterization and bioactivity profiling revealed the im-
portance of 1-, 10-, and 11-hydroxylation in the cytotoxicity of
the anthracyclinones. Installing ketone, aldehyde, and alcohol
functional groups provides chemical handles for group-transfer
enzymes, such as methyltransferases, aminotransferases, and
glycosyltransferases.
12
This opens the door for rational
metabolic engineering to generate diverse glycosylated
anthracycline analogues in the future.
RESULTS AND DISCUSSION
Development of Improved Strains and Vectors. The
yields of 14from the previous PKS cassettes ranged between
1 and 5 mg/L. We applied four distinct strategies to improve
production (Figure 2A,B).
17
First, we incorporated stronger
synthetic sp41, sp42, and sp44 promoters (Table S1)
18,19
together with ribozyme-based insulator parts (e.g., sp41-vtmoJ,
Figure 2. Four metabolic engineering strategies to increase the yields of anthracyclinones. (A) SBOL diagram of redesigned PKS cassettes.
Constructs encoded the simultaneous expression of eight to ten genes (depending on the anthracyclinone) under the expression of strong sp41,
sp42, and sp44 promoters. Constructs were insulated from external genomic promoter expression by incorporating tt-sbi-A and fd-term
transcriptional terminators. Promoters were fused to ribozyme-insulator parts to stabilize the expression of the three operons within the construct.
(B) Overexpression of ssgA,scbr2, and accA2BE for anthracyclinone enhancement. SsgA triggers sporulation and cell division, which works with the
ΔmatAB mutation to enhance biomass accumulation. Scbr2 is a pseudo-γ-butyrolactone response regulator that regulates glycolytic flux. AccA2BE
enhances the supply of malonyl-CoA for anthracyclinone biosynthesis. (C) Production titers of 9-epi-aklavinone in strains expressing pEN10002
and/or coexpressing pEAKV2 with scbr2, accA2BE, and/or ssgA. (D) Production titers of nogalamycinone in strains expressing pEN10004 and/or
coexpressing pNOG2 with scbr2, accA2BE, and/or ssgA.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
C
sp42-ltsvJ, and sp44-riboJ) to eliminate interference between
the promoters and ribosome initiation sites (Figure 2A).
6,20
Ribozyme-based insulators function by cleaving the 5
untranslated region (5-UTR) of the mRNA to form a hairpin
loop to stabilize the transcript. Furthermore, the cassettes were
bracketed with transcriptional terminators (e.g., fd phage and
ttsbi-A terminators)
21,22
The new PKS cassettes were cloned
into the pOSV802 vector, allowing single-copy chromosomal
expression in Streptomyces (Table S1).
23
Two vectors, pEAKV2
encoding the production of 9-epi-aklavinone and pNOG2
encoding the output of nogalamycinone, were expressed in S.
coelicolor M1152ΔmatAB (Table S2). The improved strains
were fermented in E1 liquid media and produced 4 mg/L of 2
and 12 mg/L of 4, which represented 1- and 6-fold
improvement over the previous vectors pEN10002 and
pEN10004, respectively (p< 0.001) (Figure 2C,D). The
enhanced transcriptional stability of the constructs appeared to
contribute to improved translation of the PKS machinery and
metabolic flux to the target molecules.
Second, we increased substrate availability via acetyl-CoA
carboxylase (e.g., accA2BE) overexpression (Figure 2B),
previously employed with tetracenomycin engineering to
achieve 3-fold yield enhancement.
24,25
The acetyl-CoA
carboxylase converts acetyl-CoA to malonyl-CoA, an essential
precursor for the biosynthesis of polyketides.
26
Integration of
the pOSV808-accA2BE expression cassette into the S. coelicolor
lines resulted in a 2-fold improvement in 2(4.4 to 9.2 mg/L, p
< 0.0001) and 4(12.1 to 23 mg/L, p< 0.0001) (Figure 2C,D).
Third, we sought to increase yields by overexpressing ssgA
from Streptomyces griseus (Figure 2B), which has been shown
to suppress sporulation and enhance the fragmented growth of
mycelia, thus resulting in faster growth kinetics and increased
production of biomass and cell products.
27,28
The over-
expression of ssgA using the pENSV3 expression vector
resulted in a nearly 3-fold increase in 2(4.4 to 11.2 mg/L, p
< 0.0001) and 4(12.1 to 31.0 mg/L, p< 0.0001) production
titers (Figure 2C,D). This result demonstrated that ssgA could
improve anthracyclinone production titers.
Fourth, we overexpressed a pseudo-γbutyrolactone (GBL)
receptor scbr2 in S. coelicolor M1152ΔmatAB::cos16F4iE
(Figure 2B). Scbr2 does not bind γ-butyrolactones but has
been shown to interact with numerous endogenous and
exogenous natural products.
29,30
The regulatory eects of
Scbr2 are mediated by promoting glycolysis via upregulation of
glyceraldehyde-3-phosphate dehydrogenase (gap1) and pyr-
uvate kinase (pyk2), which we reasoned could be important for
carbon flow and growth kinetics.
31
The scbr2 gene has been
deleted in S. coelicolor M1152 as part of the cpk cluster,
32
which
has led to increased oxidative metabolism and oxidative stress
(i.e., flux through the tricarboxylic acid cycle) based on a
genome-scale metabolic model.
33
To test this hypothesis, we
overexpressed scbr2 in S. coelicolor, which resulted in a 3-fold
Figure 3. Engineering of aromatase/cyclase enzymes to biosynthesize 2-hydroxylated anthracyclinones. (AF) HPLC-UV/vis chromatograms of
dierent strains monitored at 430 nm engineered with two plasmids producing metabolites indicated with the numbered compound. (A) S.
coelicolor M1152::acc::A2C1 (2-hydroxy-aklanonic acid, 9); (B) S. coelicolor M1152::acc::S2C1 (2-hydroxy-nogalonic acid, 10); (C) S. coelicolor
M1152::acc::A2C1::A6 (2-hydroxy-aklavinone, 11); (D) S. coelicolor M1152::acc::A2C1::S6 (2-hydroxy-9-epi-aklavinone, 12); (E) S. coelicolor
M1152::acc::S2C1::A6 (2-hydroxy-auramycinone, 13); (F) S. coelicolor M1152::acc::S2C1::S6 (2-hydroxy-nogalamycinone, 14). (G) Production
titers of 2-hydroxylated anthracyclinones from strains engineered with two plasmids. (H) The production titers of strains engineered with and
without expression of the C-12 oxygenase (snoaB). (I) The 1H,13C-HMBC, 1H,1HCOSY, and 1H,1H-NOESY two-dimensional NMR correlations
for compounds 11, 12, 13, and 13b.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
D
improvement in 2(4.4 to 13.0 mg/L, p< 0.0001) and 4(12.1
to 30.0 mg/L, p< 0.0001) titers (Figure 2C,D). This result
demonstrated that overexpression of scbr2 may have balanced
glycolytic flux and relieved oxidative stress for improved
product formation, but the exact mechanism requires further
study.
Metabolic Engineering of PKS Cassettes for 2-
Hydroxylated Anthracyclinones. We sought to develop a
means for producing 2-hydroxylated analogues of 14by
reprogramming the PKS cassettes.
6
The polyketide-derived 2-
hydroxyl group is removed during anthracycline biosynthesis
by KR enzymes (Figure 1A) and, consistently, pathways
lacking these redox enzymes have resulted in the production of
2-hydroxy-aklavinone
34
and 2-hydroxy-nogalonic acid.
15,35
Here, we reasoned that coexpression of anthracycline minPKS
genes aknBCDE2For snoa123 together with ARO and 2/3-
CYC genes from nonreducing pathways,
36
such as mtmQY
involved in mithramycin biosynthesis, would result in the
production of 2-hydroxy-aklanonic acid (9) or 2-hydroxy-
nogalonic acid (10), respectively (Figure 1C). Cloning and
transformation of gene cassettes pA2C1 (aknBCDE2F+
mtmQY) and pS2C1 (snoa123 + mtmQY) resulted in the
production of 9and 10, respectively (Figures 3A,B and S1
S2).
To extend the pathway, we cloned additional cassettes
encoding MET, 4-CYC, and 7-KR (e.g., aknGHU) and pTG1-
S6 (e.g., snoaCLF) to the host strain, which resulted in the 2-
hydroxy-aklavinone (11), 2-hydroxy-9-epi-aklavinone (12), 2-
hydroxy-auramycinone (13), and 2-hydroxy-nogalamycinone
(14) based on high-performance liquid chromatography-mass
spectrometry (HPLC-MS) (Figures 1C and 3CF and Table
S3). Although product yields were reasonable (310 mg/L)
(Figure 3G), we hypothesized that additional coexpression of
an OXY gene could enhance the folding and shaping of the
unnatural 2-hydroxylated polyketides. Indeed, the coexpression
of snoaB on a separate multicopy expression vector
(pUWL201PW) under the control of the ermE*presulted in
a 50% increase in production of 11 and 13 (Figure 3H). Based
on these findings, we refactored our constructs to generate
improved versions of pHAKV2, pHEAKV2, pHAURA2, and
pHNOG2 (e.g., encoding the production of 11, 12, 13, and
14, respectively, Table S1). pHAKV2, pHEAKV2, and
pHAURA2 were used in subsequent scale-up fermentation
experiments (see the Methods Section).
To confirm the structures of 11,12, and 13, the
fermentations were scaled up in 5 L of E1 media, the
metabolites were extracted and purified using various
chromatographic techniques, and structurally characterized
based on high-resolution electrospray ionization-MS (HRESI-
MS) and NMR spectroscopy. The chemical characterization
revealed that the early pathway intermediate SEK15 (13b) was
accumulated as a side product in the scale-up fermentation of
pHAURA2. In addition, three new 2-hydroxy-anthracyclinones
(1113) were isolated, based on one-dimensional (1D) (1H
and 13C NMR) and two-dimensional (2D) (COSY, HSQC,
HMBC, TOCSY, and NOESY) NMR spectroscopic measure-
Figure 4. Enzymatic assays and metabolic engineering of 11-hydroxylated anthracyclinones. (A) DnrF catalyzes 11-hydroxylation of 14to aord
1518. (B) HPLC-UV/vis traces at 490 nm of enzymatic reactions of 14incubated with purified DnrF and no-enzyme controls. (C) HPLC-UV/
vis traces at 490 nm of S. coelicolor lines engineered with expression constructs encoding 14and dnrF or control lines producing only 14.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
E
ments (Tables S4S5 and Figures 3I and S3S46). The
molecular formulas of 1113 were established as C22H20O9,
C22H20O9, and C21H18O9based on (+)-HRESIMS, with Δm/z
= 16 higher than those of aklavinone (1), 9-epi-aklavinone (2),
and auramycinone (3), respectively, indicative of the presence
of an extra oxygen atom in 1113 (Figures S4, S17, and S27).
Comparison of the NMR data (1H and 13C NMR) of the new
compounds 11-13 with previously reported compounds 1-3
revealed that the main dierences were observed in the
aromatic ring A, where the trisubstituted aromatic rings in
compounds 1-3were converted to tetra-substituted aromatic
rings in compounds 11-13 (with two m-coupled protons, H-1/
H-3; Tables S4 and S5). The position of the hydroxy groups in
compounds 11-13 was established to be at 2-position based on
the observed HMBC correlations from H-1 to C-12/C-4a/
CH-3; 2-OH to CH-1/C-2/CH-3 and H-3 to CH-1/C-4a
(Figure 3I). All of the remaining 2D-NMR (1H, 1HCOSY,
HMBC, TOCSY, and NOESY) correlations fully agree with
structures 1113 (Figure 3I and Supporting Information). As
new natural products and are closely related to 13,
compounds 11-13 were designated as 2-hydroxy-aklavinone
(11), 2-hydroxy-9-epi-aklavinone (12), and 2-hydroxy-auramy-
cinone (2-hydroxy-9-epi-nogalamycinone; 2-hydroxy-9-epi-no-
galavinone; 13), respectively. The structure of 14 is suggested
to be 2-hydroxy-nogalamycinone.
Anthracyclinone 11-Hydroxylation by DnrF. To probe
the substrate promiscuity of post-PKS tailoring enzymes for 1
4, we carried out parallel investigations with purified enzymes
and gene expression studies. We first cloned the dnrF gene
37
from the doxorubicin pathway into the pBAD/His B vector for
expression in Escherichia coli TOP10. After the production and
purification of recombinant 11-hydroxylase DnrF, we assayed
the conversion of 14to the 11-hydroxylated species 15, 16,
17, and 18 (Figures 1D and 4A). The incubation of 1and 3
resulted in quantitative conversion to 15 and 17, whereas 2
and 4were converted to 16 and 18 with poor eciency <5%
(Figures 4B and S47S75).
For the in vivo expression experiments, dnrF was fused to the
strong gapdhpEL promoter, cloned into expression vector
pENSV3, and transformed into cell lines producing 14.
Analysis of culture extracts demonstrated that 1was converted
with >90% eciency to 15 and maggiemycin, which is a shunt
product derived from 11-hydroxylation of aklaviketone.
38
Compound 3was also converted in >90% eciency to 17,
previously isolated from fermentations of Streptomyces
coeruleorubidis ATCC 31276.
39
However, similarly to the in
Figure 5. Enzymatic assays and metabolic engineering of 10-hydroxylated and 10-decarboxylated anthracyclinones. (A) EamC and EamK catalyze
10-decarboxylation of 14to aord 1922. EamC and RdmB catalyze 10-hydroxylation of 1and 3to produce 23 and 24. (B) HPLC-UV/vis
traces at 430 nm of enzymatic reaction of 14incubated with purified EamC + RdmB, EamC + EamK, and no-enzyme controls. (C) HPLC-UV/
vis traces at 430 nm of S. coelicolor lines engineered with expression constructs encoding 14and either eamC +eamK or eamC +rdmB.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
F
vitro analyses, 2and 4were converted to 16 and 18 with poor
510% eciency (Figures 4C and S47S75), respectively.
These results demonstrated that DnrF is more selective toward
9(R)-configured than 9(S)-configured anthracyclines.
Anthracyclinone 10-Hydroxylation by RdmB and 10-
Decarboxylation EamK. Both anthracycline 10-hydroxyla-
tion and 10-decarboxylation by the SAM-dependent hydrox-
ylase RdmB and decarboxylase EamK, respectively, require
initial 15-methylesterase activity (Figures 1E,F and 5A). Since
the rhodomycin enzymes have a preference for glycosylated
substrates,
11,40
which is in contrast to the komodoquinone
enzymes that convert aglycone substrates,
11
we utilized the 15-
methylesterase EamC in all experiments. Our in vitro (Figure
5B) and in vivo (Figure 5C) data were highly convergent and
demonstrated that 14were quantitatively converted by
EamC and EamK to 10-decarboxylated products 1922
(Figures S76S91). In contrast, only 1and 3were converted
to 10-hydroxylated species 23 and 24, respectively (Figures
5B,C and S92S97), which demonstrates that RdmB exhibits
preference toward 9(R)-configured metabolites both in vitro
and in vivo.
Anthracyclinone 10- and 11-Hydroxylation by RdmE,
RdmC, and RdmB. We next sought to incorporate the
tailoring steps for concomitant 10 and 11-hydroxylation by the
entire RdmE, RdmC, and RdmB reaction cascade (Figures 1E
and 6A). We incubated 14with purified DnrF, EamC, and
RdmB, which resulted in the conversion of substrates 1and 3
to 25 and 26 (Figures 6B and S98S104). Similarly, 1and 3
were converted to 25 and 26 in Streptomyces strains
coexpressing the appropriate PKS cassettes and a cassette
expressing rdmE, rdmC, and rdmB (Figures 6C and S98
S104). Both in vitro and in vivo, the substrates 2and 4
underwent 10-decarboxylation toward 20 and 22 as the
primary metabolic route and were not processed by the
hydroxylating enzymes. The production titers of 25 and 26
from the engineered strains in SG-TES media were 30.3 ±10
and 9.3 ±4 mg/L, respectively (Figure S105). Altogether, this
indicates robust production of the β-rhodomycinone analogues
25 and 26 in the engineered microorganisms.
Anthracyclinone 1-Hydroxylation by KstA15 and
KstA16. We next utilized the two-component monooxygenase
system, KstA15 and KstA16, from the kosinostatin pathway for
Figure 6. Enzymatic assays and metabolic engineering of 10-hydroxylated and 11-hydroxylated anthracyclinones. (A) RdmE, EamC/RdmC, and
RdmB catalyze 11-hydroxylation and 10-hydroxylation of 1and 4to generate 25 and 26, respectively. (B) HPLC-UV/vis traces at 490 nm of
enzymatic reaction of 1and 3incubated with purified DnrF, EamC, RdmB, and no-enzyme controls. (C) HPLC-UV/vis traces at 490 nm of S.
coelicolor lines engineered with expression constructs encoding 1and 3and rdmE +rdmC +rdmB and control lines producing 1and 3.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
G
1-hydroxylation (Figures 1F and 7A).
13
KstA15 is a polyketide
cyclase-like enzyme,
41
while KstA16 belongs to short-chain
alcohol dehydrogenases.
42
The enzyme assays with substrates
14resulted in the quantitative conversion to 1-hydroxylated
Figure 7. Enzymatic assays and metabolic engineering of 1-hydroxylated and 4-regioisomerized anthracyclinones. (A) KstA15 and KstA16 catalyze
1-hydroxylation of 14yielding 2730, KstA10 and KstA11 carry out asymmetric reduction and dearomatization, followed by a region-specific
reduction and dehydration yielding 3134. (B) HPLC-UV/vis traces of enzymatic reaction of 14incubated with purified KstA15 and KstA16
and no-enzyme controls. (C) HPLC-UV/vis traces of S. coelicolor lines engineered with expression constructs encoding 14and kstA15 and kstA16
and control lines producing 14. (D) HPLC-UV/vis traces of enzymatic reaction of 14incubated with purified KstA15, KstA16, KstA11, and
KstA10 and no-enzyme controls. (E) HPLC-UV/vis traces of S. coelicolor lines engineered with expression constructs encoding 14and kstA15,
kstA16, kstA10, and kstA11 and control lines producing 14.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
H
anthracyclinones 2730 (Figures 7B and S106S120).
Analogously, the heterologous expression of a cassette
encoding kstA15 and kstA16 resulted in the production of 1-
hydroxylated species 2730 (Figures 7C and S106S120).
The compounds 27,29, and 30 have been generated in
previous studies,
4143
but the anthracyclinone 28 has not been
reported in the literature. The 1-hydroxylation mechanism is a
necessary modification for 1-O-glycosylation of anthracyclines,
such as that occurs with the nogalamycin family of
compounds.
44
Anthracyclinone 4-Regioisomerization by KstA15,
KstA16, KstA11, and KstA10. KstA15 and KstA16 are part
of a four-enzyme cascade to generate a hydroxy regioiso-
merized anthracyclinone. Following the successful 1-hydrox-
ylation by KstA15 and KstA16, we reconstituted the entire 4-
regioisomerization pathway by incorporating KstA11 and
KstA10 (Figures 1F and 7A).
13
The in vitro assays with the
four enzymes converted 14to iso-anthracyclinones 3134
(Figures 7D and S121S132). Several putative intermediates
were detected in the enzymatic reactions, including the
dearomatized 1,4-diketone species (indicated with #) (Figure
7D). Similarly, coexpression of the four kst genes in
appropriate S. coelicolor strains producing 14led to
accumulation of 3134 (Figures 7E and S121S132) in
yields ranging from 8 to 30 mg/L (Figure S133). Within the
ESI-MS single ion monitoring traces, the parental substrates,
unknown intermediates (marked *), and products 3134 were
detected with the expected mass ions (Table S3). These results
indicate that the kosinostatin enzymes are remarkably flexible
toward 9(R) and 9(S)-configured anthracyclinones. Previously,
31 has been isolated from strains engineered for isoanthracy-
cline production
45
and 33 is the native substrate of the
kosinostatin pathway,
13
but compounds 32 and 34 have not
been reported to date.
HUMAN CANCER CELL VIABILITY ASSAYS
To investigate the potential anticancer activity of the
anthracyclinone extracts, the crude extracts were normalized
to 40 μg/mL concentration and were tested in a panel of
human cancer cells: A549 (nonsmall cell lung), PC3
(prostate), TC32 (Ewing sarcoma), and HCT116 (colorectal)
human cancer cell lines (Table 1 and Figure S134). As
previously reported, the anthracyclinone aglycones 3and 4are
inactive at IC50 values >30 μM in A549, PC3, MKL1, and
MCC26 cancer cell lines, though 1is slightly more active in
PC3 cells at 7 μM IC50 and A549 cells at 17 μM, respectively.
6
Most of the extracts were inactive against these cancer cell
lines; however, extracts from five engineered lines containing
compounds 15,17,18,22,25,26, and 30 exhibited
substantive cytotoxic activity in the cell lines tested (<20%
cell viability T/C) (Figures S134). Compounds 11,12, and 13
were inactive at concentrations of 3 μM (Figure S135), which
indicated that 2-hydroxylation does not enhance the
cytotoxicity of these aglycones.
Anthracyclines are considered to require the glycoside
moiety for binding to DNA and inhibition of DNA
topoisomerases.
5
Therefore, the significant cytotoxicity against
the human cancer cell lines in selected extracts can be
considered to be surprising (Table 1). The data provides
guidance regarding the rational engineering of anthracyclines
toward 11-hydroxylated, 10,11-dihydroxylated, or 1-hydroxy-
lated nogalamycinone derivatives. Indeed, anthracyclines with
C-1 or C-11 hydroxyl groups were shown to have increased
potency in L1210 leukemia cells,
46
while 11-hydroxylation of
aclacinomycins resulted in markedly improved cytotoxicity in
the NCI 60-cell line assay.
47
Rhodomycin A, which is a 10,11-
dihydroxylated anthracycline, was recently identified as an
ecient antagonist for knocking down Src-associated onco-
genes.
48
Table 1. Cytotoxic Activities of the Generated Strain Extracts in PC3, TC32, A549, and HCT116 Cell Lines
b
,
c
a
The samples were culture extracts from the strains producing the numbered compound and contain other impurities. Actinomycin D and H2O2
were used as positive controls at 20 μM and 1 mM concentration, respectively (0% viable cells, n= 3. SE = standard error).
b
The cell viability of
each line was determined as the percentage of cell viability relative to untreated controls.
c
% Viability values were obtained after 72 h incubation.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
I
CONCLUSIONS
Anthracyclines have been a cornerstone of anticancer chemo-
therapy for several decades. Despite the success of these
molecules, severe side eects have made the continued
exploration of the chemical space around anthracyclines
necessary for discovering improved congeners. For example,
developing the semisynthetic idarubicin (4-demethoxy daunor-
ubicin) for acute myeloid leukemia has demonstrated the
importance of CH functionalization of anthracyclines.
49
However, the regiospecific modification of the polyaromatic
anthracycline core has remained challenging using organic
synthesis. This is in contrast to the biosynthesis of
anthracyclines in Streptomyces bacteria, where several enzy-
matic systems have evolved for CH functionalization in the
various natural metabolic pathways.
We have used our BioBricks metabolic engineering platform
to systematically probe 1-, 2-, 10-, and 11-hydroxylations, 10-
decarboxylation, and 4-hydroxyl regioisomerization. For the
first time, we analyzed the functionality of 10 tailoring genes
with the four possible anthracyclinone core structures 14
comparatively. The results demonstrate that the in vivo activity
remained consistently high in 10-decarboxylation, 1-hydrox-
ylation, and 4-hydroxyl regioisomerization (Figure 8). In
contrast, 10-hydroxylation and 11-hydroxylation were depend-
ent on the 9R-stereochemistry but were tolerant toward both
methyl and ethyl side chains at the same location. The dual 10-
and 11-hydroxylations proved the most challenging, which may
be because the 10-hydroxylase RdmB generally prefers
glycosylated substrates (Figure 8).
Our work demonstrates that tailoring steps in anthracycline
biosynthesis are well suited for combinatorial biosynthesis for
increasing the chemical diversity of natural products. We
utilized genetic material from five distinct pathways and their
use in combination with the four aglycone possibilities led to
the generation of nine novel anthracyclinones (11, 12, 13, 14,
16, 18, 24, 26, 28). It is noteworthy that even though all of the
tailoring enzymes have been extensively studied in the past,
including by structural analysis at atomic resolution, it was not
possible to predict a priori which gene combinations resulted
in new functional metabolic pathways. Therefore, the ability to
use multiplasmid expression systems in a robust heterologous
host in combination with an expanding modular BioBricks
library is essential to allow high-throughput combinatorial
work to compensate for the limitations of an unpredictable
design space. In addition, the highly concurrent in vitro and in
vivo data indicate that combinatorial enzymatic synthesis may
be an ecient preliminary tool to guide metabolic engineering
projects. Notably, the human cancer cell line viability assays
(Table 1) provide further direction for the rational engineering
of anthracyclines based on the 15,17, 25, and 26 scaolds.
These strains will serve as valuable chassis for combinatorial
biosynthesis of TDP-deoxysugar pathways to develop “new to
nature” anthracyclines, which could be developed into potent
antitumoral metabolites.
METHODS
Bacterial Strains and Growth Conditions. E. coli
TOP10 and E. coli ET12567 were grown in LB broth or LB
agar at 37 °C as previously described.
50
E. coli TOP10 was
used for plasmid propagation, subcloning, and enzyme
expression. Enzymes were cloned into pBAD/His B vector
(Invitrogen) and enzymes were expressed using the arabinose-
inducible promoter.
51
E. coli ET12567/pUZ8002 was used as
the conjugation donor host for mobilizing expression vectors
into S. coelicolor as previously described.
52
When appropriate,
ampicillin (100 μg mL1), kanamycin (25 μg mL1),
apramycin (25 μg mL1), viomycin (25 μg mL1),
spectinomycin (100 μg mL1), hygromycin (50 μg mL1),
and nalidixic acid (30 μg mL1) were supplemented to media
to select for recombinant microorganisms.
S. coelicolor derivative strains were routinely maintained on
Soya-Mannitol Flour (SFM) agar supplemented with 10 mM
MgCl2 and International Streptomyces Project medium #4
(ISP4) (BD Difco) at 30 °C as described previously.
52
For
liquid culturing, S. coelicolor derivative strains were grown in
TSB media (3 mL) to ferment the seed culture and then grown
in a modified 50 mL SG-TES liquid medium (soytone 10 g,
glucose 20 g, yeast extract 5 g, TES free acid 5.73 g, CoCl21
mg, per liter) or 50 mL E1 medium for production for four to
5 days.
53
All media and reagents were purchased from Thermo
Fisher Scientific.
Molecular Biology Procedures. Routine genetic cloning
and plasmid manipulation were carried out in E. coli DH10B
cells (New England Biolabs). E. coli ET12567/pUZ8002 was
used as the host for intergeneric conjugation with S. coelicolor
as previously described.
52
E. coli chemically competent cells
were prepared using the Mix and Go! E. coli Transformation
Kit (Zymo Research). E. coli was transformed with plasmid
DNA via chemically competent heat-shock transformation as
described previously. Plasmid DNA was isolated via the Wizard
Plus SV Minipreps DNA Purification System by following the
manufacturer’s protocols (Promega). All molecular biology
reagents and enzymes used for plasmid construction were
purchased from New England Biolabs. The conjugation donor
host E. coli ET12567/pUZ8002 was transformed with
constructs for mobilization into S. coelicolor strains, as
previously described. For each transformation, 912 inde-
pendent exconjugants were plated to DNA plates supple-
mented with antibiotics and grown for 45 days until the
formation of vegetative mycelium.
Preparation of In Vitro Samples and In Vivo Samples
for HPLC-MS Analysis. The in vitro samples’ conditions for
protein expression, purification, and enzyme assays are
described in the Supporting Information (Supporting Methods
1, 2, and 3). All reactions were checked by UHPLC (Shimadzu
Nexera LC-40 system with a diode array detector set to 430
Figure 8. Heat map of the in vivo activity of the tailoring gene
constructs utilized in this study when coexpressed with anthracycli-
none scaolds 14.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
J
and 490 nm wavelengths) using a Phenomenex Kinetex
Phenyl-Hexyl column (2.6 μm, 100 Å, 4.6 mm ×100 mm).
Method: Solvent A: 15% CH3CN/0.1% FA; solvent B:
CH3CN; flow rate: 0.5 mL/min; 02 min, 0% B; 220 min,
040% B; 2024 min, 100% B; 2429 min, 0% B.
For each in vivo experiment, four to six biological replicates
were grown in 50 mL of SG-TES liquid media in baed
Erlenmeyer flasks as previously described.
6,24,25
The shake flask
fermentations were grown in an orbital shaker for 5 days at 30
°C at 200 rpm. The entire cultures were extracted with 3
volumes of 0.1% formic acid in ethyl acetate and the extracts
were dried down in vacuo. The extracts were resuspended in 4
mL of methanol, filtered in a 0.5 μm nylon syringe filter, and
10 μL was analyzed via HPLC-MS.
The analysis of anthracyclinones was carried out on an
Agilent 1260 Infinity II LC/MSD iQ single quadrupole
instrument. In brief, 10 μL of the sample was injected via an
autosampler onto the sample loop, separated on a Poroshell
120 Phenyl-Hexyl Column (ID 2.7 μm, 4.6 mm ×100 mm),
and analyzed in gradients of solvent A (0.1% formic acid in
water) and solvent B (0.1% formic acid in acetonitrile). The
HPLC program used a constant flow rate of 0.5 mL per minute
and the following gradient steps: 0 min, 95% solvent A and 5%
solvent B; 010 min, 95% solvent A and 5% solvent B to 5%
solvent A and 95% solvent B; 1013 min, held at 5% solvent A
and 95% solvent B; 13.1 min, reequilibrate to 95% solvent A
and 5% solvent B; 13.115.1 min, 95% solvent A and 5%
solvent B. The diode array detector (DAD) was set to monitor
UV/vis absorbance at 430 and 490 nm. The ESI-MS was set to
scan from 200 m/z500 m/zfragments in positive and
negative ionization modes.
The yields were determined by comparing them to
authenticated external standard curves of 1and 4analyzed
via HPLC-MS. The conversion percentages were determined
by measuring the area under the curve of the anthracyclinone
metabolites at 430 and 490 nm wavelengths.
General Experimental Procedures. Ultravioletvisible
(UVvis) spectra were taken directly from analytical HPLC
runs and show relative intensities. The NMR spectra were
recorded on a Bruker Avance NEO 400 MHz (Bruker BioSpin
Corporation, Billerica, MA) (1H, 400.13 MHz; 13C, 100.25
MHz), Varian 500 MHz (Agilent, Santa Clara, CA) (1H, 500
MHz; 13C, 125.7 MHz), and/or Bruker Avance NEO 600
MHz NMR (1H, 600.37 MHz; 13C, 150.96 MHz)
spectrometer, equipped with triple-channel TCI 5 mm
cryoprobe (all spectra were processed using Bruker Topspin
4.1.4 version, and 2D spectra were apodized with QSINE or
SINE window functions and zero-filled to (2048 ×1024
points)). All of the spectra were analyzed and plotted using
Mnova [where δ-values were referenced to respective solvent
signals CD3OD, δH3.31 ppm, δC49.15 ppm; DMSO-d6,δH
2.50 ppm, δC39.51 ppm]. High-resolution electrospray
ionization (HRESI) mass spectra were recorded on the AB
SCIEX Triple TOF 5600 system (AB Sciex, Framingham,
MA). HPLC-UV/MS analyses were accomplished with an
Agilent InfinityLab LC/MSD mass spectrometer (MS Model
G6125B; Agilent Technologies, Santa Clara, CA) equipped
with an Agilent 1260 Infinity II Series Quaternary LC system
and a Phenomenex NX-C18 column (250 mm ×4.6 mm, 5
μm; Phenomenex, Torrance, CA) [Method A: solvent A:
H2O/0.1% formic acid, solvent B: CH3CN; flow rate: 0.5 mL
min1; 030 min, 5100% B (linear gradient); 3035 min,
100% B; 3536 min, 100%5% B; 3640 min, 5% B]. HPLC-
UV analyses were carried out in an Agilent 1260 system
equipped with a photodiode array detector (PDA) and a
Phenomenex C18 column (250 mm ×4.6 mm, 5 μm;
Phenomenex, Torrance, CA) [Method B: solvent A: H2O/
0.1% TFA, solvent B: CH3CN; flow rate: 1.0 mL min1; 030
min, 5100% B; 3035 min, 100% B; 3536 min, 1005% B;
3640 min, 5% B]. Semi-preparative HPLC were carried out in
a Agilent 1260 Infinity II (Prep HPLC) system equipped with
a Diode Array Detector (DAD) and a Gemini 5 μm C18 110 Å,
LC column 250 mm ×10 mm (Phenomenex, Torrance, CA)
[Method C: solvent A: H2O/0.025% TFA; solvent B:
CH3CN; flow rate: 5.0 mL min1; 03 min, 25% B; 310
min, 2575% B; 1016 min, 75100% B; 1618 min, 100%
B; 1819 min, 10025% B; 1920 min, 25% B]; [Method D:
solvent A: H2O/0.025% TFA; solvent B: CH3CN; flow rate:
5.0 mL min1; 03 min, 25% B; 317 min, 25100% B; 17
18 min, 100% B; 1819 min, 10025% B; 1920 min, 25%
B]; [Method E: solvent A: H2O/0.025% TFA; solvent B:
CH3CN; flow rate: 5.0 mL min1; 03 min, 25% B; 317 min,
25100% B; 1722 min, 100% B; 2223 min, 10025% B;
2327 min, 25% B]. All solvents used were of ACS grade and
purchased from Pharmco-AAPER (Brookfield, CT). Size
exclusion chromatography was performed on Sephadex LH-
20 (25100 μm; GE Healthcare, Piscataway, NJ). A549, PC3,
and HCT116 cells were obtained from ATCC (Manassas,
VA). All other reagents used were of reagent grade and
purchased from Sigma-Aldrich (Saint Louis, MO), unless
otherwise noted.
Purification of Compounds 1113 and 13b. The
reddish-brown oily crude extract (3.82 g) produced by Strain
1[S. coelicolor M1152ΔmatAB::pSET154BB-kasOp*-snoa123-
kasOp*-mtmQY-sp44-aknGHU] was dissolved in MeOH (10
mL) followed by Sephadex LH-20 (MeOH; 2.5 cm ×50 cm)
mentored by TLC to aord six fractions. LC-MS analysis
indicates that target metabolite 13 was mainly detected in
fractions F3 and F4. Semi-prep-HPLC purification (Method
C) of F3 and F4 aorded compound 13 (2-hydroxy-
auramycinone; 24.4 mg) in pure form as red solid, and
compound 13b (SEK15; 8.5 mg) in pure form as pale-yellow
solid.
The reddish-brown oily crude extract (3.20 g) produced by
Strain 2 [S. coelicolorM1152ΔmatAB::pHEAKV2] was frac-
tionated using silica gel column (DCM/050% MeOH; 2.5
cm ×30 cm) to aord six fractions F1 (DCM; 0.5 L), F2
(DCM/2% MeOH; 0.5 L), F3 (DCM/4% MeOH; 0.5 L), F4
(DCM/10% MeOH; 0.5 L), F5 (DCM/20% MeOH; 0.5 L),
and F6 (DCM/50% MeOH; 0.5 L), followed by LC-MS
analysis. The target compound was detected in F3 and F4.
Fractions F3F4 were combined followed by semi-prep-HPLC
purification (Method D) to aord compound 12 (2-hydroxy-
9-epi-aklavinone; 4.2 mg) in pure form as red solid.
In the same manner, the crude extract (2.22 g, reddish-
brown oily) produced by Strain 3 [S. coelicolor M1152Δma-
tAB::pSET-A2M1A6] was fractionated using silica gel column
(DCM/050% MeOH; 2.0 cm ×25 cm) to yielded six
fractions F1 (DCM; 0.5 L), F2 (DCM/2% MeOH; 0.5 L), F3
(DCM/4% MeOH; 0.5 L), F4 (DCM/10% MeOH; 0.5 L), F5
(DCM/20% MeOH; 0.5 L), and F6 (DCM/50% MeOH; 0.5
L). LC-MS analysis of the obtained fractions indicates that the
target compound was detected in fractions F24, however,
with low yield. Fractions F24 have been combined, dissolved
in MeOH (2 mL) followed by Sephadex LH-20 (1 cm ×40
cm; MeOH) and semi-prep-HPLC purification (Method E) to
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
K
give compound 11 (2-hydroxy-aklavinone; 1.56 mg) in pure
form as red solid.
Statistical Analyses. The statistical significance of the
impact of genetic manipulations and combinatorially assessed
variables on production was assessed via post hoc analysis.
One-way ANOVA, two-way ANOVA, and Student’s ttest
analyses were performed using GraphPad Prism version 10.2.1
for Mac OS X, GraphPad Software, San Diego, CA, www.
graphpad.com.
Cancer Cell Line Viability Assay. Mammalian cell line
cytotoxicity [A549 (nonsmall cell lung) and PC3 (prostate),
TC32 (Ewing sarcoma), and HCT116 (colorectal) human
cancer cell lines] assays were accomplished in triplicate
following our previously reported protocols.
24,5457
Actino-
mycin D (A549 and PC3) was used as positive controls.
Physicochemical Properties of Compounds 1113. 2-
Hydroxy-aklavinone (11).C22H20O9(428); red solid; HPLC-
Rt= 24.72 min (Supporting Information, Figures S9S21);
UV/vis λmax 228, 250 (sh), 268, 290 (sh), 445 nm; 1H NMR
(DMSO-d6, 600 MHz) and 13C NMR (DMSO-d6, 150 MHz),
see Tables S1 and S2; ()-ESI-MS: m/z427 [M H];
(+)-ESI-MS: m/z411 [(M-H2O) + H]+, 393 [(M-H2O) +
H]+; (+)-HRESI-MS: m/z393.0942 [(M-2H2O) + H]+(calcd
for C22H17O7, 393.0969), 879.1973 [2M + Na]+(calcd for
C44H40O18Na, 879.2106).
2-Hydroxy-9-epi-aklavinone (12). C22H20O9(428); red
solid; HPLC-Rt= 22.47 min (Supporting Information, Figures
S22S31); UV/vis λmax 228, 268, 290 (sh), 445 nm; 1H NMR
(DMSO-d6, 600 MHz) and 13C NMR (DMSO-d6, 150 MHz),
see Tables S1 and S2; ()-ESI-MS: m/z427 [M H];
(+)-ESI-MS: m/z393 [(M-2H2O) + H]+, 879 [2M + Na]+;
(+)-HRESI-MS: m/z393.0958 [(M-2H2O) + H]+(calcd for
C22H17O7, 393.0969), 451.0986 [M + Na]+(calcd for
C22H20O9Na, 451.0999), 879.2020 [2M + Na]+(calcd for
C44H40O18Na, 879.2106).
2-Hydroxy-auramycinone (2-Hydroxy-9-epi-nogalamyci-
none; 2-Hydroxy-9-epi-nogalavinone; 13). C21H18O9(414);
red solid; HPLC-Rt= 30.39 min (Supporting Information,
Figures S32S40); UV/vis λmax 228, 258, 270, 290, 430 nm;
1H NMR (DMSO-d6, 600 MHz) and 13C NMR (DMSO-d6,
150 MHz), see Tables S1 and S2; ()-ESI-MS: m/z413 [M
H]; (+)-ESI-MS: m/z397 [(M-H2O) + H]+, 379 [(M-
2H2O) + H]+, 851 [2M + Na]+; (+)-HRESI-MS: m/z
379.0806 [(M-2H2O) + H]+(calcd for C21H15O7,
379.0812), 851.1746 [2M + Na]+(calcd for C42H36O18Na,
851.1793).
SEK15 (13b). C20H16O8(384); pale-yellow solid; HPLC-Rt
= 26.81 min (Supporting Information, Figures S42S51);
UV/vis λmax 210, 290, 320 (sh) nm; 1H NMR (CD3OD, 600
MHz) and 13C NMR (CD3OD, 150 MHz), see Tables S1 and
S2; ()-ESI-MS: m/z383 [M H], 767 [2M H];
(+)-ESI-MS: m/z385 [M + H]+; (+)-HRESI-MS: m/z
385.0893 [M + H]+(calcd for C20H17O8, 385.0918),
407.0679 [M + Na]+(calcd for C20H16O8Na, 407.0737),
769.1791 [2M + H]+(calcd for C40H32O16Na, 767.1763).
ASSOCIATED CONTENT
*
Supporting Information
The Supporting Information is available free of charge at
https://pubs.acs.org/doi/10.1021/acssynbio.4c00043.
Plasmid and strain information; chemical schemas;
NMR data for compounds 1113 and SEK15; HPLC
chromatograph traces of the in vitro and in vivo
comparisons; mass spectra; high-resolution mass spectra,
and mammalian cell cytotoxicity data (PDF)
AUTHOR INFORMATION
Corresponding Authors
Khaled A. Shaaban Center for Pharmaceutical Research and
Innovation and Department of Pharmaceutical Sciences,
College of Pharmacy, University of Kentucky, Lexington,
Kentucky 40536, United States; orcid.org/0000-0001-
7638-4942; Email: Khaled_shaaban@uky.edu
Mikko Metsä-Ketelä Department of Life Technologies,
University of Turku, FIN-20014 Turku, Finland;
orcid.org/0000-0003-3176-2908; Email: mianme@utu.fi
S. Eric Nybo Department of Pharmaceutical Sciences,
College of Pharmacy, Ferris State University, Big Rapids,
Michigan 49307, United States; orcid.org/0000-0001-
7884-7787; Email: EricNybo@Ferris.edu
Authors
Rongbin Wang Department of Life Technologies, University
of Turku, FIN-20014 Turku, Finland
Benjamin Nji Wandi Department of Life Technologies,
University of Turku, FIN-20014 Turku, Finland;
orcid.org/0000-0003-1071-3111
Nora Schwartz Department of Pharmaceutical Sciences,
College of Pharmacy, Ferris State University, Big Rapids,
Michigan 49307, United States
Jacob Hecht Department of Pharmaceutical Sciences,
College of Pharmacy, Ferris State University, Big Rapids,
Michigan 49307, United States
Larissa Ponomareva Center for Pharmaceutical Research
and Innovation and Department of Pharmaceutical Sciences,
College of Pharmacy, University of Kentucky, Lexington,
Kentucky 40536, United States
Kendall Paige Department of Pharmaceutical Sciences,
College of Pharmacy, Ferris State University, Big Rapids,
Michigan 49307, United States
Alexis West Department of Pharmaceutical Sciences, College
of Pharmacy, Ferris State University, Big Rapids, Michigan
49307, United States
Kathryn Desanti Department of Pharmaceutical Sciences,
College of Pharmacy, Ferris State University, Big Rapids,
Michigan 49307, United States
Jennifer Nguyen Department of Pharmaceutical Sciences,
College of Pharmacy, Ferris State University, Big Rapids,
Michigan 49307, United States; orcid.org/0000-0002-
4557-1638
Jarmo Niemi Department of Life Technologies, University of
Turku, FIN-20014 Turku, Finland; orcid.org/0000-
0002-7447-8379
Jon S. Thorson Center for Pharmaceutical Research and
Innovation and Department of Pharmaceutical Sciences,
College of Pharmacy, University of Kentucky, Lexington,
Kentucky 40536, United States; orcid.org/0000-0002-
7148-0721
Complete contact information is available at:
https://pubs.acs.org/10.1021/acssynbio.4c00043
Author Contributions
R.W., B.N.W., N.S., and J.H. contributed equally to this work.
M.M.-K, K.A.S., and S.E.N. conceived and designed the study.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
L
R.W. and B.N.W. performed cloning, protein expression, and
enzymatic assays. N.S., J.H., A.W., K.P., J.N., and K.S.
performed molecular biology, actinomycete transformation,
metabolic engineering, and chemical profiling of extracts from
engineered Streptomyces lines. S.E.N., J.H., and N.S. scaled up
fermentations and isolated compounds. K.A.S. purified and
performed NMR spectroscopic analyses and HRESI mass
spectrometric studies. L.P. carried out cancer cell line viability
assays and curated data. M.M.-K., K.A.S., S.E.N., L.P., and
J.S.T. wrote, edited, and revised the manuscript.
Notes
The authors declare no competing financial interest.
ACKNOWLEDGMENTS
Research reported in this publication was supported by the
National Science Foundation under grant nos. ENG-2015951
and ENG-2321976 (S.E.N.), by the National Cancer Institute
of the National Institutes of Health under Award No.
R15CA252830 (S.E.N.), by National Institutes of Health
grant R37 AI052218 (J.S.T.), the Center of Biomedical
Research Excellence (COBRE) for Translational Chemical
Biology (CTCB, NIH P20 GM130456), the National Institute
of Food and Agriculture (USDA-NIFA-CBGP, Grant No.
2023-38821-39584), the University of Kentucky College of
Pharmacy, the University of Kentucky Markey Cancer Center,
and the National Center for Advancing Translational Sciences
(UL1TR000117 and UL1TR001998). This work was
supported by the Research Council of Finland (grants
340013 and 354998 to M.M.-K.) The authors also thank the
College of Pharmacy PharmNMR Center for analytical
support. NMR data was acquired on a Bruker AVANCE
NEO 400 MHz NMR spectrometer funded or a Bruker
AVANCE NEO 600 MHz high-performance digital NMR
spectrometer [supported, in part, by NIH grants P20
GM130456 (J.S.T.) and S10 OD28690].
REFERENCES
(1) Minotti, G.; Menna, P.; Salvatorelli, E.; Cairo, G.; Gianni, L.
Anthracyclines: Molecular Advances and Pharmacologic Develop-
ments in Antitumor Activity and Cardiotoxicity. Pharmacol. Rev.
2004,56 (2), 185229.
(2) Pang, B.; Qiao, X.; Janssen, L.; Velds, A.; Groothuis, T.;
Kerkhoven, R.; Nieuwland, M.; Ovaa, H.; Rottenberg, S.; van
Tellingen, O.; Janssen, J.; Huijgens, P.; Zwart, W.; Neefjes, J. Drug-
Induced Histone Eviction from Open Chromatin Contributes to the
Chemotherapeutic Effects of Doxorubicin. Nat. Commun. 2013,4(1),
No. 1908.
(3) Hulst, M. B.; Grocholski, T.; Neefjes, J. J. C.; van Wezel, G. P.;
Metsä-Ketelä, M. Anthracyclines: Biosynthesis, Engineering and
Clinical Applications. Nat. Prod. Rep. 2022,39, 814841,
DOI: 10.1039/D1NP00059D.
(4) Bayles, C. E.; Hale, D. E.; Konieczny, A.; Anderson, V. D.;
Richardson, C. R.; Brown, Kv.; Nguyen, J. T.; Hecht, J.; Schwartz, N.;
Kharel, M. K.; Amissah, F.; Dowling, T. C.; Nybo, S. E. Upcycling the
Anthracyclines: New Mechanisms of Action, Toxicology, and
Pharmacology. Toxicol. Appl. Pharmacol. 2023,459, No. 116362.
(5) Brown, K. V.; Wandi, B. N.; Metsä-Ketelä, M.; Nybo, S. Pathway
Engineering of Anthracyclines: Blazing Trails in Natural Product
Glycodiversification. J. Org. Chem. 2020,85 (19), 1201212023.
(6) Wang, R.; Nguyen, J.; Hecht, J.; Schwartz, N.; Brown, K. V.;
Ponomareva, L. V.; Niemczura, M.; Dissel, D.; Van; Wezel, G. P.;
Van; Thorson, J. S.; Metsä-Ketelä, M.; Shaaban, K. A.; Nybo, S. E. A
BioBricks Metabolic Engineering Platform for the Biosynthesis of
Anthracyclinones in Streptomyces Coelicolor. ACS Synth. Biol. 2022,
11, 41934209, DOI: 10.1021/acssynbio.2c00498.
(7) Niemi, J.; Mantsala, P. Nucleotide Sequences and Expression of
Genes from Streptomyces Purpurascens That Cause the Production
of New Anthracyclines in Streptomyces Galilaeus. J. Bacteriol. 1995,
177 (10), 29422945.
(8) Filippini, S.; Solinas, M. M.; Breme, U.; Schluter, M. B.;
Gabellini, D.; Biamonti, G.; Colombo, A. L.; Garofano, L.
Streptomyces Peucetius Daunorubicin Biosynthesis Gene, DnrF:
Sequence and Heterologous Expression. Microbiology 1995,141 (4),
10071016.
(9) Niemi, J.; Wang, Y.; Airas, K.; Ylihonko, K.; Hakala, J.; Mäntsälä,
P. Characterization of Aklavinone-11-Hydroxylase from Streptomyces
Purpurascens. Biochim. Biophys. Acta, Protein Struct. Mol. Enzymol.
1999,1430 (1), 5764.
(10) Wang, Y.; Niemi, J.; Mäntsälä, P. Modification of Aklavinone
and Aclacinomycins in Vitro and in Vivo by Rhodomycin Biosynthesis
Gene Products. FEMS Microbiol. Lett. 2002,208 (1), 117122.
(11) Grocholski, T.; Yamada, K.; Sinkkonen, J.; Tirkkonen, H.;
Niemi, J.; Metsä-Ketelä, M. Evolutionary Trajectories for the
Functional Diversification of Anthracycline Methyltransferases. ACS
Chem. Biol. 2019,14 (5), 850856.
(12) Dinis, P.; Tirkkonen, H.; Wandi, B. N.; Siitonen, V.; Niemi, J.;
Grocholski, T.; Metsä-Ketelä, M. Evolution-Inspired Engineering of
Anthracycline Methyltransferases. PNAS Nexus 2023,2(2),
No. pgad009, DOI: 10.1093/pnasnexus/pgad009.
(13) Zhang, Z.; Gong, Y.-K.; Zhou, Q.; Hu, Y.; Ma, H.-M.; Chen, Y.-
S.; Igarashi, Y.; Pan, L.; Tang, G.-L. Hydroxyl Regioisomerization of
Anthracycline Catalyzed by a Four-Enzyme Cascade. Proc. Natl. Acad.
Sci. U.S.A. 2017,114 (7), 15541559.
(14) Gullón, S.; Olano, C.; Abdelfattah, M. S.; Brana, A. F.; Rohr, J.;
Méndez, C.; Salas, J. A. Isolation, characterization, and heterologous
expression of the biosynthesis gene cluster for the antitumor
anthracycline steffimycin. Appl. Environ. Microbiol. 2006,72 (6),
41724183.
(15) Zabala, D.; Song, L.; Dashti, Y.; Challis, G. L.; Salas, J. A.;
Méndez, C. Heterologous Reconstitution of the Biosynthesis Pathway
for 4-Demethyl-Premithramycinone, the Aglycon of Antitumor
Polyketide Mithramycin. Microb. Cell Fact. 2020,19 (1), No. 111.
(16) Wang, G.; Chen, J.; Zhu, H.; Rohr, J. One-Pot Enzymatic Total
Synthesis of Presteffimycinone, an Early Intermediate of the
Anthracycline Antibiotic Steffimycin Biosynthesis. Org. Lett. 2017,
19 (3), 540543.
(17) Metsä-Ketelä, M. Evolution Inspired Engineering of Antibiotic
Biosynthesis Enzymes. Org. Biomol. Chem. 2017,15, 40364041.
(18) Bai, C.; Zhang, Y.; Zhao, X.; Hu, Y.; Xiang, S.; Miao, J.; Lou,
C.; Zhang, L. Exploiting a Precise Design of Universal Synthetic
Modular Regulatory Elements to Unlock the Microbial Natural
Products in Streptomyces. Proc. Natl. Acad. Sci. U.S.A. 2015,112 (39),
1218112186, DOI: 10.1073/pnas.1511027112.
(19) Wang, W.; Li, X.; Wang, J.; Xiang, S.; Feng, X.; Yang, K. An
Engineered Strong Promoter for Streptomycetes. Appl. Environ.
Microbiol. 2013,79 (14), 44844492.
(20) Lou, C.; Stanton, B.; Chen, Y. J.; Munsky, B.; Voigt, C. A.
Ribozyme-Based Insulator Parts Buffer Synthetic Circuits from
Genetic Context. Nat. Biotechnol. 2012,30 (11), 11371142.
(21) Otsuka, J.; Kunisawa, T. Characteristic Base Sequence Patterns
of Promoter and Terminator Sites in ΦX174 and Fd Phage DNAs. J.
Theor. Biol. 1982,97 (3), 415436.
(22) Huff, J.; Czyz, A.; Landick, R.; Niederweis, M. Taking Phage
Integration to the next Level as a Genetic Tool for Mycobacteria.
Gene 2010,468 (12), 819.
(23) Aubry, C.; Pernodet, J. L.; Lautru, S. Modular and Integrative
Vectors for Synthetic Biology Applications in Streptomyces Spp. Appl.
Environ. Microbiol. 2019,85 (16), No. e00485-19, DOI: 10.1128/
AEM.00485-19.
(24) Tirkkonen, H.; Brown, K. V.; Niemczura, M.; Faudemer, Z.;
Brown, C.; Ponomareva, L. V.; Helmy, Y. A.; Thorson, J. S.; Nybo, S.
E.; Metsä-Ketelä, M.; Shaaban, K. A. Engineering BioBricks for
Deoxysugar Biosynthesis and Generation of New Tetracenomycins.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
M
ACS Omega 2023,8, 2123721253, DOI: 10.1021/acsome-
ga.3c02460.
(25) Nguyen, J. T.; Riebschleger, K. K.; Brown, K. V.; Gorgijevska,
N. M.; Nybo, S. E. A BioBricks Toolbox for Metabolic Engineering of
the Tetracenomycin Pathway. Biotechnol. J. 2022,17, No. 2100371.
(26) Ryu, Y. G.; Butler, M. J.; Chater, K. F.; Lee, K. J. Engineering of
Primary Carbohydrate Metabolism for Increased Production of
Actinorhodin in Streptomyces Codicolor. Appl. Environ. Microbiol.
2006,72 (11), 71327139.
(27) van Wezel, G. P.; Krabben, P.; Traag, B. A.; Keijser, B. J. F.;
Kerste, R.; Vijgenboom, E.; Heijnen, J. J.; Kraal, B. Unlocking
Streptomyces Spp. for Use as Sustainable Industrial Production
Platforms by Morphological Engineering. Appl. Environ. Microbiol.
2006,72 (8), 52835288.
(28) Sevillano, L.; Vijgenboom, E.; van Wezel, G. P.; Díaz, M.;
Santamaría, R. I. New Approaches to Achieve High Level Enzyme
Production in Streptomyces Lividans. Microb. Cell Fact. 2016,15 (1),
No. 28.
(29) Xu, G.; Wang, J.; Wang, L.; Tian, X.; Yang, H.; Fan, K.; Yang,
K.; Tan, H. Pseudo” γ-Butyrolactone Receptors Respond to Antibiotic
Signals to Coordinate Antibiotic Biosynthesis. J. Biol. Chem. 2010,285
(35), 2744027448.
(30) Wang, W.; Ji, J.; Li, X.; Wang, J.; Li, S.; Pan, G.; Fan, K.; Yang,
K. Angucyclines as Signals Modulate the Behaviors of Streptomyces
Coelicolor. Proc. Natl. Acad. Sci. U.S.A. 2014,111 (15), 56885693.
(31) Li, X.; Wang, J.; Li, S.; Ji, J.; Wang, W.; Yang, K. ScbR-and
ScbR2-Mediated Signal Transduction Networks Coordinate Complex
Physiological Responses in Streptomyces Coelicolor. Sci. Rep. 2015,5
(1), No. 14831.
(32) Gomez-Escribano, J. P.; Bibb, M. J. Engineering Streptomyces
Coelicolor for Heterologous Expression of Secondary Metabolite
Gene Clusters. Microb. Biotechnol. 2011,4(2), 207215.
(33) Kumelj, T. S.; Sulheim, S.; Wentzel, A.; Almaas, E. Predicting
Strain Engineering Strategies Using IKS1317: A Genome-Scale
Metabolic Model of Streptomyces Coelicolor. Biotechnol. J. 2019,
14, No. 1800180, DOI: 10.1002/biot.201800180.
(34) Tsukamoto, N.; Fujii, I.; Ebizuka, Y.; Sankawa, U. Nucleotide
Sequence of the AknA Region of the Aklavinone Biosynthetic Gene
Cluster of Streptomyces Galilaeus. J. Bacteriol. 1994,176 (8), 2473
2475.
(35) Gullón, S.; Olano, C.; Abdelfattah, M. S.; Brana, A. F.; Rohr, J.;
Méndez, C.; Salas, J. A. Isolation, Characterization, and Heterologous
Expression of the Biosynthesis Gene Cluster for the Antitumor
Anthracycline Steffimycin. Appl. Environ. Microbiol. 2006,72 (6),
41724183.
(36) Caldara-Festin, G.; Jackson, D. R.; Barajas, J. F.; Valentic, T. R.;
Patel, A. B.; Aguilar, S.; Nguyen, M.; Vo, M.; Khanna, A.; Sasaki, E.;
Liu, H.; Tsai, S.-C. Structural and Functional Analysis of Two Di-
Domain Aromatase/Cyclases from Type II Polyketide Synthases.
Proc. Natl. Acad. Sci. U.S.A. 2015,112 (50), E6844E6851.
(37) Yang, D.; Jang, W. D.; Lee, S. Y. Production of Carminic Acid
by Metabolically Engineered Escherichia Coli.J. Am. Chem. Soc. 2021,
143 (14), 53645377.
(38) Connors, N. C.; Bartel, P. L.; Strohl, W. R. Biosynthesis of
Anthracydines: Enzymic Conversion of Aklanonic Acid to Aklavinone
and -Rhodomycinone by Anthracycline-Producing Streptomycetes. J.
Gen. Microbiol. 1990,136 (9), 18871894.
(39) Hoshino, T.; Fujiwara, A. Microbial Conversion of Anthracy-
cline Antibiotics. II. Characterization of the Microbial Conversion
Products of Auramycinone by Streptomyces Coeruleorubidus ATCC
31276. J. Antibiot. 1983,36 (11), 14631467.
(40) Grocholski, T.; Dinis, P.; Niiranen, L.; Niemi, J.; Metsä-Ketelä,
M. Divergent Evolution of an Atypical S-Adenosyl-L-Methionine-
Dependent Monooxygenase Involved in Anthracycline Biosynthesis.
Proc. Natl. Acad. Sci. U.S.A. 2015,112 (32), 98669871.
(41) Beinker, P.; Lohkamp, B.; Peltonen, T.; Niemi, J.; Mäntsälä, P.;
Schneider, G. Crystal Structures of SnoaL2 and AclR: Two Putative
Hydroxylases in the Biosynthesis of Aromatic Polyketide Antibiotics.
J. Mol. Biol. 2006,359 (3), 728740.
(42) Siitonen, V.; Blauenburg, B.; Kallio, P.; Mäntsälä, P.; Metsä-
Ketelä, M. Discovery of a Two-Component Monooxygenase SnoaW/
SnoaL2 Involved in Nogalamycin Biosynthesis. Chem. Biol. 2012,19
(5), 638646.
(43) Fujiwara, A.; Tazoe, M.; Hoshino, T.; Sekine, Y.; Masuda, S.;
Nomura, S. New Anthracycline Antibiotics, 1-Hydroxyauramycins
and 1-Hydroxysulfurmycins. J. Antibiot. 1981,34 (7), 912915.
(44) Siitonen, V.; Claesson, M.; Patrikainen, P.; Aromaa, M.;
Mäntsälä, P.; Schneider, G.; Metsä-Ketelä, M. Identification of Late-
Stage Glycosylation Steps in the Biosynthetic Pathway of the
Anthracycline Nogalamycin. ChemBioChem 2012,13 (1), 120128.
(45) Hu, Y.; Zhang, Z.; Yin, Y.; Tang, G.-L. Directed Biosynthesis of
Iso-Aclacinomycins with Improved Anticancer Activity. Org. Lett.
2020,22, 150154.
(46) Matsuzawa, Y.; Oki, T.; Toshikazu, O.; Takeuchi, T. Structure-
Activity Relationships of Anthracyclines Relative to Cytotoxicity and
Effects on Macromolecular Synthesis in L1210 Leukemia Cells. J.
Antibiot. 1981,34 (12), 15961607.
(47) Kim, H. S.; Hong, Y. S.; Kim, Y. H.; Yoo, O. J.; Lee, J. J. New
Anthracycline Metabolites Produced by the Aklavinone 11-Hydrox-
ylase Gene in Streptomyces Galilaeus ATCC 31133. J. Antibiot. 1996,
49 (4), 355360.
(48) Lai, Y. H.; Chen, M. H.; Lin, S. Y.; Lin, S. Y.; Wong, Y. H.; Yu,
S. L.; Chen, H. W.; Yang, C. H.; Chang, G. C.; Chen, J. J.
Rhodomycin A, a Novel Src-Targeted Compound, Can Suppress
Lung Cancer Cell Progression via Modulating Src-Related Pathways.
Oncotarget 2015,6(28), 2625226265.
(49) Mao, Y.-y.; Cai, H.-c.; Shen, K.-n.; Chang, L.; Zhang, L.; Zhang,
Y.; Feng, J.; Wang, W.; Yang, C.; Zhu, T.-n.; Duan, M.-h.; Zhou, D.-
b.; Cao, X.-x.; Li, J. Benefit of High-Dose Idarubicin as Induction
Therapy in Acute Myeloid Leukemia: A Prospective Phase 2 Study.
Ann. Hematol. 2022,101 (4), 831836.
(50) Sambrook, J.; Russell, D. W. Molecular Cloning: A Laboratory
Manual; Cold Spring Harbor Laboratory Press: Cold Spring Harbor,
NY, 2001; p 999.
(51) Kallio, P.; Sultana, A.; Niemi, J.; Mäntsälä, P.; Schneider, G.
Crystal Structure of the Polyketide Cyclase AknH with Bound
Substrate and Product Analogue: Implications for Catalytic
Mechanism and Product Stereoselectivity. J. Mol. Biol. 2006,357
(1), 210220.
(52) Kieser, T.; Bibb, M. J.; Buttner, M. J.; Chater, K. F.; Hopwood,
D. A. Practical Streptomyces Genetics; John Innes Centre Ltd., 2000; p
529.
(53) Ylihonko, K.; Hakala, J.; Niemi, J.; Lundell, J.; Mantsala, P.
Isolation and Characterization of Aclacinomycin A-Non-Producing
Streptomyces Galilaeus (ATCC 31615) Mutants. Microbiology 1994,
140 (6), 13591365.
(54) Shaaban, K. A.; Wang, X.; Elshahawi, S. I.; Ponomareva, L. V.;
Sunkara, M.; Copley, G. C.; Hower, J. C.; Morris, A. J.; Kharel, M. K.;
Thorson, J. S.; Herbimycins, D.-F. Ansamycin Analogues from
Streptomyce s Sp. RM-715. J. Nat. Prod. 2013,76 (9), 16191626.
(55) Wang, X.; Shaaban, K. A.; Elshahawi, S. I.; Ponomareva, L. V.;
Sunkara, M.; Zhang, Y.; Copley, G. C.; Hower, J. C.; Morris, A. J.;
Kharel, M. K.; Thorson, J. S. Frenolicins CG, Pyranonaphthoqui-
nones from Streptomyces sp. RM-4-15. J. Nat. Prod. 2013,76, 1441
1447, DOI: 10.1021/np400231r.
(56) Savi, D. C.; Shaaban, K. A.; Wilke, F. M.; Gos, R.; Ponomareva,
L. V.; Thorson, J. S.; Glienke, C.; Rohr, J. Phaeophleospora Vochysiae
Savi & Glienke Sp. Nov. Isolated from Vochysia Divergens Found in
the Pantanal, Brazil, Produces Bioactive Secondary Metabolites. Sci.
Rep. 2018,8, No. 3122, DOI: 10.1038/s41598-018-21400-2.
(57) Shaaban, K. A.; Elshahawi, S. I.; Wang, X.; Horn, J.; Kharel, M.
K.; Leggas, M.; Thorson, J. S. Cytotoxic Indolocarbazoles from
Actinomadura Melliaura ATCC 39691. J. Nat. Prod. 2015,78 (7),
17231729.
ACS Synthetic Biology pubs.acs.org/synthbio Research Article
https://doi.org/10.1021/acssynbio.4c00043
ACS Synth. Biol. XXXX, XXX, XXXXXX
N
Article
An efficient and convenient pathway to synthesize diversified 3 or 4-oxo-alkyls substituted 1,2,4-triazine-3,5(2H, 4H)-diones has been reported via the atom- and step-economic cross-dehydrogenation coupling strategy with good functional group tolerance...
Article
Full-text available
Tetracenomycins and elloramycins are polyketide natural products produced by several actinomycetes that exhibit antibacterial and anticancer activities. They inhibit ribosomal translation by binding in the polypeptide exit channel of the large ribosomal subunit. The tetracenomycins and elloramycins are typified by a shared oxidatively modified linear decaketide core, yet they are distinguished by the extent of O-methylation and the presence of a 2',3',4'-tri-O-methyl-α-l-rhamnose appended at the 8-position of elloramycin. The transfer of the TDP-l-rhamnose donor to the 8-demethyl-tetracenomycin C aglycone acceptor is catalyzed by the promiscuous glycosyltransferase ElmGT. ElmGT exhibits remarkable flexibility toward transfer of many TDP-deoxysugar substrates to 8-demethyltetracenomycin C, including TDP-2,6-dideoxysugars, TDP-2,3,6-trideoxysugars, and methyl-branched deoxysugars in both d- and l-configurations. Previously, we developed an improved host, Streptomyces coelicolor M1146::cos16F4iE, which is a stable integrant harboring the required genes for 8-demethyltetracenomycin C biosynthesis and expression of ElmGT. In this work, we developed BioBricks gene cassettes for the metabolic engineering of deoxysugar biosynthesis in Streptomyces spp. As a proof of concept, we used the BioBricks expression platform to engineer biosynthesis for d-configured TDP-deoxysugars, including known compounds 8-O-d-glucosyl-tetracenomycin C, 8-O-d-olivosyl-tetracenomycin C, 8-O-d-mycarosyl-tetracenomycin C, and 8-O-d-digitoxosyl-tetracenomycin C. In addition, we generated four new tetracenomycins including one modified with a ketosugar, 8-O-4'-keto-d-digitoxosyl-tetracenomycin C, and three modified with 6-deoxysugars, including 8-O-d-fucosyl-tetracenomycin C, 8-O-d-allosyl-tetracenomycin C, and 8-O-d-quinovosyl-tetracenomycin C. Our work demonstrates the feasibility of BioBricks cloning, with the ability to recycle intermediate constructs, for the rapid assembly of diverse carbohydrate pathways and glycodiversification of a variety of natural products.
Article
Full-text available
Streptomyces soil bacteria produce hundreds of anthracycline anticancer agents with a relatively conserved set of genes. This diversity depends on the rapid evolution of biosynthetic enzymes to acquire novel functionalities. Previous work has identified S-adenosyl-l-methionine-dependent methyltransferase-like proteins that catalyze 4-O-methylation, 10-decarboxylation, or 10-hydroxylation, with additional differences in substrate specificities. Here we focused on four protein regions to generate chimeric enzymes using sequences from four distinct subfamilies to elucidate their influence in catalysis. Combined with structural studies we managed to depict factors that influence gain-of-hydroxylation, loss-of-methylation, and substrate selection. The engineering expanded the catalytic repertoire to include novel 9,10-elimination activity, and 4-O-methylation and 10-decarboxylation of unnatural substrates. The work provides an instructive account on how the rise of diversity of microbial natural products may occur through subtle changes in biosynthetic enzymes.
Article
Full-text available
Actinomycetes produce a variety of clinically indispensable molecules, such as antineoplastic anthracyclines. However, the actinomycetes are hindered in their further development as genetically engineered hosts for the synthesis of new anthracycline analogues due to their slow growth kinetics associated with their mycelial life cycle and the lack of a comprehensive genetic toolbox for combinatorial biosynthesis. In this report, we tackled both issues via the development of the BIOPOLYMER (BIOBricks POLYketide Metabolic EngineeRing) toolbox: a comprehensive synthetic biology toolbox consisting of engineered strains, promoters, vectors, and biosynthetic genes for the synthesis of anthracyclinones. An improved derivative of the production host Streptomyces coelicolor M1152 was created by deleting the matAB gene cluster that specifies extracellular poly-β-1,6-N-acetylglucosamine (PNAG). This resulted in a loss of mycelial aggregation, with improved biomass accumulation and anthracyclinone production. We then leveraged BIOPOLYMER to engineer four distinct anthracyclinone pathways, identifying optimal combinations of promoters, genes, and vectors to produce aklavinone, 9-epi-aklavinone, auramycinone, and nogalamycinone at titers between 15-20 mg/L. Optimization of nogalamycinone production strains resulted in titers of 103 mg/L. We structurally characterized six anthracyclinone products from fermentations, including new compounds 9,10-seco-7-deoxy-nogalamycinone and 4-O-β-d-glucosyl-nogalamycinone. Lastly, we tested the antiproliferative activity of the anthracyclinones in a mammalian cancer cell viability assay, in which nogalamycinone, auramycinone, and aklavinone exhibited moderate cytotoxicity against several cancer cell lines. We envision that BIOPOLYMER will serve as a foundational platform technology for the synthesis of designer anthracycline analogues.
Article
Full-text available
Idarubicin 12 mg/m2 has been recommended as a standard induction therapy for acute myeloid leukemia (AML). It is unknown whether a higher dose of idarubicin can improve the remission rate. This phase 2 prospective single-arm study enrolled 45 adults with newly diagnosed AML between September 2019 and May 2021 (NCT 04,069,208). Induction therapy included administration of idarubicin 14 mg/m2 for 3 days and cytarabine 100 mg/m2 every 12 h subcutaneously for 7 days. The primary endpoint was the composite complete response rate (complete response (CR) plus complete response with incomplete blood count recovery (CRi)). The median age was 45 years (range 14–60 years). Forty (88.9%) patients had CR or CRi, including 39 patients with CR and 1 patient with CRi after one course of induction therapy. The median times to recovery of absolute neutrophil and platelet counts were 21 days. Only 1 patient died of intracranial hemorrhage during induction therapy. After a median follow-up of 14 months (range 3.5–24 months), the estimated 18-month overall survival and disease-free survival (DFS) were 66.9% and 57.5%, respectively. In conclusion, idarubicin 14 mg/m2 plus cytarabine was a safe and efficient intensive regimen for younger and fit patients with newly diagnosed AML.
Article
Full-text available
Background: Mithramycin is an anti-tumor compound of the aureolic acid family produced by Streptomyces argillaceus. Its biosynthesis gene cluster has been cloned and characterized, and several new analogs with improved pharmacological properties have been generated through combinatorial biosynthesis. To further study these compounds as potential new anticancer drugs requires their production yields to be improved significantly. The biosynthesis of mithramycin proceeds through the formation of the key intermediate 4-demethyl-premithramycinone. Extensive studies have characterized the biosynthesis pathway from this intermediate to mithramycin. However, the biosynthesis pathway for 4-demethyl-premithramycinone remains unclear. Results: Expression of cosmid cosAR7, containing a set of mithramycin biosynthesis genes, in Streptomyces albus resulted in the production of 4-demethyl-premithramycinone, delimiting genes required for its biosynthesis. Inactivation of mtmL, encoding an ATP-dependent acyl-CoA ligase, led to the accumulation of the tricyclic intermediate 2-hydroxy-nogalonic acid, proving its essential role in the formation of the fourth ring of 4-demethyl-premithramycinone. Expression of different sets of mithramycin biosynthesis genes as cassettes in S. albus and analysis of the resulting metabolites, allowed the reconstitution of the biosynthesis pathway for 4-demethyl-premithramycinone, assigning gene functions and establishing the order of biosynthetic steps. Conclusions: We established the biosynthesis pathway for 4-demethyl-premithramycinone, and identified the minimal set of genes required for its assembly. We propose that the biosynthesis starts with the formation of a linear decaketide by the minimal polyketide synthase MtmPKS. Then, the cyclase/aromatase MtmQ catalyzes the cyclization of the first ring (C7-C12), followed by formation of the second and third rings (C5-C14; C3-C16) catalyzed by the cyclase MtmY. Formation of the fourth ring (C1-C18) requires MtmL and MtmX. Finally, further oxygenation and reduction is catalyzed by MtmOII and MtmTI/MtmTII respectively, to generate the final stable tetracyclic intermediate 4-demethyl-premithramycinone. Understanding the biosynthesis of this compound affords enhanced possibilities to generate new mithramycin analogs and improve their production titers for bioactivity investigation.
Article
The anthracyclines are a family of natural products isolated from soil bacteria with over 2000 chemical representatives. Since their discovery seventy years ago by Waksman and co-workers, anthracyclines have become one of the best-characterized anticancer chemotherapies in clinical use. The anthracyclines exhibit broad-spectrum antineoplastic activity for the treatment of a variety of solid and liquid tumors, however, their clinical use is limited by their dose-limiting cardiotoxicity. In this review article, we discuss the toxicity of the anthracyclines on several organ systems, including new insights into doxorubicin-induced cardiotoxicity. In addition, we discuss new medicinal chemistry developments in the biosynthesis of new anthracycline analogs and the synthesis of new anthracycline analogs with diminished cardiotoxicity. Lastly, we review new studies that describe the repurposing of the anthracyclines, or "upcycling" of the anthracyclines, as anti-infective agents, or drugs for niche indications. Altogether, the anthracyclines remain a mainstay in the clinic with a potential new "lease on life" due to deeper insight into the mechanism underlying their cardiotoxicity and new developments into potential new clinical indications for their use. Keywords: Anthracycline, chemotherapy, toxicology, medicinal chemistry, biosynthesis.
Article
Covering: January 1995 to June 2021Anthracyclines are glycosylated microbial natural products that harbour potent antiproliferative activities. Doxorubicin has been widely used as an anticancer agent in the clinic for several decades, but its use is restricted due to severe side-effects such as cardiotoxicity. Recent studies into the mode-of-action of anthracyclines have revealed that effective cardiotoxicity-free anthracyclines can be generated by focusing on histone eviction activity, instead of canonical topoisomerase II poisoning leading to double strand breaks in DNA. These developments have coincided with an increased understanding of the biosynthesis of anthracyclines, which has allowed generation of novel compound libraries by metabolic engineering and combinatorial biosynthesis. Coupled to the continued discovery of new congeners from rare Actinobacteria, a better understanding of the biology of Streptomyces and improved production methodologies, the stage is set for the development of novel anthracyclines that can finally surpass doxorubicin at the forefront of cancer chemotherapy.
Article
Background/Goal/Aim The tetracenomycins are aromatic anticancer polyketides that inhibit peptide translation via binding to the large ribosomal subunit. Here, we expressed the elloramycin biosynthetic gene cluster in the heterologous host Streptomyces coelicolor M1146 to facilitate the downstream production of tetracenomycin analogs. Main Methods and Major Results We developed a BioBricks® genetic toolbox of genetic parts for substrate precursor engineering in S. coelicolor M1146::cos16F4iE. We cloned a series of integrating vectors based on the VWB, TG1, and SV1 integrase systems to interrogate gene expression in the chromosome. We genetically engineered three separate genetic constructs to modulate tetracenomycin biosynthesis: 1) the vhb hemoglobin from obligate aerobe Vitreoscilla stercoraria to improve oxygen utilization; (2) the accA2BE acetyl-CoA carboxylase to enhance condensation of malonyl-CoA; (3) lastly, the sco6196 acyltransferase, which is a “metabolic regulatory switch” responsible for mobilizing triacylglycerols to β-oxidation machinery for acetyl-CoA. In addition, we engineered the tcmO 8-O-methyltransferase and newly identified tcmD 12-O-methyltransferase from Amycolatopsis sp. A23 to generate tetracenomycins C and X. We also co-expressed the tcmO methyltransferase with oxygenase urdE to generate the analog 6-hydroxy-tetracenomycin C. Conclusions and Implications Altogether, this system is compatible with the BioBricks® [RFC 10] cloning standard for the co-expression of multiple gene sets for metabolic engineering of Streptomyces coelicolor M1146::cos16F4iE. This production platform improves access to potent analogs, such as tetracenomycin X, and sets the stage for the production of new tetracenomycins via combinatorial biosynthesis. This article is protected by copyright. All rights reserved
Article
Carminic acid is an aromatic polyketide found in scale insects (i.e., Dactylopius coccus) and is a widely used natural red colorant. It has long been produced by the cumbersome farming of insects followed by multistep purification processes. Thus, there has been much interest in producing carminic acid by the fermentation of engineered bacteria. Here we report the complete biosynthesis of carminic acid from glucose in engineered Escherichia coli. We first optimized the type II polyketide synthase machinery from Photorhabdus luminescens, enabling a high-level production of flavokermesic acid upon coexpression of the cyclases ZhuI and ZhuJ from Streptomyces sp. R1128. To discover the enzymes responsible for the remaining two reactions (hydroxylation and C-glucosylation), biochemical reaction analyses were performed by testing enzyme candidates reported to perform similar reactions. The two identified enzymes, aklavinone 12-hydroxylase (DnrF) from Streptomyces peucetius and C-glucosyltransferase (GtCGT) from Gentiana triflora, could successfully perform hydroxylation and C-glucosylation of flavokermesic acid, respectively. Then, homology modeling and docking simulations were performed to enhance the activities of these two enzymes, leading to the generation of beneficial mutants with 2-5-fold enhanced conversion efficiencies. In addition, the GtCGT mutant was found to be a generally applicable C-glucosyltransferase in E. coli, as was showcased by the successful production of aloesin found in Aloe vera. Simple metabolic engineering followed by fed-batch fermentation resulted in 0.63 ± 0.02 mg/L of carminic acid production from glucose. The strategies described here will be useful for the design and construction of biosynthetic pathways involving unknown enzymes and consequently the production of diverse industrially important natural products.
Article
The anthracyclines are structurally diverse anticancer natural products that bind to DNA and poison the topoisomerase II - DNA complex in cancer cells. Rational modifications in the deoxysugar functionality are especially advantageous for synthesizing drugs with improved potency. Combinatorial biosynthesis of glycosyltransferases and deoxysugar synthesis enzymes is indispensable for the generation of glycodiversified anthracyclines. This Synopsis considers recent advances in glycosyltransferase structural biology and site-directed mutagenesis, pathway engineering, and deoxysugar combinatorial biosynthesis with a focus on generation of "new-to-nature" anthracycline analogs.