ArticlePDF Available

Poisoning of Mitochondrial Topoisomerase I by Lamellarin D

Authors:

Abstract and Figures

Lamellarin D (Lam-D) is a hexacyclic pyrole alkaloid isolated from marine invertebrates whose biological properties have been attributed to mitochondrial targeting. Mitochondria contain their own DNA (mtDNA), and the only specific mitochondrial topoisomerase in vertebrates is mitochondrial topoisomerase I (Top1mt). Here, we show that Top1mt is a direct mitochondrial target of Lam-D. In vitro Lam-D traps Top1mt and induces Top1mt cleavage complexes (Top1mtcc). Using single molecule analyses we also show that Lam-D slows down supercoil relaxation of Top1mt and strongly inhibits Top1mt religation in contrast to the inefficacy of camptothecin on Top1mt. In living cells, we show that Lam-D accumulates rapidly inside mitochondria, induces cellular Top1mtcc and leads to mtDNA damage. This study provides evidence that Top1mt is a direct mitochondrial target of Lam-D and suggests that developing Top1mt inhibitors represent a novel strategy for targeting mitochondrial DNA.
Content may be subject to copyright.
1521-0111/86/2/193199$25.00 http://dx.doi.org/10.1124/mol.114.092833
MOLECULAR PHARMACOLOGY Mol Pharmacol 86:193199, August 2014
U.S. Government work not protected by U.S. copyright
Poisoning of Mitochondrial Topoisomerase I by Lamellarin D
Salim Khiati, Yeonee Seol, Keli Agama, Ilaria Dalla Rosa, Surbhi Agrawal, Katherine Fesen,
Hongliang Zhang, Keir C. Neuman, and Yves Pommier
Developmental Therapeutics Branch and Laboratory of Molecular Pharmacology, Center for Cancer Research, National Cancer
Institute (S.K., K.A., I.D.R., S.A., K.F., H.Z., Y.P.) and Laboratory of Molecular Biophysics, National Heart, Lung, and Blood
Institute (Y.S., K.C.N.), National Institutes of Health, Bethesda, Maryland
Received March 21, 2014; accepted May 28, 2014
ABSTRACT
Lamellarin D (Lam-D) is a hexacyclic pyrole alkaloid isolated
from marine invertebrates, whose biologic properties have been
attributed to mitochondrial targeting. Mitochondria contain their
own DNA (mtDNA), and the only specific mitochondrial topo-
isomerase in vertebrates is mitochondrial topoisomerase I
(Top1mt). Here, we show that Top1mt is a direct mitochondrial
target of Lam-D. In vitro Lam-D traps Top1mt and induces Top1mt
cleavage complexes (Top1mtcc). Using single-molecule analyses,
we also show that Lam-D slows down supercoil relaxation of
Top1mt and strongly inhibits Top1mt religation in contrast to the
inefficacy of camptothecin on Top1mt. In living cells, we show that
Lam-D accumulates rapidly inside mitochondria, induces cellular
Top1mtcc, and leads to mtDNA damage. This study provides
evidence that Top1mt is a direct mitochondrial target of Lam-D
and suggests that developing Top1mt inhibitors represents
a novel strategy for targeting mitochondrial DNA.
Introduction
The contribution of mitochondria to cellular functions
extends beyond their critical bioenergetics role in the form
of ATP production. Indeed, they also control cell death by
apoptosis, generate critical metabolites for macromolecular
synthesis, and regulate divalent metal pools. Mitochondrial
dysfunctions have been linked to a wide range of neuro-
degenerative and metabolic diseases, cancers, and aging
(Greaves et al., 2012). Mitochondria are the only cellular
organelle containing metabolically active DNA, other than the
nucleus (Chinnery and Hudson, 2013). The mammalian
mitochondrial genome (mtDNA) consists of circular DNA
molecules with 16,569 base pairs encoding 13 essential
polypeptides for complexes I, III, and IV of the mitochondrial
oxidative phosphorylation chain. It also encodes indispens-
able components for mitochondrial translation: 22 tRNAs and
two ribosomal RNAs (12S and 16S rRNAs for mitoribosomes).
The only noncoding region of mtDNA is a small regulatory
region (5% of the genome) with critical elements for
transcription (promoters for both strands) and replication
(origin of replication). Because mtDNA is essential for
mitochondria (Chinnery and Hudson, 2013), its targeting
with drugs offers potential for new therapeutics, as well as
molecular tools to study mtDNA homeostasis and their effects
on mitochondria (Rowe et al., 2001; Neuzil et al., 2013;
Olszewska and Szewczyk, 2013).
Given that mtDNA consists of relatively small DNA
circles with bidirectional transcription and replication, and
organization in nucleoids anchored to the mitochondrial
inner membrane, it is logical to assume that mitochondrial
topoisomerases are critical to ensure mtDNA replication
and transcription. To date, three topoisomerases have been
identified in vertebrate mitochondria: Top1mt (Zhang
et al., 2001), Top2b(Low et al., 2003), and Top3a(Wang
et al., 2002). Unlike TOP2Β(Low et al., 2003) and TOP3Α,
TOP1mt is the only topoisomerase gene coding for a single
polypeptide solely targeted to mitochondria (Zhang et al.,
2001; Chinnery and Hudson, 2013). In addition to studying
Top1mts roles in regulating mtDNA replication (Zhang
and Pommier, 2008), transcription (Sobek et al., 2013), and
mtDNA integrity (Medikayala et al., 2011), it has become
possible to examine Top1mt functions genetically. Murine
embryonic fibroblasts generated from TOP1mt knockout
mice exhibit mitochondrial defects with marked increase in
reactive oxygen species production, calcium signaling,
hyperpolarization of mitochondrial membranes, and in-
creased mitophagy (Douarre et al., 2012).
Topoisomerases relax DNA supercoils by breaking the DNA
backbone and forming transient catalytic enzyme-linked DNA
breaks, which are referred to as cleavage complexes (Pommier
et al., 2010; Pommier and Marchand, 2012; Pommier, 2013). The
DNA-damaging effects and therapeutic activity of topoisomerase
inhibitors result from the trapping of topoisomerase cleavage
complexes by the binding of the inhibitors at the enzyme-DNA
interface rather than by catalytic inactivation of the enzymes
(reviewed in Nitiss and Wang, 1988; Pommier et al., 2010, 2013).
This work was supported by the Intramural Research Program of the
National Institutes of Health National Cancer Institute, Center of Cancer
Research [Z01 BC006161] and the National Heart, Lung, and Blood Institute.
dx.doi.org/10.1124/mol.114.092833.
ABBREVIATIONS: CPT, Camptothecin; ICE, immuno-complex of enzyme; Lam-D, lamellarin D; mtDNA, mitochondrial DNA; NCI_H23,
adenocarcinoma, human nonsmall cell lung cancer; SK-MEL-5, human skin melanoma cell line; Top1, topoisomerase I; Top1cc, Top1 cleavage
complexes; Top1mt, mitochondrial topoisomerase I; Top1mtcc, mitochondrial topoisomerase I cleavage complexes; Tw, twist density.
193
at ASPET Journals on February 21, 2015molpharm.aspetjournals.orgDownloaded from
Camptothecin (CPT) and its clinical derivatives (topotecan
and irinotecan) kill cancer cells by selectively trapping
(poisoning) nuclear topoisomerase I (Top1) (Hsiang et al.,
1985; Nitiss and Wang, 1988; Pommier et al., 2010;
Pommier and Marchand, 2012). Although DNA topoisom-
erases are among the most effective targets for antican-
cer and antibacterial agents (Pommier et al., 2010), no
drug has been shown to target mitochondrial topoisom-
erases, and camptothecins are inefficient Top1mt inhib-
itors, as they are rapidly inactivated in the mitochondrial
pH environment (Burke and Mi, 1994; Zhang et al., 2001;
Pommier et al., 2010) and require high drug concentrations to
target Top1mt (Zhang et al., 2001; Seol et al., 2012), which
otherwise extensively damage the nuclear genome (Hsiang et al.,
1985; Covey et al., 1989).
The purpose of the present study was to determine whether
lamellarin D (Lam-D; Fig. 1A) could poison Top1mt in
biochemical and cellular systems. Lamellarins (Bailly, 2004;
Pla et al., 2008) are hexacyclic pyrole alkaloids originally
isolated from marine invertebrates (genus Lamellaria); they
can also be obtained by total chemical synthesis (Li et al.,
2011). Lam-D is among the most cytotoxic molecules in the
series (Bailly, 2004). Its proapoptotic activity has been
associated with the generation of DNA breaks resulting from
the trapping of nuclear topoisomerase I [Top1-DNA covalent
complexes (Top1cc) (Facompre et al., 2003)]. However, Lam-D
has also been shown to directly target mitochondria (Kluza
et al., 2006; Ballot et al., 2009; Ballot et al., 2010), suggesting
that it may poison Top1mt and could act as a potential drug
targeting mtDNA by trapping Top1mt. In this study we
investigated the effects of Lam-D on Top1mt using in vitro
and in vivo DNA cleavage assays in addition to single-
molecule supercoil relaxation measurements. We found that
Lam-D indeed inhibits Top1mt by trapping the cleavage
complex efficiently, in striking contrast to the inefficacy
of CPT. More importantly, our study reveals that Lam-D
accumulates preferentially inside the mitochondria and ef-
ficiently induces Top1mtcc, resulting in mtDNA damage in
treated cells.
Materials and Methods
Mitochondrial Topoisomerase I-Mediated DNA Cleavage
Reactions. Human recombinant Top1 was purified from baculovirus
as previously described (Pourquier et al., 1999; Seol et al., 2012), and
Top1mt was expressed and purified in the same manner. DNA
Fig. 1. Lamellarin D induces Top1mt-mtDNA
cleavage complexes in vitro. (A) Chemical structure
of Lam-D. (B) Comparison of Top1mt cleavage
complex sites induced by Lam-D and CPT in a
39-end-labeled 117-bp oligonucleotide rich in Top1
sites. Lane 1: DNA alone; Lane 2: Top1mt alone;
Lane 3: Top1mt + CPT 1 mM; Lanes 46: Top1mt +
0.1, 1, and 10 mM Lam-D, respectively. Numbers
and arrows on the left indicate cleavage site
positions (see Materials and Methods). (C) Lam-D
inhibits Top1mtcc religation. The same oligonucle-
otide was reacted with Top1mt in the presence of
1mM CPT or Lam-D at 25°C for 20 minutes. DNA
cleavage was reversed by adding 0.35 M NaCl and
monitored over time. Numbers and arrows on the
left indicate cleavage site positions.
194 Khiati et al.
at ASPET Journals on February 21, 2015molpharm.aspetjournals.orgDownloaded from
cleavage reactions were prepared as previously reported with the
exception of the DNA substrate (Dexheimer and Pommier, 2008).
Briefly, a 117-bp DNA oligonucleotide contains a single 59-cytosine
overhang, which was 39end-labeled by fill-in reaction with [
32
P]-dGTP
in NEBuffer 2 (New England BioLabs Inc., Ipswich, MA) with 0.5
units of DNA polymerase I (New England BioLabs).
For reversal experiments, the addition of SDS was preceded by the
addition of NaCl (final concentration of 0.35 M at 25°C for the
indicated times). Aliquots of each reaction mixture were subjected to
20% denaturing PAGE. Gels were dried and visualized by using
a phosphorimager and ImageQuant software (Molecular Dynamics/
GE Healthcare Life Sciences, Pittsburgh, PA). For simplicity,
cleavage sites were numbered as previously described in the 161-bp
fragment (Pourquier et al., 1999).
Single-Molecule Measurement and Analysis. Generation of
23-kb coilable DNA and sample preparation were done as previously
described (Seol et al., 2012). Measurements of supercoil relaxation by
Top1mt (50500 pM) were performed in topoisomerase buffer [10 mM
Tris, pH 8, 50 mM KCl, 10 mM MgCl
2
, 0.3% w/v bovine serum albumin
(BSA), 0.04% Tween-20, 0.1 mM EDTA, and 5 mM dithiothreitol
(DTT)] with varying Lam-D concentrations (110 mM) and DNA twist
density (0.008, 0.008, 0.016, and 0.032). DNA twist density (Tw)is
controlled via the force applied on the magnetic bead (Seol et al.,
2012). The overall Top1mt activity was measured by tracking the
height of the tethered bead at 100 Hz (Seol et al., 2012). The rates of
supercoil relaxation by Top1mt were obtained by analyzing the DNA
extension change using custom-written software (Seol et al., 2012).
The slow relaxation-rate probability was determined by calculating
the fraction of rates that were less than 30% of the mean rate obtained
in the absence of Lam-D for a given DNA twist density. The religation
inhibition probability was quantified from the ratio of futile to
productive supercoiling attempts, i.e., the fraction of events for which
there was no change in DNA linking-number difference (DL
k
) when
the magnets were rotated.
Confocal Microscopy in Live Cells. Human skin melanoma
cells (SK-MEL-5) growing on chamber slides (Nalge Nunc International/
Thermo Scientific, Rochester, NY) were incubated in medium con-
taining 100 nM MitoTracker Red 580 (Molecular Probe, Eugene,
Oregon) at 37°C for 1 hour. Lam-D (1 mM) was added and fluorescent
signals were accessed after 530 minutes under confocal microscope
(Zeiss 710). Images were collected and processed using the Zeiss
AIM software and fluorescence quantification was performed using
ImageJ software.
Detection of Top1 and Top1mt Cleavage Complexes by
Immuno Complex of Enzyme Bioassay. Cells (5 10
5
per well in
six-well plate) were grown for 12 days (7080% confluence). After
treatments, cells were lysed by adding to each culture well 0.6 ml of
lysis buffer (6 M Guanidinium Thiocyanate, 10 mM Tris-HCL, pH 6.5,
20 mM EDTA, 4% Triton 100, 1% sarosyl, and 1% DTT). Lysates
were transferred to Eppendorf tubes and mixed with 0.4 ml of 100%
ethanol. After incubation at 20°C for 5 minutes, tubes were
centrifuged for 15 minutes. Supernatants were discarded and pellets
washed two times with 0.8 ml of 100% ethanol. Pellets were dissolved
in 0.2 ml 8 mM NaOH (freshly made) and sonicated for 1020 seconds
at 20% power. For immunodetection, equal volumes and concen-
trations of isolated DNA were transferred to 96-well plates and
serially diluted with 8 mM NaOH. At least 100 ml of DNA solution
were applied to polyvinylidene difluoride membranes with a slot blot
apparatus. After 1 hour blocking with 5% milk in PBST (phosphate-
buffered saline, Tween-20 0.1%), membranes were incubated over-
night with Top1mt antibody (Douarre et al., 2012) or (nuclear) Top1
antibody (#556597; BD Biosciences, San Jose, CA) (Antony et al.,
2007). After three washes in PBST, membranes were incubated
with horseradish peroxidaseconjugated goat anti-mouse (1:5000
dilution) antibody (Amersham Biosciences, Piscataway, NJ) for
1 hour and washed three times. Immunoblots were detected using
enhanced chemiluminescence detection kit (Thermo Scientific,
Rockford, IL).
Quantification of mtDNA Damage. Mitochondrial DNA dam-
age was quantified by long-range polymerase chain reaction (PCR)
(Hunter et al., 2010) from total DNA isolated from whole cell after
treatment. A 10-Kb fragment and a shorter region of mtDNA (117 pb)
were amplified. PCR reactions were limited to 14 and 16 cycles, to
ensure that amplification process was still in the exponential phase.
The damage index is determined by the ratio P/Q of long-range PCR
product (P) by the short-range PCR product (Q). Primer sequences
used for mtDNA analysis are:
Results
Induction of Top1mt Cleavage Complexes by Lamel-
larin D. To determine whether Lam-D could target Top1mt,
we tested cleavage complex formation with recombinant
Top1mt using a 117-base-pair
32
P-39-end-labeled DNA oligo-
nucleotide substrate (Zhang et al., 2001; Antony et al., 2007).
Figure 1 demonstrates the induction of Top1mt cleavage
complexes by Lam-D. The cleavage sites were common to
those trapped by CPT. Yet, Lam-D showed differential
intensity at some sites (such as sites 70 and 106 in Fig. 1),
which is most probably related to the structural difference
between the polycyclic ring drug structures that intercalate at
the break sites (Stewart et al., 1998; Ioanoviciu et al., 2005).
These results demonstrate that Lam-D can effectively target
Top1mt cleavage complexes at submicromolar concentrations.
Lamellarin D Inhibits Top1mt Cleavage Complex
Religation and Supercoil Relaxation. To determine
whether Lam-D enhances Top1mt cleavage complexes by
inhibiting DNA religation, we first measured the reversal to
Top1mt cleavage complexes after addition of salt, which shifts
the nicking-religation activity of Top1 toward religation
(Tanizawa et al., 1994). To do so, we followed the reversal
kinetics of site 92 (Fig. 1C), using CPT as positive control as it
is known to enhance Top1 cleavage complexes by religation
inhibition because of its binding at the Top1-DNA interface
(Hsiang et al., 1985; Tanizawa et al., 1994; Stewart et al.,
1998; Ioanoviciu et al., 2005). The religation kinetics of site 92
was similar for Lam-D and CPT indicating that Lam-D traps
Top1mt and inhibits religation.
In line with this, we performed single-molecule measure-
ments of Top1mt to test the effect of Lam-D on supercoil
relaxation activity (Fig. 2A). The single-molecule supercoil
relaxation assay has proven to be a useful tool to test the
efficacy of topoisomerase inhibitors, since individual steps in
the topoisomerase reaction cycle and changes in these steps
upon inhibition by topoisomerase inhibitors such as CPT can
be monitored in detail (Koster et al., 2007; Seol et al., 2012).
In our assay, a 23-kb DNA tether was formed by attaching
one end of the DNA to the surface of a flow cell via multiple
digoxigenin-antidigoxigenin linkages and the other end to a
1-mm magnetic particle via multiple biotin-streptavidin link-
ages. By rotating small magnets above the flow cell, the DNA
can be supercoiled (change in linking number: DL
k
), resulting
in a decrease in the extension as plectonemes form in the DNA
(Fig. 2A). Supercoil relaxation by Top1mt was observed
through the increase in DNA extension. Relaxation rates
Long range PCR: 59-TCTAAGCCTCCTTATTCGAGCCGA-39and
59- TTTCATCATGCGGAGATGTTGGATGG -39
Short range PCR: 59-AAGTCACCCTAGCCATCATTCTAC-39and
59- GCAGGAGTAATCAGAGGTGTTCTT-39
Poisoning of Mitochondrial Topoisomerase I by Lamellarin D 195
at ASPET Journals on February 21, 2015molpharm.aspetjournals.orgDownloaded from
were quantified by a linear fit over the region where DNA
extension increases.
To determine the effects of Lam-D on Top1mt, we performed
DNA supercoil relaxation assays with Top1mt under varying
concentrations of Lam-D (010 mM) and Tw (0.008, 0.008,
0.016, and 0.032). Varying Tw allows us to investigate the effect
of Lam-D for different topological states potentially present in
vivo. We observed two aspects of Lam-D effect on Top1mt
activity: a significant decrease in supercoil relaxation rate (Fig.
2B), which was identified as one of the signature effects of CPT
on nuclear topoisomerase (Koster et al., 2007), and religation
inhibition, i.e., events for which the DNA could not be super-
coiled upon rotating the magnets (Fig. 2B).
We quantified the effect on Top1mt attributable to Lam-D by
calculating the probabilities of religation inhibition and slow-
relaxation rate (see Materials and Methods).AsshowninFig.
2C, the religation inhibition probability generally increased as
DNA twist density increased, whereas the probability of slow
relaxation was more pronounced at lower twist density, in-
dicating that Tw modulates the effect of Lam-D on Top1mt
(Gentry et al., 2011). Surprisingly, the degree of Top1mt
inhibition at 1-mM Lam-D was similar to that higher concen-
trations, suggesting robust inhibition efficacy of Lam-D in
contrast to the poor inhibition efficiency of CPT (Seol et al.,
2012). Together, the gel-based religation assays (Fig. 1C) and
the single-molecule measurements (Fig. 2) demonstrate that
Lam-D interferes with Top1mt activity both by inhibiting
the religation of cleavage complexes and impeding the DNA
unwinding activities of Top1mt over a broad range of DNA
twist density.
Lamellarin D Accumulates Rapidly in Mitochondria
and Poisons Top1mt in Cells. To determine whether Lam-D
can directly target Top1mt in cells, we took advantage of the
fact that Lam-D is intrinsically fluorescent. This enabled us to
investigate the intracellular distribution of Lam-D in live cells
by confocal microscopy. After Lam-D treatment of SK-MEL-5
melanoma cells, which we chose for this study because they
possess large mitochondria, we observed prominent Lam-D
cytoplasmic staining coinciding with mitochondria (Fig. 3).
Localization of Lam-D to mitochondria was verified by co-
localization with the mitochondria-specific dye MitoTracker
(Fig. 3A, top panels). As shown in the magnified image
(bottompanel,Fig.3A),andinthe correlation between the
pixel intensities along a line in the two channels (Fig. 3B),
the fluorescent intensity distribution of Lam-D coincides with
that of mitochondria. These results demonstrate that Lam-D
penetrates and accumulates preferentially in mitochondria
within 530 minutes of drug exposure.
Next, we tested the trapping of Top1mt cleavage complexes
in Lam-D-treated cells using the ICE assay (immuno-complex
of enzyme assay). This technique allows the detection of
topoisomerases covalently linked to DNA by immunoblotting
after isolation of cellular DNA (Subramanian et al., 1995;
Takebayashi et al., 1999; Antony et al., 2007). After Lam-D
treatment, Top1mt signals were observed in both SK-MEL-5
melanoma cells and NCI_H23 lung cancer cells (Fig. 4, A and B,
Fig. 2. Lamellarin D inhibits Top1mt cleavage-complex religation and
hinders supercoil relaxation as determined by single-molecule analysis. (A)
Top panel: Schematic of the single-molecule assay with the 23-kb DNA tether
(blue), 1-mm magnetic particle (brown sphere), and recombinant Top1mt (red).
Bottom panel: example trace of Top1mt supercoil relaxation, which is observed
through an increase in DNA extension (blue arrow). Brown lines represent the
mechanical coiling of DNA. The resulting extension change is shown by dashed
lines. Relaxation rates were quantified by a linear fit over the region where
DNA extension increases (blue dashed lines). (B) DNA supercoil relaxation by
Top1mt in the absence (top panel) or presence of Lam-D at 1 mM(twomiddle
panels) and 10 mM (bottom panel). Lam-D interaction with Top1mt resulted in
periods of slow supercoil relaxation (red dashed lines) and periods of religation
inhibition (indicated by red arrows). (C) Quantification of religation (top panel)
and slow relaxation probabilities (bottom panel) under varying conditions of
DNA supercoil twist density and Lam-D concentration.
196 Khiati et al.
at ASPET Journals on February 21, 2015molpharm.aspetjournals.orgDownloaded from
respectively). The induction of Top1mt-mtDNA complexes
by Lam-D was concentration- and time-dependent; time-course
analysis showed that Lam-D induced Top1mt-mtDNA com-
plexes within 1 hour (Fig. 4C), which is consistent with the
rapid accumulation of Lam-D in mitochondria (Fig. 3 and
above). We also tested whether trapping of Top1mt-mtDNA
cleavage complexes by Lam-D damaged mtDNA using long-
range PCR, a well-established method to evaluate mtDNA
damage (Das et al., 2010). Accordingly, we observed a Lam-D
concentration-dependent reduction in long-range PCR product
(Fig. 4, D and E), confirming the damaging effect of Lam-D on
mtDNA.
Because Lam-D has previously been shown to also trap
nuclear Top1 (Facompre et al., 2003), we compared the
relative activity of Lam-D on the mitochondrial versus
nuclear genome. For this, we took advantage of the fact that
our Top1mt antibodies do not cross-react with nuclear Top1,
and that the monoclonal C21 Top1 antibody does not cross-
react against Top1mt (Fig. 4F). Figure 4, G and H, shows that
Lam-D is more potent at trapping Top1mt-mtDNA cleavage
complexes than camptothecin (Fig. 4G). On the other hand,
CPT is more potent at trapping nuclear Top1 than Lam-D
(Fig. 4H), which is consistent with lower accumulation of
Lam-D in the cell nucleus than in mitochondria (see Fig. 3 and
above).
Discussion
Cells typically contain hundreds to thousands of copies of
mtDNA. Replication of mtDNA is independent of nuclear
DNA synthesis and occurs throughout all phases of the
cell cycle including in postmitotic cells (Clayton, 1991). DNA
polymerase gamma and mitochondrial transcription factor A
are two essential proteins for replication and maintenance
of mammalian mtDNA. Some drugs have been reported to
deplete mitochondrial DNA by interfering with mitochon-
drial DNA polymerase gamma activity (Rowe et al., 2001;
Neuzil et al., 2013). Top1mt is also important for mitochon-
drial homeostasis (Douarre et al., 2012), mtDNA replication
(Zhang and Pommier, 2008), mtDNA transcription (Sobek
et al., 2013), and mtDNA integrity (Medikayala et al.,
2011). Yet, there are no drugs known to specifically target
Top1mt.
In this study, we provide evidence that Lam-D targets and
efficiently inhibits Top1mt by trapping the normally transient
Top1mt-DNA cleavage complexes. Trapping of Top1mt cleav-
age complexes was observed in both biochemical (Figs. 1 and 2)
and cellular assays (Fig. 4). Furthermore, confocal imaging dem-
onstrates that Lam-D accumulates rapidly (530 minutes) in
mitochondria (Fig. 3), consistent with Lam-D-induced mito-
chondrial alterations (Kluza et al., 2006; Ballot et al., 2009;
Ballot et al., 2010). In particular, Ballot et al. described a
Lam-D-induced mitochondrial cascade involving: inhibition of
complex III, decreased zeta potential, swelling of mitochondrial
matrix, and apoptosis. Our work reveals that Top1mt is a new
direct mitochondrial target of Lam-D.
The general mechanism of inhibition of Top1mt by Lam-D
appears to be similar to that of CPT. Yet, the detailed
molecularinteractionsappeartodiffersomewhatasthe
sequence-dependent cleavage patterns between CPT and
Lam-D show differences (Fig. 1), and Lam-D is a more robust
inhibitor of Top1mt compared with CPT (Fig. 2) (Seol et al.,
2012). Although a significant decrease in supercoil relaxation
rate was identified as a signature aspect of CPT inhibition of
nuclear Top1 (Koster et al., 2007; Seol et al., 2012), our study
reveals that religation inhibition by Lam-D is not absolutely
associated with a decrease in supercoil relaxation rate (Fig. 2).
This implies that DNA damage attributable to topoisomerase
I inhibition may not always be associated with hindering
positive supercoil relaxation and subsequent accumulation of
positive supercoils, as previously suggested (Koster et al.,
2007; Ray Chaudhuri et al., 2012).
Nuclear Top1 has proven to be a successful target for
anticancer drugs. Indeed, accumulation of Top1cc after Top1-
inhibitor treatment leads to DNA lesions and ultimately
triggers apoptosis and cell death (Hsiang et al., 1985; Covey
et al., 1989; Tanizawa et al., 1994; Pommier et al., 2010, 2013).
In our study we show that treating cells with Lam-D results in
both the trapping of Top1mt on, and significant damage to,
mtDNA (Fig. 4). These findings suggest a direct role of Top1mt
trapping in the generation of mtDNA damage as an alterna-
tive, or in addition, to damage resulting from mitochondrial
defects reported by Ballot et al. (2009, 2010).
Several studies have demonstrated that mtDNA muta-
tions are common in cancer (Lu et al., 2009) and numerous
Fig. 3. Lamellarin D accumulates preferentially inside mitochondria. (A)
Representative confocal images demonstrating prominent cytoplasmic
and mitochondrial staining of Lam-D in SK-MEL-5 melanoma cells,
monitored by the intrinsic fluorescence (blue channel, left and right
panels) of Lam-D, which co-localizes with MitoTracker (red channel,
middle and right panels). Right panels show the merged images. The lower
panels are expanded views of the boxed images in the upper panels. (B)
Quantitation of the pixel intensity of fluorescence along with the arrow
indicated in the expanded view of the merged image. The arrow was drawn
arbitrarily over a section of interest.
Poisoning of Mitochondrial Topoisomerase I by Lamellarin D 197
at ASPET Journals on February 21, 2015molpharm.aspetjournals.orgDownloaded from
polymorphisms and mutations of mtDNA correlate with an
increased risk of developing malignancies, including breast
(Canter et al., 2005) and prostate cancers (Petros et al., 2005).
Moreover, recent evidence revealed that mtDNA amounts
regulate response to chemotherapy (Singh et al., 1999; Naito
et al., 2008; Hsu et al., 2010). Different strategies have been
reported for targeting drugs to mitochondria (Heller et al.,
2012; Yousif et al., 2009), including nanoparticles (Marrache
and Dhar, 2012). Recently, the two most used anticancer
agents, cisplatin (Wisnovsky et al., 2013) and doxorubicin
(Chamberlain et al., 2013), have been targeted to mitochon-
drial DNA. The robust Top1mt inhibition activity of Lam-D
displayed in our study suggests that Lam-D can be poten-
tially used as an anticancer reagent targeted to Top1mt in
mitochondria.
In summary, our study demonstrates that the marine
alkaloid lamellarin D is the first drug to target mitochondrial
DNA by trapping Top1mt cleavage intermediates (Top1mt
cleavage complexes) and suggests that developing Top1mt
inhibitors could be an alternative strategy for targeting
mitochondrial DNA.
Acknowledgments
The authors thank Dr. Carmen Cuevas Marchante and José M.
Fernandez Sousa-Faro, PharmaMar, for providing lamellarin D.
Authorship Contributions
Participated in research design: Khiati, Seol, Agama, Dalla Rosa,
Zhang, Neuman, Pommier.
Conducted experiments: Khiati, Seol, Agama, Agrawal, Fesen.
Performed data analysis: Khiati, Seol, Agama, Dalla Rosa, Neuman,
Pommier.
Wrote or contributed to the writing of the manuscript: Khiati, Seol,
Agama, Dalla Rosa, Neuman, Pommier.
Fig. 4. Lamellarin D traps Top1mt in
human cells. (A and B) Top1mt cleavage
complexes in melanoma SK-MEL-5 and
lung carcinoma NCI_H23 cells treated
with Lam-D for 3 hours, respectively. (C)
Time course of Top1mtcc after Lam-D
treatment at 10 mM in SK-MEL-5 cells.
Top1mtcc were detected by ICE bioassay.
(D) Induction of mtDNA damage. Repre-
sentative agarose gel images of mtDNA
long- and short-fragment PCR (Long F
PCR and Short F PCR, respectively) in
SK-MEL-5 cells treated with Lam-D for
24 hours. (E) Quantification of mtDNA
damage using the ratio of the long versus
short fragment PCR products (Tukeys
multiple comparison test; **P,0.01 and
***P,0.001). (F) Specificity of the
Top1mt and Top1 antibodies demon-
strated by a representative Western blot
with recombinant Top1mt and Top1 blot-
ted with Top1mt antibody (left panel) and
Top1 antibody (right panel). (G) Compar-
ison of Top1mtcc induced by CPT and
Lam-D. (H) Comparison of nuclear Top1cc
induced by CPT and Lam-D under similar
conditions.
198 Khiati et al.
at ASPET Journals on February 21, 2015molpharm.aspetjournals.orgDownloaded from
References
Antony S, Agama KK, Miao ZH, Takagi K, Wright MH, Robles AI, Varticovski L,
Nagarajan M, Morrell A, and Cushman M et al. (2007) Novel indenoisoquinolines
NSC 725776 and NSC 724998 produce persistent topoisomerase I cleavage com-
plexes and overcome multidrug resistance. Cancer Res 67:1039710405.
Bailly C (2004) Lamellarins, from A to Z: a family of anticancer marine pyrrole
alkaloids. Curr Med Chem Anticancer Agents 4:363378.
Ballot C, Kluza J, Martoriati A, Nyman U, Formstecher P, Joseph B, Bailly C,
and Marchetti P (2009) Essential role of mitochondria in apoptosis of cancer cells
induced by the marine alkaloid Lamellarin D. Mol Cancer Ther 8:33073317.
Ballot C, Kluza J, Lancel S, Martoriati A, Hassoun SM, Mortier L, Vienne JC, Briand
G, Formstecher P, and Bailly C et al. (2010) Inhibition of mitochondrial respiration
mediates apoptosis induced by the anti-tumoral alkaloid lamellarin D. Apoptosis
15:769781.
Burke TG and Mi Z (1994) The structural basis of camptothecin interactions with
human serum albumin: impact on drug stability. J Med Chem 37:4046.
Canter JA, Kallianpur AR, Parl FF, and Millikan RC (2005) Mitochondrial DNA
G10398A polymorphism and invasive breast cancer in African-American women.
Cancer Res 65:80288033.
Chamberlain GR, Tulumello DV, and Kelley SO (2013) Targeted delivery of doxo-
rubicin to mitochondria. ACS Chem Biol 8:13891395.
Chinnery PF and Hudson G (2013) Mitochondrial genetics. Br Med Bull 106:135159.
Clayton DA (1991) Replication and transcription of vertebrate mitochondrial DNA.
Annu Rev Cell Biol 7:453478.
Covey JM, Jaxel C, Kohn KW, and Pommier Y (1989) Protein-linked DNA strand
breaks induced in mammalian cells by camptothecin, an inhibitor of topoisomerase
I. Cancer Res 49:50165022.
Das BB, Dexheimer TS, Maddali K, and Pommier Y (2010) Role of tyrosyl-DNA
phosphodiesterase (TDP1) in mitochondria. Proc Natl Acad Sci USA 107:
1979019795.
Dexheimer TS and Pommier Y (2008) DNA cleavage assay for the identification of
topoisomerase I inhibitors. Nat Protoc 3:17361750.
Douarre C, Sourbier C, Dalla Rosa I, Brata Das B, Redon CE, Zhang H, Neckers L,
and Pommier Y (2012) Mitochondrial topoisomerase I is critical for mitochondrial
integrity and cellular energy metabolism. PLoS ONE 7:e41094.
Facompré M, Tardy C, Bal-Mahieu C, Colson P, Perez C, Manzanares I, Cuevas C,
and Bailly C (2003) Lamellarin D: a novel potent inhibitor of topoisomerase I.
Cancer Res 63:73927399.
Gentry AC, Juul S, Veigaard C, Knudsen BR, and Osheroff N (2011) The geometry of
DNA supercoils modulates the DNA cleavage activity of human topoisomerase I.
Nucleic Acids Res 39:10141022.
Greaves LC, Reeve AK, Taylor RW, and Turnbull DM (2012) Mitochondrial DNA and
disease. JPathol 226:274286.
Heller A, Brockhoff G, and Goepferich A (2012) Targeting drugs to mitochondria. Eur
J Pharm Biopharm 82:118.
Hsiang YH, Hertzberg R, Hecht S, and Liu LF (1985) Camptothecin induces protein-
linked DNA breaks via mammalian DNA topoisomerase I. J Biol Chem 260:
1487314878.
Hsu CW, Yin PH, Lee HC, Chi CW, and Tseng LM (2010) Mitochondrial DNA content
as a potential marker to predict response to anthracycline in breast cancer
patients. Breast J 16:264270.
Hunter SE, Jung D, Di Giulio RT, and Meyer JN (2010) The QPCR assay for analysis
of mitochondrial DNA damage, repair, and relative copy number. Methods 51:
444451.
Ioanoviciu A, Antony S, Pommier Y, Staker BL, Stewart L, and Cushman M (2005)
Synthesis and mechanism of action studies of a series of norindenoisoquinoline
topoisomerase I poisons reveal an inhibitor with a flipped orientation in the ter-
nary DNA-enzyme-inhibitor complex as determined by X-ray crystallographic
analysis. J Med Chem 48:48034814.
Kluza J, Gallego MA, Loyens A, Beauvillain JC, Sousa-Faro JM, Cuevas C, Marchetti
P, and Bailly C (2006) Cancer cell mitochondria are direct proapoptotic targets for
the marine antitumor drug lamellarin D. Cancer Res 66:31773187.
Koster DA, Palle K, Bot ESM, Bjornsti M-A, and Dekker NH (2007) Antitumour
drugs impede DNA uncoiling by topoisomerase I. Nature 448:213217.
Li Q, Jiang J, Fan A, Cui Y, and Jia Y (2011) Total synthesis of lamellarins D, H, and
R and ningalin B. Org Lett 13:312315.
Low RL, Orton S, and Friedman DB (2003) A truncated form of DNA topoisomerase
IIbeta associates with the mtDNA genome in mammalian mitochondria. Eur J
Biochem 270:41734186.
Lu J, Sharma LK, and Bai Y (2009) Implications of mitochondrial DNA mutations
and mitochondrial dysfunction in tumorigenesis. Cell Res 19:802815.
Marrache S and Dhar S (2012) Engineering of blended nanoparticle platform for
delivery of mitochondria-acting therapeutics. Proc Natl Acad Sci USA 109:
1628816293.
Medikayala S, Piteo B, Zhao X, and Edwards JG (2011) Chronically elevated glucose
compromises myocardial mitochondrial DNA integrity by alteration of mitochon-
drial topoisomerase function. Am J Physiol Cell Physiol 300:C338C348.
Naito A, Carcel-Trullols J, Xie CH, Evans TT, Mizumachi T, and Higuchi M (2008)
Induction of acquired resistance to antiestrogen by reversible mitochondrial DNA
depletion in breast cancer cell line. Int J Cancer 122:15061511.
Neuzil J, Dong LF, Rohlena J, Truksa J, and Ralph SJ (2013) Classification of
mitocans, anti-cancer drugs acting on mitochondria. Mitochondrion 13:199208.
Nitiss J and Wang JC (1988) DNA topoisomerase-targeting antitumor drugs can be
studied in yeast. Proc Natl Acad Sci USA 85:75017505.
Olszewska A and Szewczyk A (2013) Mitochondria as a pharmacological target:
magnum overview. IUBMB Life 65:273281.
Petros JA, Baumann AK, Ruiz-Pesini E, Amin MB, Sun CQ, Hall J, Lim S, Issa MM,
Flanders WD, and Hosseini SH et al. (2005) mtDNA mutations increase tumori-
genicity in prostate cancer. Proc Natl Acad Sci USA 102:719724.
Pla D, Albericio F, and Alvarez M (2008) Recent advances in lamellarin alkaloids:
isolation, synthesis and activity. Anticancer Agents Med Chem 8:746760.
Pommier Y (2013) Drugging topoisomerases: lessons and challenges. ACS Chem Biol
8:8295.
Pommier Y, Leo E, Zhang H, and Marchand C (2010) DNA topoisomerases and their
poisoning by anticancer and antibacterial drugs. Chem Biol 17:421433.
Pommier Y and Marchand C (2012) Interfacial inhibitors: targeting macromolecular
complexes. Nat Rev Drug Discov 11:2536.
Pourquier P, Ueng LM, Fertala J, Wang D, Park HJ, Essigmann JM, Bjornsti MA,
and Pommier Y (1999) Induction of reversible complexes between eukaryotic DNA
topoisomerase I and DNA-containing oxidative base damages. 7, 8-dihydro-8-
oxoguanine and 5-hydroxycytosine. J Biol Chem 274:85168523.
Ray Chaudhuri A, Hashimoto Y, Herrador R, Neelsen KJ, Fachinetti D, Bermejo R,
Cocito A, Costanzo V, and Lopes M (2012) Topoisomerase I poisoning results in
PARP-mediated replication fork reversal. Nat Struct Mol Biol 19:417423.
Rowe TC, Weissig V, and Lawrence JW (2001) Mitochondrial DNA metabolism tar-
geting drugs. Adv Drug Deliv Rev 49:175187.
Seol Y, Zhang H, Pommier Y, and Neuman KC (2012) A kinetic clutch governs reli-
gation by type IB topoisomerases and determines camptothecin sensitivity. Proc
Natl Acad Sci USA 109:1612516130.
Singh KK, Russell J, Sigala B, Zhang Y, Williams J, and Keshav KF (1999) Mito-
chondrial DNA determines the cellular response to cancer therapeutic agents.
Oncogene 18:66416646.
Sobek S, Dalla Rosa I, Pommier Y, Bornholz B, Kalfalah F, Zhang H, Wiesner RJ, von
Kleist-Retzow JC, Hillebrand F, Schaal H, Mielke C, Christensen MO and Boege F
(2013) Negative regulation of mitochondrial transcription by mitochondrial top-
oisomerase I. Nucleic Acids Res 41:98489457.
Stewart L, Redinbo MR, Qiu X, Hol WG, and Champoux JJ (1998) A model for the
mechanism of human topoisomerase I. Science 279:15341541.
Subramanian D, Kraut E, Staubus A, Young DC, and Muller MT (1995) Analysis of
topoisomerase I/DNA complexes in patients administered topotecan. Cancer Res
55:20972103.
Takebayashi Y, Pourquier P, Yoshida A, Kohlhagen G, and Pommier Y (1999) Poi-
soning of human DNA topoisomerase I by ecteinascidin 743, an anticancer drug
that selectively alkylates DNA in the minor groove. Proc Natl Acad Sci USA 96:
71967201.
Tanizawa A, Fujimori A, Fujimori Y, and Pommier Y (1994) Comparison of top-
oisomerase I inhibition, DNA damage, and cytotoxicity of camptothecin derivatives
presently in clinical trials. J Natl Cancer Inst 86:836842.
Wang Y, Lyu YL, and Wang JC (2002) Dual localization of human DNA top-
oisomerase IIIalpha to mitochondria and nucleus. Proc Natl Acad Sci USA 99:
1211412119.
Wisnovsky SP, Wilson JJ, Radford RJ, Pereira MP, Chan MR, Laposa RR, Lippard
SJ, and Kelley SO (2013) Targeting mitochondrial DNA with a platinum-based
anticancer agent. Chem Biol 20:13231328.
Yousif LF, Stewart KM, and Kelley SO (2009) Targeting mitochondria with
organelle-specific compounds: strategies and applications. ChemBioChem 10:
19391950.
Zhang H, Barceló JM, Lee B, Kohlhagen G, Zimonjic DB, Popescu NC, and Pommier
Y (2001) Human mitochondrial topoisomerase I. Proc Natl Acad Sci USA 98:
1060810613.
Zhang H and Pommier Y (2008) Mitochondrial topoisomerase I sites in the regulatory
D-loop region of mitochondrial DNA. Biochemistry 47:1119611203.
Address correspondence to: Dr. Yves Pommier, Laboratory of Molecular
Pharmacology and Developmental Therapeutics Branch, Center for Cancer
Research, National Cancer Institute, NIH, Bldg. 37, Rm. 5068, 37 Convent
Drive, MSC 4255, Bethesda, MD 20814. E-mail: pommier@nih.gov
Poisoning of Mitochondrial Topoisomerase I by Lamellarin D 199
at ASPET Journals on February 21, 2015molpharm.aspetjournals.orgDownloaded from
... In contrast to camptothecin, lamellarin D slows down the relaxation of mitochondrial Topo-I and strongly inhibits DNA relegation by this mitochondrial enzyme. 35 The initial study of Faulkner and coll. on the Topo-I inhibition by a series of ascidian alkaloids, i.e., the lamellarins involved HIV-1 integrase and the topoisomerase of the Molluscum contagiosum virus (MCV), 36 indicated that whereas none of the studied compounds as 20-sulfate of lamellarin-a, U, and V (31)(32)(33), and non-sulfated forms of lamellarin N (9), T (34), and W (35) (Fig. 2) inhibited MCV topoisomerase at concentration <100 mM, only lamellarin-a 20-sulfate showed selective inhibitory activity against HIV-1 integrase and HIV-1 virus in cell culture (IC 50 = 8 mM), being less toxic in HeLa cells (LD 50 = 274 mM). ...
... Subsequently, it was found that lam-H (16) had a potent antitopoisomerase activity (IC 50 = 0.23 mM) but lacked the speci-city required to be medicinally useful given that it was quite cytotoxic toward HeLa cells (LD 50 = 5.7 mM). 37 In addition, lamellarins M (5), N (9), H (16), X (11), W (35), and B (36) ( Fig. 1 and 2) also showed potent inhibition of human Topo-I. 2 On the other hand, lam-N analogs were utterly inactive against Topo-I even at 50 mM and less potent than lam-D. ...
Article
Full-text available
Fused pyrrolo[2,1-a]isoquinolines have emerged as compelling molecules with remarkably potent cytotoxic activity and topoisomerase inhibitors. This comprehensive review delves into the intricate world of this family of compounds, analyzing the natural marine lamellarins known for their diverse and complex chemical structures, exploring structure–activity relationships (SARs), and highlighting their remarkable versatility. The review emphasizes their fundamental role as topoisomerase inhibitors and cytotoxic agents, as well as some crucial aspects of the chemistry of pyrrolo[2,1-a]isoquinolines, exploring synthetic strategies in total synthesis and molecular diversification trends, highlighting their importance in the field of medicinal chemistry and beyond.
... Bidirectional transcription of mtDNA results in the accumulation of negative supercoils behind the RNA polymerases due to its circular nature. To address this, TOP1mt activity is required, which is similar to the known role of nuclear TOP1 in transcription [25,26]. In addition to regulating mtDNA, TOP1MT also interacts with small mitochondrial subunits to influence mitochondrial translation [27]. ...
Article
Full-text available
Background Abnormal glucose metabolism is one of the determinants of maintaining malignant characteristics of cancer. Targeting cancer metabolism is regarded as a new strategy for cancer treatment. Our previous studies have found that TOP1MT is a crucial gene that inhibits glycolysis and cell metastasis of gastric cancer (GC) cells, but the mechanism of its regulation of glycolysis remains unclear. Methods Transcriptome sequencing data, clinic-pathologic features of GC from a variety of public databases, and WGCNA were used to identify novel targets of TOP1MT. Immunohistochemical results of 250 patients with GC were used to analyze the relative expression relationship between TOP1MT and PDK4. The function of TOP1MT was investigated by migration assays and sea-horse analysis in vitro. Results We discovered a mitochondrial topoisomerase I, TOP1MT, which correlated with a higher risk of metastasis. Functional experiments revealed that TOP1MT deficiency promotes cell migration and glycolysis through increasing PDK4 expression. Additionally, the stimulating effect of TOP1MT on glycolysis may be effectively reversed by PDK4 inhibitor M77976. Conclusions In brief, our work demonstrated the critical function of TOP1MT in the regulation of glycolysis by PDK4 in gastric cancer. Inhibiting glycolysis and limiting tumor metastasis in GC may be accomplished by suppressing PDK4.
... Pharmacologically, lamellarins pyrrole alkaloids (LPAs) displayed a myriad of intriguing powerful biomedical potentialities including, cytotoxicity and topoisomerase I-mediated DNA cleavage antitumor activities (Bailly, 2015;Imperatore et al., 2014;Kluza et al., 2006;Tardy et al., 2004), mitochondrial function inhibitors (Ballot et al., 2010;Bayet-Robert et al., 2010;Khiati et al., 2014), protein kinases inhibitors (Baunbaek et al., 2008;Neagoie et al., 2012;Yoshida et al., 2013), multidrug resistance reversal activity (Huang et al., 2014;Quesada et al., 1996;Zhang et al., 2012). ...
Article
Full-text available
The new coronavirus variant (SARS-CoV-2) and Zika virus are two worldwide health pandemics. Along history, natural products-based drugs have always crucially recognised as a main source of valuable medications. Considering the SARS-CoV-2 and Zika main proteases (Mpro) as the reproduction key element of the viral cycle and its main target, herein we report an intensive computer-aided virtual screening for a focused list of 39 marine lamellarins pyrrole alkaloids, against SARS-CoV-2 and Zika main proteases (Mpro) using a set of combined modern computational methodologies including molecular docking (MDock), molecule dynamic simulations (MDS) and structure-activity relationships (SARs) as well. Indeed, the molecular docking studies had revealed four promising marine alkaloids including [lamellarin H (14)/K (17)] and [lamellarin S (26)/Z (39)], according to their notable ligand-protein energy scores and relevant binding affinities with the SARS-CoV-2 and Zika (Mpro) pocket residues , respectively. Consequentially, these four chemical hits were further examined thermodynamically though investigating their MD simulations at 100 ns, where they showed prominent stability within the accommodated (Mpro) pockets. Moreover, in-deep SARs studies suggested the crucial roles of the rigid fused polycyclic ring system, particularly aromatic A-and F-rings, position of the phenolic-OH and d-lactone functionalities as essential structural and pharmacophoric features. Finally, these four promising lamellarins alkaloids were investigated for their in-silico ADME using the SWISS ADME platform , where they displayed appropriated drug-likeness properties. Such motivating outcomes are greatly recommending further in vitro/vivo examinations regarding those lamellarins pyrrole alkaloids (LPAs).
... [6][7][8][9][10][11] In detail, lamellarin D (Fig. 1), a leading compound in the family of type I, induces its anticancer activity through topoisomerase I inhibition and mitochondrial targeting that triggers cell death. [12][13][14][15] Lamellarin O was investigated by Huang et al. 16 As reported, cytotoxicity effects of some natural lamellarins were assessed towards colorectal cancer SW620 and its multi-drug resistant daughter cell line SW620/Ad300. Lamellarin O exhibited moderate cytotoxicity against aforementioned cells at IC 50 of 20.0 mM and 22.3 mM respectively. ...
Article
Full-text available
A library of pyrazole-based lamellarin O analogues was synthesized from easily accessible 3(5)-aryl-1H-pyrazole-5(3)-carboxylates which were subsequently modified by bromination, N-alkylation and Pd-catalysed Suzuki cross-coupling reactions. Synthesized ethyl and methyl 3,4-diaryl-1-(2-aryl-2-oxoethyl)-1H-pyrazole-5-carboxylates were evaluated for their physicochemical property profiles and in vitro cytotoxicity against three human colorectal cancer cell lines HCT116, HT29, and SW480. The most active compounds inhibited cell proliferation in a low micromolar range. Selected ethyl 3,4-diaryl-1-(2-aryl-2-oxoethyl)-1H-pyrazole-5-carboxylates were further investigated for their mode of action. Results of combined viability staining via Calcein AM/Hoechst/PI and fluorescence-activated cell sorting data indicated that cell death was triggered in a non-necrotic manner mediated by mainly G2/M-phase arrest.
... ETC's disorder of mitochondrial inner membrane contributes to excessive production of ROS and subsequent OS following mitochondrial dysfunction (Sorrentino et al., 2018). Mitochondria are the main target for toxicity (Khiati et al., 2014) because they perform an essential function in detoxification. Previous studies showed that intracellular ROS accumulation is the main reason for the damage of mitochondrial ETC (Cheraghi et al., 2019). ...
Article
Full-text available
This research aimed to evaluate the curative impact of nobiletin (NOB) on paraquat (PQ) induced mito-chondrial impairment in rats' hepatic tissues. Twenty-four male rats were divided into control, PQ, PQ + NOB, and NOB treated groups. The alanine aminotransferase levels, aspartate aminotransferase, and alkaline phosphatase were increased in the experimental animals after exposure with PQ. A significant decrease in the level of catalase, glutathione peroxidase, superoxide dismutase, glutathione, thiobar-bituric acid reactive substances, and reactive oxygen species were recorded in rats exposed to PQ. The mitochondrial TCA cycle enzyme activities, including isocitrate-dehydrogenase, succinate-dehydrogenase, alpha-ketoglutarate dehydrogenase, and malate-dehydrogenase, were reduced in PQ exposed rats. The activities of mitochondrial enzymes including NADH dehydrogenase, coenzyme Q-cytochrome reductase, succinic-coenzyme Q and cytochrome c-oxidase were significantly decreased after the PQ-exposure. PQ administration also reduced the mitochondrial membrane potential. However, the administration of mitochondria with NOB can decrease the toxic effect of the PQ in isolated mitochondria. NOB treatment potentially reduced the damaging effects of the PQ in the mitochondria isolated from hep-atic tissues. Thus, the current study revealed that the NOB could attenuate PQ-induced mitochondrial damage in rats' hepatic tissues. Ó 2021 The Author(s). Published by Elsevier B.V. on behalf of King Saud University. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Article
Mitochondrial Topisomerase 1 (Top1mt) is critical for mtDNA replication, transcription, and energy production. Here, we investigate the carrier-mediated targeted delivery of the anticancer drug Irinotecan into the mitochondria to selectively...
Article
Full-text available
In a growing multidrug-resistant environMent, the identification of potential new drug candidates with an acceptable safety profile is a substantial crux in pharmaceutical discovery. This review discusses several aspects and properties of approved Marine Natural Products derived from ascidian sources (phylum Chordata, subphylum Tunicata) and/or their deduced analogues including their biosynthetic origin, (bio)chemical preclinical assessments and known efficacy-safety profiles, clinical status in trials, but also translational developments, opportunities and final conclusions. The review also describes the preclinical assessments of a large number of other ascidian compounds that have not been involved in clinical trials yet. Finally, the emerging research on the connectivity of the ascidian hosts and their independent or obligate symbiotic guests is discussed. The review covers the latest information on the topic of ascidian-derived Marine Natural Products over the last two decades including 2022, with the majority of publications published in the last decade.
Article
Full-text available
Since the discovery of anticancer properties of a naturally occurring hexacyclic marine alkaloid Lamellarin D, the attempts have been made to prepare its synthetic analogues and elucidate the effects of each structural component on their activity profile. While F‐ring‐free, A‐ring‐free and B‐ring‐open lamellarins are known, E‐ring‐free analogues have never been investigated. In this work, we developed a facile and straightforward synthetic method toward E‐ring‐free lamellarin analogues based on the [3+2]‐cycloaddition. For the first time, we prepared several pentacyclic lamellarin analogues without E‐ring in their structure and assessed their cytotoxicity in a panel of cancer cell lines in comparison with several hexacyclic lamellarins. E‐ring‐free lamellarins were devoid of cytotoxicity due to their poor solubility in cellular environment.
Article
Full-text available
Mitochondrial topoisomerase I is a genetically distinct mitochondria-dedicated enzyme with a crucial but so far unknown role in the homeostasis of mitochondrial DNA metabolism. Here, we present data suggesting a negative regulatory function in mitochondrial transcription or transcript stability. Deficiency or depletion of mitochondrial topoisomerase I increased mitochondrial transcripts, whereas overexpression lowered mitochondrial transcripts, depleted respiratory complexes I, III and IV, decreased cell respiration and raised superoxide levels. Acute depletion of mitochondrial topoisomerase I triggered neither a nuclear mito-biogenic stress response nor compensatory topoisomerase IIβ upregulation, suggesting the concomitant increase in mitochondrial transcripts was due to release of a local inhibitory effect. Mitochondrial topoisomerase I was co-immunoprecipitated with mitochondrial RNA polymerase. It selectively accumulated and rapidly exchanged at a subset of nucleoids distinguished by the presence of newly synthesized RNA and/or mitochondrial RNA polymerase. The inactive Y559F-mutant behaved similarly without affecting mitochondrial transcripts. In conclusion, mitochondrial topoisomerase I dampens mitochondrial transcription and thereby alters respiratory capacity. The mechanism involves selective association of the active enzyme with transcriptionally active nucleoids and a direct interaction with mitochondrial RNA polymerase. The inhibitory role of topoisomerase I in mitochondrial transcription is strikingly different from the stimulatory role of topoisomerase I in nuclear transcription.
Data
Full-text available
Human tyrosyl-DNA phosphodiesterase (TDP1) hydrolyzes the phosphodiester bond at a DNA 3′-end linked to a tyrosyl moiety and has been implicated in the repair of topoisomerase I (Top1)-DNA covalent complexes. TDP1 can also hydrolyze other 3′-end DNA alterations including 3′-phosphoglycolate and 3′-abasic sites, and exhibits 3′-nucleosidase activity indicating it may function as a general 3′-end-processing DNA repair enzyme. Here, using laser confocal microscopy, subcellular fractionation and biochemical ana-lyses we demonstrate that a fraction of the TDP1 encoded by the nuclear TDP1 gene localizes to mitochondria. We also show that mitochondrial base excision repair depends on TDP1 activity and provide evidence that TDP1 is required for efficient repair of oxidative damage in mitochondrial DNA. Together, our findings provide evidence for TDP1 as a novel mitochondrial enzyme. DNA repair ∣ ligase III ∣ oxidative DNA damage ∣ topoisomerase I ∣ mitochondrial BER M itochondrial DNA (mtDNA) is an essential component of eukaryotic cells because it encodes a critical subset of mitochondrial proteins for the production of cellular ATP. Each mitochondrion contains 4–6 copies of the double-stranded circu-lar 16 kilobase long mitochondrial genome. MtDNA needs to be tightly preserved because more than 93% of the mitochondrial genome has to be accurately transcribed into 13 individual mes-senger RNAs coding for the essential mitochondrial proteins that are part of the mitochondrial electron transport chain (oxidative phosphorylation) (1). MtDNA also encodes 2 mitochondrial-specific ribosomal RNAs and 22 transfer RNAs that are essential for protein synthesis inside mitochondria. Mitochondria go through alternative rounds of fission and fusion to maintain mitochondrial integrity and mtDNA copy number (2). Because of the proximity of reactive oxygen species (ROS) generated by the mitochondrial oxidative phosphorylation chain, mtDNA is potentially exposed to oxidative DNA damage (3, 4). Accumu-lation of mtDNA damage has been involved in neurodegenera-tive disorders (Parkinson, Alzheimer's, and Huntington diseases), myopathies and diabetes, and has been associated with aging, cancer, and age-related degenerative disorders (5, 6). Mitochondria are dependent on the nucleus for all the mito-chondrial proteins necessary for mtDNA replication, repair, and maintenance (6, 7). Noticeably, some nuclear genes encode pro-teins exclusively for the mitochondria such as polymerase gamma (POLγ), mitochondrial helicase (TWINKLE), transcription factor A (TFAM) (8, 9), and vertebrate mitochondrial topoisomerase IB (Top1mt) (10, 11). Moreover, the mitochondrial genome does not encode DNA repair proteins, and is thus completely dependent on proteins encoded in the nucleus for its repair and integrity. The corresponding genes encode proteins for both nuclear and mitochondrial repair (12–14). Yet, the full spectrum of nuclear DNA repair proteins involved in mtDNA repair is not entirely known. MtDNA repair primarily uses base excision repair (BER) and lacks nucleotide excision repair (6, 13, 15). However, mismatch repair (16) and DNA double-strand break (DSB) end joining (17–19) activities have been reported in mitochondria. Recently, several reports have focused on the existence of long-patch BER (LP-BER) by the flap endonucleases FEN-1 and DNA2 in mito-chondria (13, 20–23). ROS can also generate 3′-deoxyribose residues that are oxidized, unsaturated or fragmentary (e.g., 3′-phosphoglycolates and 3′-phosphate). The DNA glycosylases with associated AP-lyase activity would generate 2,3-unsaturated deoxyribose as an indirect product of DNA oxidation (24). Abasic endonuclease 1 (APE1) is the a well characterized enzyme in mammals for the repair of 3′-phosphoglycolate esters (3′-PG) dur-ing oxidative DNA damage (25–27). Recently, APE1 has been shown to be localized both in nuclei and mitochondria (14, 28). However, the 3′-PG removal activity of APE1 is highly selective and APE1 is relatively ineffective when the 3′-PG is in single-stranded DNA, at 3′-overhangs or at blunt or recessed 3′-ends of DSBs (27, 29–32). Furthermore, APE1 is not able to remove topoisomerase-DNA complexes (33, 34), which can be trapped by the endogenous lesions generated by ROS (oxidized bases, abasic sites, and strand breaks) (13, 35–37). Human tyrosyl-DNA phosphodiesterase (TDP1) typically hydrolyzes the phosphodiester bond between a tyrosyl moiety and a DNA 3′-end (33, 34). TDP1 was originally discovered in yeast (34) and has been implicated in the repair of stalled Top1-DNA covalent complexes (38, 39). The ability of TDP1 to resolve 3′-phosphotyrosyl linkages is consistent with its role in protecting cells against Top1-DNA lesions (40–43). Homozygous mutation of TDP1 causes spinocerebellar ataxia with axonal neuropathy (SCAN1), an autosomal recessive neurodegenerative syndrome (44). Cells from SCAN1 patients are hypersensitive to the specific Top1 poison camptothecin and accumulate elevated Top1-asso-ciated DNA breaks in response to camptothecin (39, 43, 45). TDP1 activity is not limited to the removal of Top1 adducts. TDP1 can also process other 3′-DNA end blocking groups: 3′-aba-sic sites and 3′-phosphoglycolates (45–49). Accordingly, TDP1-deficient cells are deficient in the removal of 3′-phosphoglycolate and are hypersensitive to bleomycin in addition to their hypersen-sitivity to camptothecin (39, 41–43, 49). TDP1 also possesses a limited DNA and RNA 3′-nucleosidase activity in which a single nucleoside is removed from the 3′-hydroxyl end of the substrate (45). Thus, TDP1 may function to remove a variety of adducts from 3′-DNA ends during DNA repair (50). Yeast TDP1 has also been shown to process Top2-DNA adducts (51). In this study, we employed Immunofluorescence staining, cellular fractionation, Western blotting analysis, and biochemical assays to demonstrate the presence of TDP1 in mitochondria and its function in repairing mtDNA.
Article
Full-text available
Introduction In the last 10 years the field of mitochondrial genetics has widened, shifting the focus from rare sporadic, metabolic disease to the effects of mitochondrial DNA (mtDNA) variation in a growing spectrum of human disease. The aim of this review is to guide the reader through some key concepts regarding mitochondria before introducing both classic and emerging mitochondrial disorders. Sources of data In this article, a review of the current mitochondrial genetics literature was conducted using PubMed (http://www.ncbi.nlm.nih.gov/pubmed/). In addition, this review makes use of a growing number of publically available databases including MITOMAP, a human mitochondrial genome database (www.mitomap.org), the Human DNA polymerase Gamma Mutation Database (http://tools.niehs.nih.gov/polg/) and PhyloTree.org (www.phylotree.org), a repository of global mtDNA variation. Areas of agreement The disruption in cellular energy, resulting from defects in mtDNA or defects in the nuclear-encoded genes responsible for mitochondrial maintenance, manifests in a growing number of human diseases. Areas of controversy The exact mechanisms which govern the inheritance of mtDNA are hotly debated. Growing points Although still in the early stages, the development of in vitro genetic manipulation could see an end to the inheritance of the most severe mtDNA disease.
Article
Camptothecin was recently identified as an inhibitor of mammalian topoisomerase I. Similar to inhibitors of topoisomerase II, camptothecin produces DNA single-strand breaks (SSB) and DNA-protein cross-links (DPC) in mammalian cells. However, their one-to-one association, expected for trapped topoisomerase complexes, has not previously been demonstrated. We have studied camptothecin-induced SSB and DPC in Chinese hamster DC3F cells and their isolated nuclei, using the DNA alkaline elution technique. It was found that the SSB and DPC frequencies detected following camptothecin treatment depend upon the conditions used for lysis. When lysis was with sodium dodecyl sulfate, the observed frequencies of SSB and DPC were 2- to 3-fold greater than when sodium dodecyl sarkosinate (Sarkosyl) was used. In either case, the SSB:DPC ratio was close to 1. All of the camptothecin-induced SSB were protein linked, as indicated by the absence of DNA elution under nondeproteinizing conditions. DNA cleavage assays with purified topoisomerase I also indicated that the weaker Sarkosyl detergent fails to trap all of the enzyme-DNA complexes. In contrast, lysis conditions had little effect on levels of SSB or DPC produced by 4&apos;-(9-acridinylamino)-methanesulfon-m-anisidide, suggesting that trapping of topoisomerase II complexes occurs equally well with either detergent. In experiments using isolated nuclei, it was found that the camptothecin-induced SSB, in contrast to trapped topoisomerase II complexes, can form and reverse within minutes at 4 degrees C. The activity of camptothecin at low temperature was also seen with purified topoisomerase I. These results support the hypothesis that the SSB and DPC induced by camptothecin in mammalian cells are due to an action on topoisomerase I.
Article
An analog of the anticancer drug cisplatin (mtPt) was delivered to mitochondria of human cells using a peptide specifically targeting this organelle. mtPt induces apoptosis without damaging nuclear DNA, indicating that mtDNA damage is sufficient to mediate the activity of a platinum-based chemotherapeutic. This study demonstrates the specific delivery of a platinum drug to mitochondria and investigates the effects of directing this agent outside the nucleus.
Article
Several families of highly effective anticancer drugs are selectively toxic to cancer cells because they disrupt nucleic acids synthesis in the nucleus. Much less is known, however, about whether interfering with nucleic acids synthesis in the mitochondria would have significant cellular effects. In this study, we explore this with a mitochondrially-targeted form of the anticancer drug doxorubicin, which inhibits DNA topoisomerase II - an enzyme that is both in mitochondria and nuclei of human cells. When doxorubicin is attached to a peptide that targets mitochondria, it exhibits significant toxicity. However, when challenged with a cell line that overexpresses a common efflux pump, it does not exhibit the reduced activity of the nuclear-localized parent drug and resists being removed from the cell. These results indicate that targeting drugs to the mitochondria provides a means to limit drug efflux, and provide evidence that a mitochondrially-targeted DNA topoisomerase poison is active within the organelle.
Article
Mitochondria, responsible for energy metabolism within the cell, act as signaling organelles. Mitochondrial dysfunction may lead to cell death and oxidative stress and may disturb calcium metabolism. Additionally, mitochondria play a pivotal role in cardioprotective phenomena and a variety of neurodegenerative disorders ranging from Parkinson's to Alzheimer's disease. Mitochondrial DNA mutations may lead to impaired respiration. Hence, targeting the mitochondria with drugs offers great potential for new therapeutic approaches. The purpose of this overview is to present the recent state of knowledge concerning the interactions of various substances with mitochondria. © 2013 IUBMB Life, 65(3):273-281, 2013.
Article
Topoisomerases are ubiquitous enzymes that control DNA supercoiling and entanglements. They are essential during transcription and replication and topoisomerase inhibitors are among the most effective and most commonly used anticancer and antibacterial drugs. This review consists in two parts. In the first part ("Lessons"), it gives background information on the catalytic mechanisms of the different enzyme families (6 different genes in humans and 4 in most bacteria), describes the "interfacial inhibition" by which topoisomerase-targeted drugs act as topoisomerase poisons and describes clinically relevant topoisomerase inhibitors. It generalizes the interfacial inhibition principle, which was discovered from the mechanism of action of topoisomerase inhibitors, and discusses how topoisomerase inhibitors kill cells by trapping topoisomerases on DNA rather than by classical enzymatic inhibition. Trapping protein-DNA complexes extends to a novel mechanism of action of PARP inhibitors and could be applied to the targeting of transcription factors. The second part of the review focuses on the challenges for discovery and precise use of topoisomerase inhibitors, including targeting topoisomerase inhibitors using chemical coupling and encapsulation for selective tumor delivery, use of pharmacodynamic biomarkers to follow drug activity, complexity of the response determinants for anticancer activity and patient selection, prospects of rational combinations with DNA repair inhibitors targeting tyrosyl-DNA-phosphodiesterases 1 and 2 (TDP1 and TDP2) and PARP, and the unmeet need to develop inhibitors for type IA enzymes.
Article
Mitochondrial dysfunctions cause numerous human disorders. A platform technology based on biodegradable polymers for carrying bioactive molecules to the mitochondrial matrix could be of enormous potential benefit in treating mitochondrial diseases. Here we report a rationally designed mitochondria-targeted polymeric nanoparticle (NP) system and its optimization for efficient delivery of various mitochondria-acting therapeutics by blending a targeted poly(d,l-lactic-co-glycolic acid)-block (PLGA-b)-poly(ethylene glycol) (PEG)-triphenylphosphonium (TPP) polymer (PLGA-b-PEG-TPP) with either nontargeted PLGA-b-PEG-OH or PLGA-COOH. An optimized formulation was identified through in vitro screening of a library of charge- and size-varied NPs, and mitochondrial uptake was studied by qualitative and quantitative investigations of cytosolic and mitochondrial fractions of cells treated with blended NPs composed of PLGA-b-PEG-TPP and a triblock copolymer containing a fluorescent quantum dot, PLGA-b-PEG-QD. The versatility of this platform was demonstrated by studying various mitochondria-acting therapeutics for different applications, including the mitochondria-targeting chemotherapeutics lonidamine and α-tocopheryl succinate for cancer, the mitochondrial antioxidant curcumin for Alzheimer's disease, and the mitochondrial uncoupler 2,4-dinitrophenol for obesity. These biomolecules were loaded into blended NPs with high loading efficiencies. Considering efficacy, the targeted PLGA-b-PEG-TPP NP provides a remarkable improvement in the drug therapeutic index for cancer, Alzheimer's disease, and obesity compared with the nontargeted construct or the therapeutics in their free form. This work represents the potential of a single, programmable NP platform for the diagnosis and targeted delivery of therapeutics for mitochondrial dysfunction-related diseases.