Vous êtes sur la page 1sur 13

Chemico-Biological Interactions 173 (2008) 84–96

Contents lists available at ScienceDirect

Chemico-Biological Interactions
journal homepage: www.elsevier.com/locate/chembioint

Copper–adenine complex, a compound, with multi-biochemical


targets and potential anti-cancer effect
Hassan H. Hammud a , Georges Nemer b , Walid Sawma b , Jhonny Touma b , Pascale Barnabe b ,
Yolla Bou-Mouglabey b , Amer Ghannoum a , Jida El-Hajjar b , Julnar Usta b,∗
a Beirut Arab University, Department of Chemistry, Beirut, Lebanon
b American University of Beirut, Department of Biochemistry, School of Medicine, Abdul-Aziz Street, Hamra, Beirut, Lebanon

a r t i c l e i n f o a b s t r a c t

Article history: A series of adenine–copper complexes (1–6) with various ligands (Cl− , SCN− , BF4 − and
Received 14 September 2007 acac [acetylacetonate ion]) have been synthesized and characterized by elemental analysis,
Received in revised form 20 February 2008
infrared spectroscopy and thermal analysis. Among the six complexes only complex (1),
Accepted 12 March 2008
Cu2 (adenine)4 Cl4 ·2EtOH (abbreviated as Cu–Ad), demonstrated some toxic effect on differ-
Available online 21 March 2008
ent cell lines. In vitro investigations of the biological effect of Cu–Ad complex have shown
that it: (1) binds genomic DNA; (2) decreases significantly, the viability of cells in culture in a
Keywords:
Copper–adenine concentration (15–125 ␮M)-dependant manner; an estimated IC50 of: 45 ␮M with HepG2;
HepG2 73 ␮M with C2C12; 103 ␮M with NIH3T3; and 108 ␮M with MCF7. Cu–Ad had no effect on
PCR A549 cells; (3) inhibits Taq polymerase-catalyzed reaction; (4) inhibits the binding of the
GATA-5 transcription factor GATA-5 to labeled DNA probes; (5) inhibits mitochondrial NADH-UQ-
TbX20 reductase with an estimated IC50 of 2.8 nmol, but had no effect on succinate dehydrogenase
WND activity; (6) increases reactive oxygen species (60%) at 45 ␮M Cu–Ad; and (7) decreases ATP
(80%) at 50 ␮M Cu–Ad. The new compound Cu2 (adenine)4 Cl4 ·2EtOH (Cu–Ad), belongs to a
class of copper–adenylate complexes that target many biochemical sites and with potential
anti-cancer activity.
© 2008 Elsevier Ireland Ltd. All rights reserved.

1. Introduction dination of the metal to residues at/or outside the active


site of the enzyme, which may block the substrate binding
Metal coordination complexes have been extensively or interfere in the catalytic cycle. Alternatively metal com-
used in clinical applications as enzyme inhibitors [1], plexes targeting enzymes in the life cycle of the virus have
anti-bacterial [2,3], antiviral [4–6] and as anti-cancerous been used as antiviral [5,6]. The human immunodeficiency
[7–9]. Different kinds of metals have been employed in virus for example, encodes three critical enzymes: a reverse
these complexes including platinum, gold, vanadium, iron, transcriptase, an integrase and a protease. The metallo-
molybdenum, cobalt, tin, gallium, copper and many others thiosemicarbazones were recently characterized and found
[8]. to inhibit the reverse transcriptase; thus interfering with
Metals may have structural role in many enzymes. This virus life cycle.
makes them natural targets in drug therapy and develop- Organo-arsenicals were used as anti-bacterial in the
ment of enzyme inhibitors. The structural integrity and treatment of syphilis and as anti-parasitic as feed additives
hence function of the enzyme is influenced by the coor- in livestock [2,10,11].
In addition metal coordination complexes have been
employed in the treatment of cancer. The hunt for new
∗ Corresponding author. drugs designed against specific proteins that are differ-
E-mail address: justa@aub.edu.lb (J. Usta). entially expressed in cancer cells remains an active field

0009-2797/$ – see front matter © 2008 Elsevier Ireland Ltd. All rights reserved.
doi:10.1016/j.cbi.2008.03.005
H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96 85

of research. Cisplatin discovered in 1965, has been use- 2. Methods


ful in the treatment of testicular carcinomas and ovarian,
head and neck, bladder and lung cancers [8]. Gallium 2.1. Chemical preparation of the copper–adenine
complexes inhibited specifically ornithine decarboxylase complexes (1–6)
activity inducible by phorbol esters [1,11]. Seleno com-
pounds have been found to exhibit an inhibitory effect on Scheme 1 summarizes the starting metal salt, the lig-
protein kinase-C (PKC) one of the widely tackled targets of and(s) and the solvent used in the synthesis of compounds
anti-tumor agents [12–14]. Copper(II)–anthracyclin com- (1–6). The experimental details are described below. The
plex exerted various effects on DNA topoisomerase-II [1,15]. formula of compounds (1–6) were identified and ver-
An interesting group of chemotherapeutic agents used ified by: (a) C, H, N elemental analysis performed by
in cancer therapy comprises molecules that interact with Microanalytical Service, Department of Chemistry, Uni-
DNA. The covalent or non-covalent association of agents versity of Surrey, Guilford, Surrey, GU2 5XH and the
with DNA may inhibit mitotic progression and thus would American University of Beirut, Environmental CORE Lab-
be used as antineoplastic drug [15,16]. Alternatively, the oratory, Beirut, Lebanon, (b) IR spectra were recorded
cytotoxic effect exhibited by some cancer therapeutics may with Shimadzu 8300 FTIR spectrophotometer using the
be attributed to damage at nuclear DNA level, thereby inter- KBr pellet method, (c) thermal analysis was performed on
fering with the replication and transcription machinery. Setaram-Labsys instrument using thermogravimetric and
In addition, the anti-tumor effect may involve forma- differential scan calorimetric curves (TG–DSC). The heating
tion of DNA-adduct, site specific cleavage of DNA-strand, rate was 3 ◦ C/min. Dry nitrogen was passed over 15 mg of
alkylation of DNA [17,19] and shape distortion of the the sample placed in a platinum crucible, and (d) UV–vis
genetic material which would interfere with RNA syn- spectra was obtained on Shimadzu UV-2101 PC UV–vis
thesis [20]. Among the most common metal binding scanning spectrophotometer.
sites to DNA, are the hetero-atoms of the nucleoside
bases that form strong complexes with transition metal 2.1.1. Preparation of Cu2 (adenine)4 Cl4 ·2EtOH:
ions [8]. complex (1) (Cu–Ad)
Copper is an essential trace element required for Adenine (0.20 g, 1.48 mmol) was dissolved in boil-
enzymes and animal tissues in biological systems in both ing ethanol (30 ml). To this solution was added 0.04 g
the +1/+2 valence states; hence involved in redox met- (0.74 mmol) of NH4 Cl solution in 10 ml ethanol and 0.39 g
alloenzymes and in oxygenation, and oxygen-carrying (1.49 mmol) of copper acetylacetone solution, {Cu(acac)2 },
proteins. in 20 ml chloroform. The turbid mixture was refluxed for 2 h
The binding of copper ions to specific sites can modify and the greenish-blue precipitate was filtered, washed with
the conformational structures of proteins, polynucleotides, ethanol and dried in an oven under vacuum (38% yield). Ele-
DNA and bio-membranes [18,19]. The binding of Cu to mental analysis for Cu2 C24 H32 N20 O2 Cl4 , % Calcd: C, 31.66;
DNA occurs with high affinity. The crystal structure of the H, 3.55; N, 31.08; and % Found: C, 30.30; H, 3.36; N, 30.35
complex formed between CuCl2 and DNA forms a pseudo- (MW: 901.1 g/mol). Complex (1) has low solubility in both
octahedral geometry and involves copper binding to N7 of DMF and dimethyl sulfoxide (DMSO). Throughout the study
guanine residue. No interaction with ribose hydroxyls is complex (1) was referred to as copper–adenine complex or
observed. It has been found that in aqueous solution coor- Cu–Ad.
dination of Cu to the N7 and N1 sites of purine rings is
pH dependant that diminishes as the pH of the solution
increases [21]. 2.1.2. Preparation of Cu(acac)2 (adenine)·EtOH:
In this study we describe the synthesis of six complex (2)
copper–adenine complexes (1–6). Their ligands and struc- The complex is prepared by refluxing adenine and cop-
tural components were identified by elemental analysis, per acetylacetone (1 mmol) in a mixture of chloroform
infrared spectroscopy and thermal analysis. The cytotoxic (50 ml) and ethanol (25 ml) for 2 h. The blue solution was
effect of these compounds was assessed initially on dif- evaporated for few weeks at room temperature to yield blue
ferent cell lines using trypan blue exclusion dye test. crystals (25% yield). Elemental analysis for CuC17 H25 N5 O5 ,
Only compound (1), Cu2 (adenine)4 Cl4 ·2EtOH, exhibited a % Calcd: C, 46.10; H, 5.69; N, 15.81; and % Found: C, 49.96; H,
cytotoxic effect and resulted in significant cell death in 5.77; N, 16.05 (MW: 442.96 g/mol). Complex (2) is soluble
HepG2 cells, verified as well by 3-(4,5-dimethyl thiazol- in both DMF and DMSO.
2-yl)-2,5-diphenyl tetrazolium bromide (MTT) assay. In
addition the compound was found to bind genomic DNA 2.1.3. Preparation of Cu(acac)(adenine)·H2 O: complex (3)
and to exhibit significant inhibitory effect on different To adenine (0.20 g, 1.48 mmol) in 40 ml boiling water,
biochemical targets including: Taq polymerase-catalyzed a solution of copper acetylacetone (0.39 g, 1.49 mmol) in
reaction, binding of DNA to the transcription factor GATA- 20 ml chloroform was added drop wise. An immediate
5 and complex I of the mitochondrial electron transport violet precipitate was formed; the mixture was refluxed
chain. Cu–Ad is a compound with potential anti-cancer for 1 h following which the precipitate was filtered and
effects; further design and synthesis of structural analogues dried under vacuum (17% yield). Elemental analysis for
is needed to establish proper structure function relation CuC10 H14 N5 O3 , % Calcd: C, 38.03; H, 4.44; N, 22.19; and %
that would increase the specificity towards biochemical Found: C, 38.45; N, 24.1 (MW: 315.5 g/mol). Complex (3)
targets. was neither soluble in DMF nor in DMSO.
86 H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96

Scheme 1. Synthesis of copper–adenine complexes.

2.1.4. Preparation of Cu(adenine)2 Cl2 ·2H2 O: complex (4) mental analysis for CuC8.5 H9 N6 O2.5 S, % Calcd: C, 30.86; H,
To adenine (1.35 g, 1 mmol) solution in 30 ml of boiling 2.72; N, 25.42; and % Found: C, 29.1; H, 2.8; N, 26.49 (MW:
ethanol, a solution of copper chloride dihydrate (0.086 g, 315.55 g/mol). The complex is sparingly soluble in DMF or
0.50 mmol) was added in 5 ml ethanol. The mixture was DMSO.
refluxed for 1 h and the blue precipitate was filtered and
dried under vacuum (32% yield). Elemental analysis for 2.2. In vitro assays
CuC10 H14 N10 Cl2 O2 , % Calcd: C, 27.24; H, 3.1; N, 31.79; and %
Found: C, 25.34; H, 3.04; N, 29.81 (MW: 440.45 g/mol). The 2.2.1. Trypan blue exclusion assay
complex is soluble in both DMF and DMSO. The cytotoxicity of the various complexes on different
cell lines was initially assessed qualitatively using trypan
2.1.5. Preparation of Cu(adenine)2 (BF4 )2 ·2EtOH: blue exclusion test. In 12-well plates, six cell lines (C2C12,
complex (5) HepG2, A549, NIH3T3, MCF7 and MCF7-BcL2) were cul-
To adenine (1.35 g, 1 mmol) solution in boiling ethanol tured in appropriate media (1 ml) under standard incubator
(30 ml), tetrabutyl ammonium tetrafluoroborate (0.68 g, conditions of humidified atmosphere, 95% air, 5% CO2 and
3 mmol) and copper chloride (0.17 g, 1 mmol) in 30 ml temperature 37 ◦ C. Cu–Ad complex at varying concentra-
ethanol were added. The resulting bright yellow solution tion (15–125 ␮M) was added to cells (5 × 104 ). Cell toxicity,
was refluxed for half hour, and the precipitated green com- expressed as %cell death was estimated after 24 h, by deter-
plex was filtered and dried (28% yield). Elemental analysis mining the ratio of dead cells (acquiring a blue color) to the
for CuC14 H22 N10 O2 B2 F8 , % Calcd: C, 28.04; H, 3.67; N, 23.37; total number of cells and compared to a control treated with
and % Found: C, 28.13; H, 3.6; N, 23.51 (MW: 599.17 g/mol). the vehicle DMSO/ethanol not exceeding 2.5%.
The complex is soluble in both DMF and DMSO.
2.2.2. MTT viability assay
2.1.6. Preparation of Cell viability was assessed using MTT, as described by
Cu(acetate)(adenine)(SCN)·0.5CH3 OH: complex (6) the Kit manual (Roche). HepG2 cells (1 × 104 ) cultured in
To copper acetate dihydrate (0.60 g, 3.00 mmol) in 50 ml 96-well plates, were treated for 24 h with varying con-
methanol, adenine (0.44 g, 3.26 mmol) in 25 ml methanol centrations of: Cu–Ad (15–125 ␮M), adenine (15, 50 and
and ammonium thiocyanate (1.00 g, 13 mmol) in 25 ml 125 ␮M) and copper chloride (15, 50 and 125 ␮M), fol-
methanol were added. The resulting mixture was refluxed lowing which the formation of purple formazan by the
for 1 h and the green precipitate was filtered and dried. Ele- metabolically active cells was quantified at 595 nm using
H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96 87

an ELISA reader, Thermo Electron Corporation Multiskan 800 nm with 10 or 50 nmol of Cu–Ad; however, because of
Ex multi-well spectrophotometer. possible low absorptivity at higher wavelength and lim-
ited solubility, Cu–Ad spectrum at higher concentration
2.2.3. Reactive oxygen species (ROS) measurement (2.2 ␮mol) was performed by adsorbing it on a filter paper
The intracellular ROS generation was measured as saturated with 1 M Tris buffer pH 7.0, in the absence and
described [22] using the nitroblue tetrazolium (NBT) reduc- presence of DNA (1600 ng).
tion to formazan assay. NBT, at a final concentration of
1 mg/ml, was added to the control and Cu–Ad-treated 2.3.1. Gel shift assay
(24 h) HepG2 cells (1 × 104 ) in culture. Cells were lysed and The effect of Cu–Ad on the binding of DNA probes to
formazan was dissolved with 2 M KOH and 1.4 V DMSO. nuclear factors, hence on transcription, was determined
The absorbance was monitored spectrophotometrically at using electrophoretic mobility shift assay (EMSA). In this
630 nm. experiment, we investigated the effect of Cu–Ad on the
binding of GATA-5 to 32 P-labeled double-stranded oligonu-
2.2.4. Intracellular ATP level cleotides harboring the GATA-5 consensus binding motifs.
The intracellular ATP level was determined in control
and Cu–Ad-treated cells following the instruction manual 2.3.2. Preparation of DNA probe
of the ATP Bioluminescence Assay Kit (Roche). In 12-well The double-stranded primer of GATA-5 was phosphory-
plates, seeded HepG2 cells (5 × 104 per well) were treated lated using T4 kinase that adds 32 P on the 5 -end. Briefly,
for 24 h with different concentrations of Cu–Ad. The cells the phosphorylation reaction contained in a final volume
were then collected, lysed using the kit lysis reagent, then of 5 ␮l: 2 ␮l of the annealed GATA-5 primers {5 pmol, of
50 ␮l were transferred to white MTP plates followed by each of the forward: 5 -AGCCAGAGATAAGACCCG-3 ; and
the addition of the luciferase reagent (50 ␮l). ATP biolu- reverse: 5 -CGGGTCTTATCTCTGGCT-3 primers}, 1 ␮l of T4
minescence was measured, compared to a control, as per kinase, 1 ␮l kinase buffer (10×), 20 ␮l of 32 P-␥ATP and 1 ␮l
kit instruction manual using the Luminometer; Ascent FL, of water; all were mixed and incubated for 1 h at 37 ◦ C.
Fluroskan-Thermo Lab Systems. The probe was then loaded on 12% bis-acrylamide gel in
TBE 1× buffer (0.1 M Tris:0.1 M borate:2 mM EDTA pH 8.00)
2.3. Binding of Cu–Ad to DNA at 110 V for 30 min. The pure double-stranded probe was
located after exposure to an XOMAT film and cut. The acry-
The binding of Cu–Ad to genomic DNA was assessed by lamide gel containing the probe was incubated overnight
monitoring the change in absorption of Cu–Ad complex at in 400 ␮l TES buffer (Tris:EDTA:NaCl) at 37 ◦ C with agita-
260 nm in the presence or absence of genomic DNA. tion. The content of the tube was then added to a quick
Genomic DNA was prepared following standard pro- spin column (Costar) and centrifuged for 10 s at maximum
cedures. Blood sample (5 ml) was collected, added to speed. The probe was precipitated in 2.5× volume of probe
red blood cell lysis buffer (6 ml RCLB) and incubated at in 1 ml of ethanol, i.e. about 1 ml ethanol (100%) and cen-
37 ◦ C for 10–12 min. The mixture was then centrifuged trifuged at 4 ◦ C at 14,000 rpm for 15 min. The supernatant
for 2 min at 4000 rpm and the supernatant was decanted. was discarded and the pellet was washed with 700 ␮l of
To the pellet, RCLB (2 ml) was added, incubated again 70% ethanol and centrifuged at 14,000 rpm for 10 min. The
for 10–12 min at 37 ◦ C and re-centrifuged for 2 min at supernatant was discarded and ethanol was allowed to air
4000 rpm, the supernatant was discarded and the white dry, then finally suspended in 50 ␮l TE (1×). The radioactiv-
cells pellet was re-suspended in white cell lysis buffer ity of the probe was measured by adding 2 ␮l of the probe
(1.8 ml) containing sodium dodecyl sulfate (24 ␮l SDS 10%) to 4 ml scintillation liquid and counted using a ␤ counter.
and proteinase K (18 ␮l, 20 mg/ml). The suspension was
incubated 2 h at 55 ◦ C (or overnight at 37 ◦ C) following 2.3.3. Assembly for gel casting: sample preparation
which sodium chloride (3.3 mM) was added, vortexed and The EMSA unit was assembled using 6% polyacrylamide
centrifuged for 15 min at maximal speed. The supernatant gel. A total volume of 100 ml of the gel was prepared
was transferred into a tube to which 3 ml of absolute by adding 19 ml of acrylamide:bis-acrylamide (29:1) to
ethanol were added. The DNA flocculent was transferred 2.5 ml of 10× TBE (1 M Tris base:1 M borate:20 mM EDTA,
into 1.5 ml eppendorf tube containing 200 ␮l of ethanol pH 8.00), 4 ml APS (1.6%), 74.5 ml water and finally 50 ␮l
(70%), centrifuged for 5 min at maximal speed and the TEMED. After polymerization, the gel was pre-run for
supernatant was decanted. DNA was left to dry at room 45 min at 205 V (or current intensity 15 mA) in a 0.25×
temperature, then re-suspended in TE (1×) buffer (200 ␮l TBE buffer. Sample was prepared in a final volume of 20 ␮l
of 1 mM Tris:0.1 mM EDTA). DNA concentration was containing 4 ␮l of the binding buffer (4 mM Tris:24 mM
determined using UV–vis spectrophotometer SHIMADZU KCl:0.4 mM EDTA:0.4 mM DTT:5 mM MgCl2 :glycerol 5%),
UV-1201 at wavelength of 260 nm. 1 ␮l poly-dI/dC (Amersham 1 ␮g/␮l), 3 ␮l of the inhibitor
To determine if Cu–Ad binds genomic DNA, the UV–vis of non-specific binding and the labeled probe (50,000 cpm)
spectra for each of Cu–Ad complex (10 nmol); genomic DNA then finally 3 ␮l of the nuclear extract and water. The
(800 ng); and copper–adenine:DNA (10 nmol:800 ng) mix- contents were incubated for 15 min at room temperature
ture was monitored in Tris–HCl (1 M) pH 7.0 using quartz following which samples (20 ␮l) were loaded on the gel.
micro-cuvette (1 ml). All samples were premixed, incu- Separation was carried by running the current (2–3 h) at
bated for 15 min at 37 ◦ C prior to spectral scanning. No a constant voltage (205 V). The gel was then dried using
peaks were observed in the visible region between 400 and a Biorad drying system for 90 min at 80 ◦ C under vacuum
88 H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96

and finally it was exposed and developed using X-ray film were frozen (−80 ◦ C) and thawed three times before using
(XOMAT). The effect of Cu–Ad complex at 10 and 25 nmol them in activity assay. The effect of Cu–Ad on respira-
was compared to controls of adenine, copper chloride and tory chain complexes I and II, the main entry sites of the
DMSO. reducing equivalents NADH and FADH2 , respectively, was
investigated. Activities of the rotenone sensitive NADH-UQ-
2.3.4. Effect of Cu–Ad on Taq polymerase reductase (complex I) and antimycin sensitive succinate
The effect of Cu–Ad on Taq polymerase-catalyzed dehydrogenase (complex II) were assayed spectrophoto-
gene amplification reaction was investigated. Two exons metrically [24]. The rate of absorbance change (Abs) of
namely: exon-11 of the ATP7B gene (Wilson gene) and NADH oxidation at 340 nm and K3 Fe(CN)6 reduction at
exon-1 of the transcription factor TbX20 gene were ampli- 400 nm, respectively, by mitochondria (30 ␮g) was moni-
fied in the presence and absence of the Cu–Ad complex tored with time (30 min). Mitochondria was pre-incubated
and compared to controls of DMSO, copper chloride and for 5 min with Cu–Ad (0.5–4.5 nmol) prior to initiation of
adenine. Briefly, the amplification reaction, of final vol- the reaction. The absorbance change was recorded and was
ume of 50 ␮l, contained 25 ␮l of IQ-Supermix (containing compared to a vehicle control at a final concentration of
100 mM KCl, 40 mM Tris–HCl, pH 8.4, 1.6 mM dNTPs, iTaq 1%.
DNA polymerase (Biorad) 50 units/ml, 6 mM MgCl2 ), 22 ␮l
of autoclaved H2 O, 1 ␮l of the forward and reverse primers 3. Results
(250 ng/␮l, listed below) for exon-11 of ATP7B and exon-
1 of TbX20 genes and 1 ␮l of genomic DNA (250 ng/␮l). The synthesis of various copper–adenine complexes
The amplification program was as follows: Taq activation (1–6) was achieved starting with different copper salts,
(2 min at 94 ◦ C) and 38 cycles of denaturation (30 s at 94 ◦ C), ligands and solvents. The complexes’ composition was con-
annealing (30 s at 60 ◦ C) and extension (30 s at 72 ◦ C) and a firmed by elemental analysis, infrared spectroscopy and
final extension (7 min at 72 ◦ C). The following primers were thermal analysis. A summary of the characteristic infrared
used in the amplification of E-11: ATP7B: forward: 5 -GCT spectral bands for the various complexes 1–6 are listed in
GTC AGG TCA CAT GAG TGC-3 ; reverse: 5 -CTG ATT TCC Table 1. Thermal analysis of complexes 1–6 using TG and
CAG AAC TCT TCA CAT-3 ; E-1: TbX20: forward: 5 -TGT TTC DSC mode are presented in Table 2 showing the %mass loss
GGG TCT TTG TCT CC-3 ; reverse: 5 -TGC ACA TTC ACA GCA (theoretical and experimental) and the enthalpy change
TTC AA-3 . (H) of the process.
The amplified PCR products were then separated by Complexes 1, 2, 4 and 5 (15–125 ␮M) were qualitatively
electrophoresis at 80 V, on a 2% agarose gel in TBE 1× buffer screened initially for their cytotoxic effect on different cell
and the bands were visualized by UV illumination com- lines in culture, using the trypan blue exclusion test (data
pared to molecular weight ladder. not shown). Complexes 3 and 6 were not included because
of their relative insolubility in aqueous medium; they used
2.4. Effect of Cu–Ad on mitochondrial complexes I and II to come out of solution even at low concentration. Among
the six complexes, only complex 1, Cu–Ad, demonstrated
Mitochondria were isolated from HepG2 cells as cytotoxic effects. Hence in this study we have investigated
described from isolated rat liver cells [23]. Mitochondria the in vitro biological effects of Cu–Ad only.

Table 1
Infrared spectral data (KBr) for adenine and its copper complexes (1–6)

Band assignment Adenine 1 2 3 4 5 6

(OH) , NH2 3327 3332 3390 3356 3398 3415


NH2 , 2ıNH2 3294 3186 3267 3290 3171 3300 3338
(C8–H) , (C2–H) , NH2 3118 3112 3080 3095 3111 3130 3200
ıNH2 1668 1652 1653 1650 1657 1666 1653
(C O) 1582 1578 – – 1583
(C4–C5) , (C8–N9) + ı(C8–H) 1558 1558 1556 1558 1558 1575 1544
ı(N1–H) 1489 1489 1474 1489 1480 1489 1489
ı(C2–H) , (C8–N9) + ı(C8–H) 1450 1458 1451 1458 1458 1452 1464
(N1–C6N6) 1420 1437 1415 1398 1408 1404 1425
(C5–N7–C8) 1335 1335 1337 1335 1348 1346 1344
(N9–C8) + (N3–C2) + ı(C–H) 1308 1317 1300 1305 1319 1312 1308
ring , ı(C–H) 1252 1278 1256 1272 1246 – 1281
ı(C8–H) + (N7–C8) 1232 1211 1236 1236 1215 1215 1209
ı(C8–N7) + ı(C6–N6–H) 1157 1145 1157 1151 1147 1172 1153
(C2–N3) 1125 1138 1126 1118 1110 –
NH2 1024 1020 1020 1020 1018 1083 1031
NH2 + N1–C6 939 928 934 935 937 929 949
ı(N1–C2–N3) + (C5–N7) + (N9–H) 870 866 800 860 – – 858
NH2 ring deformation 640 654 656 657 684 665 638
o.o.p. deformation N9–H 621 617 619 611 611 594 623
R +  (N9–H) 666 680 680 670 672 665 671
WNH2 + ıNH2 542 575 540 547 547 550 554
ı(C O) – 557 561 – – 579
H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96 89

Table 2
Thermal analysis data: thermogravimetry (TG) and differential scan calorimetry (DSC) for the complexes (1–6)

Complex no. and formula T (◦ C) Experimental loss Theoretical loss Expected fragment Enthalpy change,
(%mass) (%mass) H (J/g) (process)

(1) Cu2 (adenine)4 Cl4 ·2EtOH 71.76 10.96 10.21 2EtOH 251.74 (endo)
233.01 16.01 15.76 4Cl 9.31 (exo)
339.64 19.22 18.65 4N2 CH 123.47 (endo)

(2) Cu(acac)2 (adenine)·EtOH 100.35 10.93 10.38 EtOH 89.06 (endo)


207.53 42 44.7 2acac 207.25 (endo)
260–400 12.92 13.32 N2 CH2 , NH2 (endo)

(3) Cu(acac)(adenine)·H2 O 186.52 29.29 31.38 Acac 168.2 (endo)


229.4 5.64 5.7 H2 O 8.2 (endo)
316.38 12.92 13.3 N2 CH2 12.87 (endo)
394.59 120.9 (endo)

(4) Cu(adenine)2 Cl2 ·2H2 O 69.53 6.5 8.17 2H2 O 106.89 (endo)
216 7.85 8.06 Cl 11.24 (endo)
248.53 7.89 8.06 Cl 8.91 (exo)
275.76 35.96 31.3 2NH2 24.94 (endo)
378.99 2C2 N2 H 204.57 (endo)

(5) Cu(adenine)2 (BF4 )2 ·2EtOH 112.49 6.08 7.68 EtOH 18.5 (endo)
255.06 17.4 17.36 EtOH, N2 CH2 , NH2 131.82 (endo)
387.41 19.37 21.5 N2 CH2 , BF4 172.61 (endo)
450–750 42.87 45.54 BF4 , C4 N3 H3 , C4 N2 H (endo)

(6) Cu(acetate)(adenine)(SCN)·0.5CH3 OH 79.14 9.19 6.82 0.5CH3 OH 160.3 (endo)


258.46 16.41 17.48 Acetate 124.68 (endo)
331.97 14.13 17.19 SCN 92.01 (endo)
410.22 11.82 12.15 N2 CH 62.28 (endo)

Cu–Ad complex came out as a powder from the reaction confirming their presence in the Cu–Ad complex (Table 2).
mixture. Attempts to crystallize complex 1 from solvent The high temperature needed for adenine fragmentation is
systems failed which limited its structural characterization indicative of the relative stability of copper coordination to
by X-ray and mass spectroscopy. Spectral analysis of Cu–Ad adenine in the complex Cu–Ad.
(2.2 ␮mol) in Nujol revealed (data not included) the follow-
ing max at 288 nm (due to ␲–␲* of adenine ligand), 395 nm 3.1. In vitro biological effects of Cu–Ad complex
(shoulder) and 580 nm (due to d–d transition). Nujol how-
ever interacted with DNA thus could not be used to monitor 3.1.1. Cytotoxicity of Cu–Ad
the spectral changes occurring upon binding to Cu–Ad. The effect of varying Cu–Ad concentration (15–125 ␮M)
Complex 1 was stable under all biological assay condi- on different cell lines in culture was evaluated using trypan
tions; thermal analysis findings have shown (Fig. 1) a mass blue exclusion test. Our results (Fig. 2a) show a significant
loss corresponding to 2EtOH molecules at 75 ◦ C; 4Cl groups (p < 0.005) increase in cell death with increasing concentra-
at 233 ◦ C, and 4N CH–N fragments of adenine at 339 ◦ C tion of Cu–Ad. A maximal cell death of: 96% with HepG2,

Fig. 1. Thermal analysis of Cu–Ad complex using TG and DSC mode and showing %mass loss and the enthalpy change of the process.
90 H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96

Fig. 3. Representative UV-absorption spectra of Cu–Ad (yellow tracing),


Cu–Ad + DNA (blue tracing) and DNA (red tracing). Genomic DNA (800 ng)
was incubated with 10 nmol of Cu–Ad for 15 min at 37 ◦ C in 1 M Tris buffer
pH 7.0. Absorption spectrum (200–400 nm) was monitored using Shi-
madzu spectrophotometer. (For interpretation of the references to color in
this figure legend, the reader is referred to the web version of the article.)

pH 7.0, in the presence and absence of DNA (1620 ng) then


scanned between 200 and 800 nm. Cu–Ad alone exhibited
the following max at 261, 343 and 590 nm; whereas when
mixed with DNA we observed an increase in the absorbance
at 261 nm, a shift in the absorption from 343 to 402 nm and
Fig. 2. Effect of Cu–Ad on different cell lines in culture. Cells cultured
under optimal conditions of humidity and using appropriate media were a small shift from 590 to 587 nm (data not shown).
treated for 24 h with Cu–Ad (15–125 ␮M). %cell death was compared to a
control, considered as 100% treated with vehicle. The effect of Cu–Ad was 3.1.2. Cu–Ad inhibits DNA binding to transcription factor
estimated using: (a) trypan blue exclusion assay compared to a control
Nuclear extracts expressing the rat GATA proteins
treated with vehicle and (b) MTT assay performed only on HepG2 cells.
were incubated with 32 P-radio-labeled double-stranded
oligonucleotide containing GATA consensus binding motifs.
84% with C2C12, 74% with NIH3T3, 69% with MCF7, 60% The addition of 5, 10 and 25 nmol of Cu–Ad inhibited com-
with MCF7-Bcl2, 28% with A549, was obtained with final pletely GATA-5 binding to DNA (Fig. 4) whereas a similar
concentration of 125 ␮M Cu–Ad. Estimated IC50 for the dif- effect was obtained with CuCl2 , neither adenine nor DMSO
ferent cell lines were: 45 ␮M HepG2; 73 ␮M C2C12; 103 ␮M exerted any on transcription.
NIH3T3; 108 ␮M MCF7; and 115 ␮M MCF7-BcL2, indicating
that HepG2 cells are the most sensitive, while A549 cells are 3.1.3. Cu–Ad inhibits Taq polymerase activity in vitro
the least sensitive. The toxicity of Cu–Ad on HepG2 cells was The effect of Cu–Ad on replication was assessed
further verified using the MTT assay. A decrease (Fig. 2b) in indirectly via its effect on amplification by Taq polymerase-
HepG2 viability was obtained with increasing Cu–Ad con-
centration with an estimated IC50 of 40 ␮M. Neither free
adenine nor copper chloride salt, when added to HepG2
cells, at a final concentration of 30, 100 and 200 ␮M, exerted
any effect (data not shown) on cell viability indicating the
effect is due to intact Cu–Ad complex.

3.1.1.1. Cu–Ad binds genomic DNA. The interaction of the


copper–adenine complex with genomic DNA was investi-
gated by monitoring the changes in the absorption intensity
of free Cu–Ad complex compared to that of DNA pre-
incubated with Cu–Ad. The UV spectra of Cu–Ad showed a
single peak at 260 nm characteristic of the aromatic ade-
nine base. Incubation of Cu–Ad complex (10 nmol) with
genomic DNA (800 ng/ml) isolated from human blood,
decreased the absorption intensity by 30%; the yellow and
blue tracings in Fig. 3 are for Cu–Ad complex alone and
Cu–Ad complex with DNA, respectively. No other peaks at
higher wavelength were observed at the Cu–Ad concentra-
tions (10 and 50 nmol) used in our study. To examine if spec-
Fig. 4. Representative figure of the effect of Cu–Ad on gel shift assay. Dif-
tral changes, at higher wavelength, occur upon the binding
ferent concentrations of Cu–Ad were added to nuclear extracts treated
of Cu–Ad to DNA, high amount of Cu–Ad (2.2 ␮mol) were with GATA-5 32 P-labeled DNA probe, then separated, exposed and com-
adsorbed on a filter paper saturated with 1 M Tris buffer, pared to CuCl2 , adenine and DMSO (vehicle) as described in Section 2.
H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96 91

Fig. 5. Representative figure of the effect of Cu–Ad on Taq polymerase-catalyzed reaction. The effect of Cu–Ad on the amplification of exon-11 of the ATP7B
(Wilson gene) and exon-1 of the TbX20 gene was investigated and compared to CuCl2 , adenine and DMSO.

catalyzed reactions in vitro. The amplification of exon-11


of the Wilson gene (WND, ATP7B) and exon-1 of the
TbX20 gene was examined. The complex was added to the
PCR reaction with the appropriate forward and reverse
primers for each of the amplified exons. A concentration-
dependant decrease in the intensity of the amplified bands
in both WND (54%) and TbX20 (68%) genes (Fig. 5) was
observed with 0.5 nmol of Cu–Ad complex. Similar results
were obtained when Cu–Ad complex was subjected to
the PCR program, heating–cooling cycles and prior to
adding it to the Taq polymerase amplification reaction
indicating thus its stability under the reaction conditions.
On the other hand CuCl2 (0.5 and 1 nmol) inhibited the Taq
polymerase reaction whereas adenine and DMSO had no
effect.

3.1.4. Effect of Cu–Ad on mitochondrial complexes


Copper–adenine effect on mitochondrial function
was assessed via testing its effect on the activity of
both complexes I and II, the main entry site of NADH
and FADH2 , respectively. Our results (Fig. 6a) show a
concentration-dependent decrease in the activity of com-
plex I reaching a maximal inhibition (74.8%) at 4.0 nmol
Cu–Ad, with an estimated IC50 of 2.8 nmol. On the other
hand, Cu–Ad (1–4 nmol) exerted no effect on complex II
activity (Fig. 6a). Control enzyme activities of 104 ± 10.3
and 94.8 ± 8.4 abs/(mg/h) (considered as 100%) were for
complex I (NADH-UQ-reductase) and complex II (succinate
dehydrogenase), respectively, treated with the vehicle of
Cu–Ad.

3.1.5. Cu–Ad increases ROS generation and decreases


intracellular ATP level
Inhibition of complex I by Cu–Ad will decrease
electron transfer through the electron transport chain;
consequently will decrease ATP level and increase ROS gen-
eration. We examined the level of ATP in Cu–Ad-treated
HepG2 cells. Our results show 80% decrease in ATP level at Fig. 6. The effect of Cu–Ad on (a) NADH-UQ-reductase () and
50 ␮M Cu–Ad (Fig. 6b). succinate dehydrogenase () activities in mitochondria (30 ␮g) iso-
Similarly Cu–Ad decreased NBT reduction into for- lated from HepG2 cells. Activities were compared to a control
value (100%) of 104 ± 10.3 abs/(mg h) for NADH-UQ-reductase and
mazan. Our results (Fig. 6c) show a decrease in NBT
94.8 ± 8.4 abs/(mg h) for succinate dehydrogenase treated with vehicle.
reduction with increasing concentration of Cu–Ad, indicat- (b) ATP bioluminescence assay as per supplier instructions and (c) NBT
ing an increase in ROS generation. A 60% increase in ROS reduction into formazan. Results are average ± S.E.M. of three determina-
was obtained at 45 ␮M Cu–Ad. tions from three different experiments.
92 H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96

4. Discussion

In this study the toxicity and biochemical targets of


a copper coordination complex synthesized with adenine
as ligand, were assessed in vitro. The compound Cu–Ad
induced cell death in many cell lines and bound genomic
DNA. In addition it inhibited Taq polymerase-catalyzed
reactions, the binding of transcription factor to DNA and
mitochondrial NADH oxidation in vitro. Cu–Ad increased
also ROS generation and decreased ATP level that may be
one of the possible underlying mechanisms of its cytotox-
icity. These findings identify Cu–Ad as a compound with
several biochemical targets and potential anti-cancerous
activity.
Copper complexes 1–6 were synthesized (Scheme 1)
from different ligands and solvents and were initially quali-
tatively (data not shown) screened for their cytotoxic effect
on different cell lines in culture. Out of the six complexes,
only complex 1 (Cu–Ad), induced cell death in all the tested
cell lines, except for A549 (Fig. 2a). The most sensitive were
HepG2 cells showing a dose-dependant effect (Fig. 2b).
Elemental analysis, comparative IR spectra and thermal
analysis profile of complexes 2–6 verified the structural
components of Cu–Ad.
Several bands appeared in the IR spectra of adenine
and its copper complexes (Table 1). Bands in the region
3500–3000 cm−1 could be attributed to stretching vibra-
tion for –NH group of adenine and OH group of water,
ethanol or methanol which are present in the lattice of
the complexes [25,26]. The peak at 3390 cm−1 could be
attributed to OH of water in complex 3 and at 3415 cm−1
to OH of methanol in complex 6. The carbonyl stretch of Fig. 7. Crystal structure of Cu(acac)2 (adenine)·EtOH compound (2).

acetylacetone is clearly observed at 1582 and 1578 cm−1 in


complexes 2 and 3, respectively, while the carbonyl stretch cation also coordinates through ring nitrogen N(9)
of the acetate appeared at 1583 cm−1 in complex 6. in [Cu(adenineH)2 Cl2 ]2+ [33]. On the other hand, the
The NH2 stretches of adenine at 3290–3118 cm−1 were crystal structure of [Cu(MOBIDA)(adenineH)(H2 O)]. H2 O
slightly shifted upon adenine complexation with metal reveals the selective formation of rare Cu–N(3) adenine-H
(Table 1). The ıNH2 mode of free adenine at 1668 cm−1 bound, (MOBIDA = N-(p-methoxybenzyl)-imidiacetato(2-))
undergoes shifts to about 1650 cm−1 as expected for cor- ligand [33]. In a recent X-ray study the structure of
responding N-bounded adenine complexes [25–27]. Thus Cu(acetylacetone)2 (adenine)·EtOH shows that the complex
NH2 group (exocyclic NH2 nitrogen) is coordinated to the is square pyramidal with acetylacetone ligands occupying
metal ion in the complexes 1–4 and 6. Adenine coordi- the equatorial positions and adenine in the axial position
nate through ring nitrogen with appreciable shifts and coordinating through its N(7) nitrogen [34,35] (Fig. 7).
occasional splitting of C C, C N and ring vibrations of Further evidence on the formula of adenine–metal
the ligand (1605–1300 cm−1 ) [28–30]. The NH region, complexes comes from thermal analysis (Table 2). The
2900–2500 cm−1 , suffered considerable changes in the %mass loss with expected fragments and involved heat dur-
metal adenine complexes resulting in two weak maxima in ing thermal analysis TG–DSC of various complexes (1–6)
this region [26]. The 1232 cm−1 of adenine due to N7–C8 are listed in Table 2. Thermal analysis of free adenine
shift to lower wavelength upon complexation in the case shows the presence of two decomposition steps: one cen-
of complex 1 and 4–6 indicating the binding of adenine is tered at 383 ◦ C and the second at 667 ◦ C, with associated
through ring nitrogen [25,29]. endothermic heat of 892 and 376 J/g due to loss of C4 N3 H3
Various coordinated sites have been observed for ade- and N2 CH2 , respectively. Coordinated adenine although
nine in copper complexes as indicated by X-ray studies: showed two degradation steps in all the studied complexes
among the four nitrogens N(1), N(3), N(7) and N(9) of ade- at similar temperature regions 250–400 ◦ C and 400–800 ◦ C,
nine, the N(9) is the most basic hence bears a proton ren- the %mass loss shifts to lower values in the first step and to
dering it the most preferred metal binding site. An example larger values in the second step. The binding of N-adenine
is the crystal structure of [Cu(tren)(adeninato)]ClO4 to copper introduces some sort of rigidity to the ring struc-
where N(9) of adeninate anion is bound to Cu at ture that degrades in the second step with increased %mass
the axial position {tren = Tris(2-aminoethyl)amine} [31]. loss. While N CH–NH of five-member ring of adenine is
The same mode of binding is observed in the crystal lost in the first step in most cases, TG–DSC curve for com-
structure of [Cu(tren)(adeninato)]Cl·2H2 O [32]. Adenine plex 1 is shown in Fig. 1. In addition thermal analysis of
H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96 93

Fig. 8. Proposed structure of Cu–Ad (1) with formula [Cu(adenine)2 Cl]2 Cl2 ·2EtOH when L = Cl− , and [Cu(adenine)2 (EtOH)]2 Cl4 when L = ethanol.

copper bis(acetylacetone) shows a sharp endothermic peak transformed with time. However no effect on cell viability
at 200 ◦ C (H = 465 J/g). The same pattern is observed in was obtained when cells were treated with either CuCl2
complexes 2 and 3 which clearly confirm the presence of or free adenine; thus indirectly ruling out the dissociation
acetyl acetone in both. of Cu–Ad into its primary components. We hence opted to
Previous studies have investigated the interaction of further investigate the target Cu–Ad in vitro.
Cu(II) with adenine in solution. These studies reported Coordination metal complexes among which is copper,
that copper adeninate complexes were stable involving N3 have been designed and used in the treatment of many
and N9 as binding sites simultaneously [36]. The binding diseases including cancer. Of the many suggested mech-
constants for copper and related metal complexes have anisms is the possible interaction with DNA and disruption
indicated the high stability of these complexes [36–38]. of its tertiary structure. This may influence the recognition
In this study the prepared Cu–Ad complex remained sta- of DNA by specific factors or proteins leading ultimately to
ble in DMSO. No decomposition or change in its spectral inhibition of DNA synthesis and or transcription [2].
properties or thermal data was obtained even after a In this study we show (Fig. 3) a decrease in the
long time exposure to atmosphere. The crystal structure absorbance of Cu–Ad compared to that pre-incubated with
(Fig. 7) of complex 2, [34] has shed some light on ade- genomic DNA, suggesting binding of the complex to the
nine binding to copper in Cu–Ad; both complexes 1 and DNA. On the other hand the observed increase in the
2 were prepared similarly; with one exception the addi- absorbance of the Cu–Ad–DNA (blue tracing) compared to
tion of ammonium chloride to Cu–Ad. A dimeric structure that of genomic DNA (red tracing) may result from dis-
is proposed for Cu–Ad with four bridging adenine and ruption of the double-stranded DNA into single strand by
chloride or ethanol as axial ligand (L) suggesting the fol- Cu–Ad. A shift in the structure of DNA from double stranded
lowing formulae, respectively [Cu(adenine)2 Cl]2 Cl2 ·2EtOH into single is known to increase the absorbance. Alterna-
or [Cu(adenine)2 (EtOH)]2 Cl4 (Fig. 8). Previous investiga- tively, possible base pairing, favored by ␲–␲ interactions,
tors have reported with related complexes [39,40] similar between the adenine moieties of the Cu–Ad complex and
dimeric structures involving bridging of adenine through the thymine bases of the genomic DNA would facilitate
its N3 and N9 with the axial sites occupied by water or the separation of the double-stranded DNA favoring single-
chloride ion. stranded chains, subsequently increasing absorbance [18].
Of the entire complexes (1–6) only Cu–Ad complex Molecules that modify nucleic acids have received con-
significantly decreased cell viability in a concentration- siderable attention because of their potential application
dependant manner. HepG2 and C2C12 were the most as chemotherapeutic agents. Transition metal complexes
sensitive whereas A549 was the least sensitive (Fig. 2a). endowed with redox properties and DNA affinity have
The obtained effects indicate that Cu–Ad is bioavailable, been developed as chemical nucleases. DNA cleavers such
capable of overcoming the physical barriers and entering as iron–bleomycin, Mn(III)–porphyrin, nickel, Co, Rh, and
the cells. Prolonged treatment (data not shown) of HepG2 Ru phenanthroline/bipyridine and Cu(phen)2 2+ complexes
with Cu–Ad for 36 and 48 h resulted in less cell death when were able to mediate oxidative damage to nucleobases/or to
compared to 24 h, which may be attributed to proliferation deoxyribose moiety [41]. Among the many possible mech-
of HepG2, instability of the Cu–Ad or possible metabolism anisms of metal carcinogenesis, is the direct or indirect
of Cu–Ad into less toxic compound. Our experiments do binding of the metal ion to DNA. The direct binding to
not provide direct evidence, as to whether Cu–Ad gets bio- DNA involves the binding of the metal to either phosphates
94 H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96

and/or nucleobases. The coordination of the divalent metal through oxidative scission or via hydrolytic pathway.
ion to N7 of guanine in DNA promotes the proton switch Induced DNA strand scission has been reported to occur
of the hydrogen from guanine to cytosine, and increases by reactive radical species generated in the presence of
the probability of mispairing hence mutations [42]; similar transition metal ion catalysts and redox [50]. Of all essen-
situation may be occurring with the Cu–Ad complex. tial metallo elements, copper has the highest bonding
The ability of copper II complexes to interact with DNA affinity for nucleic acids. Nucleic acid phosphate and base
and induce strand scission was reported [43]. Copper- donor sites yield Cu chelates that are much more sta-
adenylated polymer templates resulted in the complete ble than Cu chelates of most amino acids and near the
conversion of supercoiled into nicked DNA. The above stability of protein chelates. A higher physiological concen-
finding prompted us to investigate the in vitro effect of tration accounts for inhibition of ribonuclease syntheses:
Cu–Ad complex on gene amplification by Taq polymerase initiation, nucleotides condensations and release of newly
(indirectly replication) and on the binding of transcrip- formed mRNA [51].
tion factors to DNA probes (indirectly transcription). Cu–Ad Contrary to Cu–Ad or complex 1, monomeric complexes
complex inhibited both. such as complex 2 (Fig. 7), complex 4 and complex 5 did
Taq polymerase is an enzyme that catalyzes the in not exert any cytotoxic effect. The proposed dimeric struc-
vitro amplification of gene/exon by polymerizing deoxynu- ture of Cu–Ad in which bridging of adenine to two copper
cleotides. Cu–Ad inhibited significantly the amplification atoms occurs via N3 and N9, would leave behind N1 and
of exon-11 of the Wilson gene, a copper ATPase encoding N6 available for H-bonding with thymine base of DNA.
gene. To rule out the effect is because it is a copper-related This substitutes the normal hydrogen bonding that exists
gene, exon-1 of TbX20 gene was amplified in the presence between adenosine and thymine in double-stranded DNA,
of Cu–Ad (Fig. 5). Again the Taq polymerase activity was contributing thus to some of the cytotoxic effects imparted
significantly inhibited by Cu–Ad. Neither adenine nor the by Cu–Ad [52].
vehicle (DMSO) demonstrated any effect. CuCl2 however Mitochondria are one of the subcellular targets of metal-
exhibited a similar effect, suggesting inhibition is proba- based toxicity drugs. For instance severe morphological
bly mediated by the copper. When subjecting the Cu–Ad alterations were reported to occur following cisplatin treat-
alone to the PCR-programmed cycle of heating and cool- ment to isolated or inside intact kidney cells. These include
ing prior to adding it to the Taq polymerase amplification mitochondrial swelling, collapse of membrane potential,
reaction, we obtained the same result (data not shown) inhibition of ADP stimulated respiration and increased for-
indicating that Cu–Ad complex is stable under these con- mation of reactive oxygen species [15,53]. Mitochondria
ditions. In addition thermal analysis data have shown the were also suggested to be involved in the maintenance
presence of two-step decomposition of the adenine at tem- of malignant phenotype by unknown mechanisms related
perature greater than 200 ◦ C. Currently our data do not to ATP production [54]. In tumorogenesis the role of
indicate if this inhibition is the result of a direct inhibition of mitochondria may involve the transfer and insertion of
Cu–Ad on the Taq polymerase or is an effect of the Cu–Ad mitochondrial DNA into nuclear DNA or altering the expres-
on DNA. Cu(II) complexes were reported to cleave inacti- sion of mt-DNA encoded proteins.
vated peptide bonds and the thermostable enzyme Taq DNA Cu–Ad inhibited NADH-UQ-reductase activity in iso-
polymerase at ␮M concentration under mild pH and tem- lated HepG2 mitochondria but had no effect on succinate
perature conditions [42,44]. Strand breaks and modified dehydrogenase activity (Fig. 6a). NADH-UQ-reductase or
bases were induced in vitro by copper ion-mediated reduc- complex I of the respiratory chain complex, is the main
tion of hydrogen peroxide in the presence of ascorbate entry site of the reducing equivalent NADH whose oxi-
[45]. Copper and copper(I)-metalated nucleobase polymers dation is coupled to ATP production. Cu–Ad (2.8 nmol)
induced DNA cleavage and promoted hydrolysis of non- inhibited 50% of the rotenone sensitive NADH oxidation of
natural and natural phosphate ester substrates [46,47]. In the ETC, which is in line with reported toxic concentration
addition Bruston et al. reported hydrogen peroxide dis- of the copper-based antineoplastic drugs, the casiopeinas
proportionation into oxygen and water by the catalytic [15]. Inhibition of NADH oxidation would increase oxy-
activity of copper–adenine complexes [48]. DNA cleavage gen radical generation, deplete intracellular ATP level
and complete conversion of supercoiled plasmid DNA into and ultimately lead to cell death. Alternatively inhibition
linearized DNA was reported to occur by the catalytic activ- of complex I would increase oxygen radical generation
ity endowed to copper complexes in the presence of a favoring oxidative stress and leading to DNA damage by
co-oxidant [46]. modifying bases or deoxyriboses [18]. In this study we show
In a previous study phenanthroline Cu(I) complex inhib- an increase in ROS generation and a decrease in ATP pro-
ited transcription by binding to a site generated by the duction (Fig. 6b and c) which may be one of the mechanisms
interaction of the RNA polymerase with the promoter that underlying the demonstrated toxicity of Cu–Ad.
may partly be attributed to its potent cytotoxicity [49]. We In summary, we report in this paper the synthesis of
hereby show inhibition (Fig. 4) of the binding of the tran- a copper–adenine complex, a compound with multi in
scription factor GATA-5 to DNA probe (harboring GATA-5 vitro biochemical effects. The Cu–Ad complex was relatively
consensus site) by Cu–Ad. Likewise neither adenine nor toxic and was found to inhibit Taq polymerase-catalyzed
DMSO had any effect whereas CuCl2 did. reaction, the binding of transcription factors to DNA probes
Among transition metal ions, copper-based artificial and mitochondrial NADH oxidase activity. The different
systems have received considerable focus because of their multi-biochemical targets and effects of Cu–Ad identify it
dual mode of action in modifying nucleic acids either as a potential anti-cancerous compound. Design of different
H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96 95

structural analogues of Cu–Ad will allow better evalua- [22] S.Y. Paik, K.H. Koh, S.M. Beak, S.H. Paek, J.A. Kim, The essential oils
tion of structure–function relationship and increases target from Zabthoxylum Schinifoliukm pericap induce apoptosis of HepG2
human hepatoma cells through increased production of reactive oxy-
specificity and potency. gen species, Biol. Pharm. Bull. 28 (5) (2005) 802–807.
[23] P. Gellerfors, B.D. Nelson, A rapid method for the isolation of intact
mitochondria from isolated rat liver cells, Anal. Biochem. 93 (1979)
References
200–203.
[24] J. Usta, S. Kreydiyyeh, K. Bajakian, H. Nakkash-Chmaisse, In vitro
[1] A.Y. Louie, T.J. Meade, Metal complexes as enzymes inhibitors, Chem. effect of eugenol and cinnamaldehyde on membrane potential and
Rev. 99 (1999) 2711–2734. respiratory chain complexes in isolated rat liver mitochondria, Food
[2] R.N. Patel, N. Singh, K. Shukla, U.K. Chauhan, S. Chakraborty, J. Niclos- Chem. Toxicol. 40 (7) (2002) 935–940.
Gutierrez, A. Castineiras, X-ray, spectral and biological (antimicrobial [25] I. Somasundaram, M. Palaniandavar, Models for enzyme–copper–
and superoxide dismutase) studies of oxalate bridged Cu-II com- nucleic acid interaction: interaction of some biomimetic copper com-
plexes with penta methyldiethylenetriamine as capping ligand, J. plexes derived from salicylaldehyde, glycine and ␣-alanine with
Inorg. Biochem. 98 (2004) 231–323. adenine and adenosine, Indian J. Chem. 32A (1993) 495–499.
[3] A.H. Fairlamb, G.B. Henderson, A. Cerami, Trypanothione is the pri- [26] R. Singh, S. Tyagi, S. Singh, S.M. Singh, U.P. Singh, Solid, solution and
mary target for arsenical drugs against African trypanosomes, Proc. antitumor activity studies of mixed ligand complexes with adenine-
Natl. Acad. Sci. A86 (1989) 2607–2611. 5-bromouracilbase pair, Synth. React. Inorg. Met. Org. Chem. 32 (5)
[4] Z. Balcarova, J. Kasparakova, A. Zakovska, O. Novakova, M.F. Sivo, G. (2002) 853–859.
Natile, V. Brabeck, DNA interactions of a novel platinum drug, cis- [27] M.S. Masoud, A.A. Soayed, A.E. Ali, Complexing properties of nucleic-
[PtCl(NH3 )2 (N7-Acyclovir)]+ , Mol. Pharmacol. 53 (1998) 846–855. acid constituents adenine and guanine complexes, Spectrochim. Acta
[5] R.L. LaFemina, Requirement of active human immunodeficiency 60 (2004) 1907–1915.
virus type 1 integrase enzyme for productive infection of human [28] R. Savoie, J.J. Jutier, L. Prizant, A.L. Beauchamp, Raman and infrared
T-lymphoid cells, J. Virol. 66 (1992) 7414–7419. spectra of methylmercury complexes of adenine, Spectrochim. Acta
[6] P.S. Moore, C.J. Jones, N. Mahmood, I.G. Evans, M. Goff, R. Cooper, A: Mol. Spectrosc. 38 (1982) 561–566.
A.J. Hay, Anti-(human immunodeficiency virus)activity of polyoxo- [29] A.N. Speca, C.M. Mikulski, F.J. Iaconianni, L.L. Pytlewski, N.M.
tungstates and their inhibition of human immunodeficiency virus Karayanni, Characterization studies of adenine complexes with 3d
reverse transcriptase, Biochem. J. 307 (1995) 129–134. metal perchlorates, J. Inorg. Nucl. Chem. 43 (11) (1981) 2771.
[7] O. Rixe, W. Ortuzar, M. Alvarez, R. Parker, E. Reed, K. Paull, T. Fojo, [30] A.N. Speca, L.L. Pytlewski, C.M. Mikulski, N.M. Karayanni, Zinc(II)
Oxaliplatin, tetraplatin, cisplatin and carboplatin: spectrum of activ- complexes of adenine, Inorg. Chim. Acta 66 (1982) 153–158.
ity in drug-resistant cell lines and in the cell lines of the National [31] M.A. Salam, K. Aoki, Metal ion interactions with nucleobases in the
Cancer Institute’s Anticancer drug screen panel, Biochem. Pharmacol. tripodal Tris(2-aminoethyl)amine (tren) ligand system, Inorg. Chim.
52 (1996) 1855–1865. Acta 314 (2001) 71–78.
[8] R. Bakhtiar, E.-I. Ochiai, Pharmacological applications of inorganic [32] A. Marzotto, A. Ciccarese, D.A. Clemente, G. Valle, Co-ordination
complexes, Gen. Pharmacol. 32 (1999) 525–540. chemistry of adenine (hade): synthesis and characterization of
[9] J.H. Murray, M.M. Harding, Organometallic anticancer agents: the [Cu(tren)(nucleobase)]X2 , J. Chem. Soc., Dalton Trans. (1995)
effect of the central metal and halide ligands on the interaction of 1461–1467.
metallocene dihalides Cp2Mx2 with nucleic acids constituents, J. [33] D.B. Brown, J.W. hall, H.M. Helis, E.G. Walton, D.J. Hodgson, W.E. Hal-
Med. Chem. 37 (1994) 1936–1941. field, The preparation and magnetic and structural examination of
[10] J.S. Thayer, Organometallic Compounds and Living Organisms (1984), monomeric, dimeric and polymeric adenine complexes of copper(II),
Monographs (Organometallic Chem. Ser.), Elsevier Science and Tech- Inorg. Chem. 16 (11) (1997) 2675–2682.
nology Books, San-Diego, CA, USA, 1984. [34] M.T. Zaworotko, H.H. Hammud, G. Mc-Manus, A.T. Kabbani, M.S.
[11] V.A. Shlykhovenko, V.A. Milinevskaya, I.L. Zagorujko, Y.V. Yanish, V.V. Masoud, A.M. Ghannoum, deposited at Cambridge Crystallographic
Kozak, V.N. Kokozey, A. Schernova, Exp. Oncol. 19 (1997) 348–358. Data Center # CCDC-281077, 2005.
[12] R. Gopalakrishna, Z.H. Chen, U. Gundimeda, Seleno compounds [35] T.A. George, H.H. Hammud, S. Isber, Structure, thermal analysis
induce a redox modulation of protein kinase C in the cell, com- and magnetism of the doubly bridged linear polymer [Ni(4-NH2 -
partmentally independent from cytosolic glutathione: its role in C6 H4 CO2 )2 (CH3 OH)2 ]·CH3 OH, Polyhedron 25 (2006) 2721–2728.
inhibition of tumor promotion, Arch. Biochem. Biophys. 348 (1997) [36] R. Tauler, M.J.A. Rainer, B.M. Rode, The interaction of H+ , Zn(II) and
37–48. Cu(II) with adenine and 9-methyladenine, Inorg. Chim. Acta 123
[13] R. Gopalakrishna, U. Gundimeda, Z.H. Chen, Cancer preventive seleno (1986) 75–82.
compounds induce a specific redox modification of cysteine-rich [37] S. Satyanarayana, K. Venugopal Reddy, Potentiometric determina-
regions in Ca(2+)-dependant isoenzymes of protein kinase-c, Arch. tion of the stability constants of ternary complexes of biologically
Biochem. Biophys. 348 (1997) 25–36. important metal ions, Indian J. Chem. 28A (1989) 630–632.
[14] E. Monti, Interaction of Cu-II anthracyclin complexes with protein [38] H. Demir, M. Pekin, A.K. CuCu, E. Dolen, H.Y. Aboul-Enein, Potentio-
kinase-C. Spectromagnetic assessment of inhibitory effect, Inorg. metric studies of mixed complexes of cobalt(II) and copper(II) with
Chim. Acta 205 (1993) 181–184. l-asparagine and adenine, Toxicol. Environ. Chem. 71 (1999) 357–367.
[15] A. Marin-Hernandez, I. Gracia-Mora, L. Ruiz-Ramirez, R. Moreno- [39] A. Terzis, A.L. Beauchamp, R. Rivest, Crystal and molecular struc-
Sanchez, Toxic effects of copper based antineoplastic drugs ture of tetra-u-adenine-diaqua-dicopper(II) perchlorate dihydrate,
(casiopeinas) on mitochondrial function, Biochem. Pharmacol. 65 [Cu2 (C5 H5 N5 )4 (H2 O)2 ](ClO4 )4 ·2H2 O, Inorg. Chem. 12 (5) (1973)
(2003) 1979–1989. 1166–1170.
[16] R. Martinez, L. Chacon-Garcia, The search of DNA-intercalators as [40] P. De Meester, A.C. Skapski, Crystal structure of dichlorotetra-p-
antitumoral drugs: what it worked and what did not work, Curr. Med. adenine-dicopper(II) chloride hexahydrate, J. Chem. Soc. (1971)
Chem. 12 (2005) 127–151. 2167–2169.
[17] X. Shui, M.E. Peek, L.A. Lipscomb, A.P. Wilkinson, L.D. Williams, M. [41] M. Pitie, C.J. Burrows, B. Meunier, Mechanisms of DNA cleavage by
Gao, C. Ogata, B.P. Roques, C. Garbay-Jaureguiberry, Effects of cationic copper complexes of 3-clip-phen and of its conjugate with a dis-
charge on three-dimensional structures of intercalative complexes: tamycin analogue, Nucleic Acids Res. 28 (2000) 4856–4864.
structure of a bis-intercalated DNA complex solved by mad phasing, [42] P.A. Ochoa, M.I. Rodriguez-Tapiador, S.S. Alexandre, C. Pastor,
Curr. Med. Chem. 7 (2000) 59–71. F. Zamora, Structural models for the interaction of Cd(II) with
[18] T. Theophanides, J. Anastassopoulou, Copper and carcinogenesis, Crit. DNA: trans-[Cd(9-rgh-N7)2 (H2 O)4 ]2+ , J. Inorg. Biochem. 99 (2005)
Rev. Oncol. Hematol. 42 (2002) 57–64. 1540–1547.
[19] T.F. Kagawa, B.H. Geierstanger, A.H. Wang, P.S. Ho, Covalent modifi- [43] S. Verma, S.G. Srivatsan, C.A. Claussen, E.C. Long, DNA strand scission
cation of guanine bases in double-stranded DNA. The 1.2- a z-DNA by a Cu-I adenylated polymeric template: preliminary mechanis-
structure of d(cgcgcg) in the presence of CuCl2 , J. Biol. Chem. 266 tic and recycling studies, Bioorg. Med. Chem. Lett. 13 (2003) 2501–
(1991) 20175–20184. 2504.
[20] R. Mital, G.M. Shah, T.S. Srivastava, R.K. Bhattacharya, The effect of [44] M.C. De-Oliveira, M. Scarpellini, A. Neves, H. Terenzi, A.J. Bortoluzzi,
some new platinum(II) and palladium(II) coordination complexes on B. Szpoganics, A. Greatti, A.S. Mangrich, E.M. de Souza, P.M. Fernan-
rat hepatic nuclear transcription in vitro, Life Sci. 50 (1992) 781–790. dez, M.R. Soares, Hydrolytic protein cleavage mediated by unusual
[21] K. Maskos, The interaction of metal ions with nucleic acids. NMR mononuclear copper(II) complexes: X-ray structures and solution
study of the copper(II) interaction with inosine derivatives, Acta studies, Inorg. Chem. 44 (2005) 921–929.
Biochim. Pol. 28 (1981) 317–335.
96 H.H. Hammud et al. / Chemico-Biological Interactions 173 (2008) 84–96

[45] H. Rodriguez, R. Drouin, G.P. Holmquist, T.R. O’Connor, S. Boiteux, [49] A. Mazumder, D.M. Perrin, K.J. Watson, D.S. Sigman, A transcription
J. Laval, J.H. Doroshow, S.A. Akman, Mapping of copper/hydrogen inhibitor specific for wound DNA in RNA polymerase promoter open
peroxide-induced DNA damage at nucleotide resolution in human complexes, Proc. Natl. Acad. Sci. 90 (1993) 8140–8144.
genomic DNA by ligation-mediated polymerase chain reaction, J. Biol. [50] S.G. Srivatsan, Modeling prebiotic catalysis with nucleic acid-like
Chem. 270 (1995) 17633–17640. polymers and its implications for the proposed RNA world, Pure Appl.
[46] S.G. Srivatsan, M. Parvez, S. Verma, Modeling of prebiotic catalysis Chem. 76 (2004) 2085–2099.
with adenylated polymeric templates: crystal structure studies and [51] J.R. Sorenson, Cu, Fe, Mn, and Zn chelates offer a medicinal chemistry
kinetic characterization of template assisted phosphate ester hydrol- approach to overcoming radiation injury, Curr. Med. Chem. 9 (2002)
ysis, Chem. Eur. J. 8 (22) (2002) 5184–5191. 639–662.
[47] S.G. Srivatsan, M. Parvez, S. Verma, Adenine–copper coordi- [52] M.E. Prater, D.J. Mindiola, X. Ouyang, K.R. Dunbar, A quadruply-
nation polymer as an oxidative nucleozyme: implications for bonded dirhenium complex bridged by 2 N1 /N6 adenate ligands,
simple prebiotic catalytic units, J. Inorg. Biochem. 97 (2003) 340– Inorg. Chem. Commun. 1 (1998) 475–477.
344. [53] J.A. Gordon, V.H. Gattone, Mitochondrial alterations in cisplatin-
[48] F. Bruston, J. Verbne, L. Grajcar, B. Drahi, R. Calvayrac, M.-H. Baron, induced acute renal failure, Am. J. Physiol. 250 (1986) F991–F998.
M.-C. Maurel, Copper–adenine catalyst for O2 production from H2 O2 , [54] J.S. Penta, F.M. Johnson, J.T. Wachsman, W.C. Copeland, Mitochondrial
Biochem. Biophys. Res. Commun. 263 (1999) 672–677. DNA in human malignancy, Mutat. Res. 488 (2001) 119–133.

Vous aimerez peut-être aussi