Accepted Manuscript Copper(II) complexes of N3O tripodal ligands appended with pyrene and polyamine groups: Anti-proliferative and nuclease activities
Doti Serre, Sule Erbek, Nathalie Berthet, Xavier Ronot, Véronique Martel-Frachet, Fabrice Thomas PII: DOI: Reference:
S0162-0134(17)30595-0 doi:10.1016/j.jinorgbio.2017.11.006 JIB 10366
To appear in:
Journal of Inorganic Biochemistry
Received date: Revised date: Accepted date:
17 August 2017 28 October 2017 4 November 2017
Please cite this article as: Doti Serre, Sule Erbek, Nathalie Berthet, Xavier Ronot, Véronique Martel-Frachet, Fabrice Thomas , Copper(II) complexes of N3O tripodal ligands appended with pyrene and polyamine groups: Anti-proliferative and nuclease activities. The address for the corresponding author was captured as affiliation for all authors. Please check if appropriate. Jib(2017), doi:10.1016/j.jinorgbio.2017.11.006
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT Copper(II) complexes of N3O tripodal ligands appended with pyrene and polyamine groups: Anti-proliferative and nuclease activities
Doti Serre,a Sule Erbek,b Nathalie Berthet,a Xavier Ronot,b Véronique Martel-Frachet,b and Fabrice
PT
Thomasa*
RI
a. Département de Chimie Moléculaire, Université Grenoble Alpes, UMR-5250 CNRS UGA, CS 40700, 38058 Grenoble Cedex 9, France. Email :
[email protected].
AC
CE
PT E
D
MA
NU
SC
b. EPHE, PSL Research University, IAB, INSERM UGA U1209 - CNRS UMR 5309, 38700 La Tronche, France
ACCEPTED MANUSCRIPT Research highlights: - N3O tripodal ligands were appended by DNA targeting groups. - The copper complexes interact tightly with DNA. - The complexes exhibit micromolar nuclease activity.
AC
CE
PT E
D
MA
NU
SC
RI
PT
- The best nucleases show IC50 lower than cis-platin.
ACCEPTED MANUSCRIPT Abstract: A series of tripodal ligands based on the 2-tert-butyl-4-R-6-phenol was synthesized, where R = aldehyde (HL1), R = putrescine-pyrene (HL2) and R = putrescine (HL3). A dinucleating ligand wherein a putrescine group connects two tripodal moieties was also prepared (H2L4). The corresponding copper complexes (1, 2, 3, and 4, respectively) were prepared and characterized. We determined the phenol’s pKas in the range 2.47-3.93. The DNA binding constants were determined at
PT
6×106, 5.5×105 and 2.7×106 for 2, 3 and 4, respectively. The complexes display a metal-centered reduction wave at Epc,red = -0.45 to -0.5 V vs. saturated calomel electrode, as well as a ligand-centered
RI
oxidation wave above 0.57 V at pH 7. In the presence of ascorbate they promote an efficient cleavage
SC
of DNA, with for example a concentration required to cleave 50% of supercoiled DNA of 1.7 M for 2. The nuclease activity is affected by the nature of the R group: putrescine-pyrene ≈ bis-ligating >
NU
putrescine > aldehyde. The species responsible for strand scission is the hydroxyl radical. The cytotoxicity of the complexes was evaluated on bladder cancer cell lines sensitive or resistant to cis-
MA
platin. The IC50 of complexes 2 and 4 span over a short range (1.3 – 2 M) for the two cell lines. They are lower than those of the other complexes (3.1-9.7 M) and cis-platin. The most active compounds
CE
PT E
D
block the cell cycle at the G1 phase and promote apoptosis.
AC
Keywords: Copper; nuclease; DNA; tripodal ligand; phenol; anti-tumour.
Abbreviations: Boc: tert-ButylOxyCarbonyl; CT: Charge Transfer; CT DNA: Calf Thymus DeoxyriboNucleic Acid; CV: Cyclic Voltammetry; DFT: Density Functional theory; DMEM: Dulbecco's Modified Eagle's Medium; DMF: N,N-dimethylformamide; DMSO: dimethylsulfoxide; DNA: DeoxyriboNucleic Acid; EDTA: Ethylene Diamine Tetraacetic Acid; EI/DCI: Electron Ionization / Desorption Chemical Ionization; EPR: Electron Paramagnetic Resonance; ESI-MS:
ACCEPTED MANUSCRIPT ElectroSpray Ionization Mass Spectrometry; EtBr: Ethidium Bromide; FCS: Fetal Calf Serum; FITC: Fluorescein IsoThioCyanate; HR-MS: High Resolution Mass Spectrometry; L DNA: Linear DNA; MTT: 3-(4,5-DiMethylThiazol-2-yl)-2,5-diphenylTetrazolium bromide; NC DNA: Nicked Circular DNA; NMR: Nuclear Magnetic Resonance; NTO: Natural Transition Orbital; PBS: Phosphate Buffered Saline; X174 RF1: X174 Replicative Form 1; PI: Propidium Iodide; Q-TOF : Quadrupole
PT
Time-Of-Flight; RF: Resistance Factor; RPMI: Roswell Park Memorial Institute; SC DNA: Supercoiled DNA; SCE: Saturated Calomel Electrode; TBE: Tris - Boric acid - Ethylene Diamine
AC
CE
PT E
D
MA
NU
SC
RI
Tetraacetic Acid; TD-DFT: Time-Dependent Density Functional theory; UV-Vis: UV-Visible.
ACCEPTED MANUSCRIPT 1. Introduction Inorganic complexes provide a versatile platform for the design of anti-cancer agents.[1-17] Perhaps the most representative example is cis-platin (and its most recent derivatives oxaliplatin and carboplatin), which is one of the most potent anti-tumour drugs.[2, 6-9] Its biological activity results from the formation of platinum-DNA adducts inside the nucleus, which hamper cellular processes and
PT
ultimately lead to cellular apoptosis.[2, 6-9] Despite cis-platin being one of the most efficient anti-cancer agents, severe side-effects (nephrotoxicity, emetogenesis and neurotoxicity)[18, 19] and drug resistance limit its clinical use. In contrast to platinum, copper is an essential bio-element.[23] It is involved
RI
[20-22]
SC
in several natural processes such as dioxygen transport, oxidation catalysis and signalling. By following the logic that biologically relevant metal ions may be less toxic than non-biological ones,
NU
several groups have designed new families of anti-cancer agents based on copper complexes.[10-12, 15-17] These investigations are also bolstered by the fact that correlations have been recently established
MA
between the serum copper levels and the tumor incidence, malignant progression [24-27] as well as the response to treatment in some human cancers. [28] In addition, copper was found to be an essential
PT E
D
cofactor for the tumour angiogenesis process. [29]
Generally, the copper complexes that exhibit an interesting anti-proliferation activity also show a nuclease activity. [10-12, 14, 16, 30] The in vitro DNA cleavage reaction proceeds according to two main
CE
mechanisms, which usually depend on the presence or absence of exogenous agents. In the absence of exogenous agents, it is a hydrolytic pathway which is favoured, [31-42] although some examples of
AC
oxidative cleavage have been reported.[43-54] In the hydrolytic pathway the nucleophilic attack at the DNA phosphates is facilitated by metal coordination. This is instead an oxidative pathway that is observed in the presence of a reductant or oxidant.[55-77] Typically, the copper(II) is reduced to copper(I), which reacts with dioxygen to form toxic reactive oxygen species (O22-, O2•- and/or OH•) responsible for DNA strand scission. We recently reported a first generation of copper nucleases based on N3O tripodal ligands bearing a sterically hindered phenolate moiety, whose archetype is HL0 (scheme 1). We showed that the
ACCEPTED MANUSCRIPT complexes exhibit promising anti-proliferative activity (more active than cis-platin) against the bladder tumor cell lines RT112 and RT112CP (cis-platin resistant).[78]
Scheme 1
PT
We herein report a second generation of compounds, which incorporate either a putrescine or a putrescine-pyrene arm (Scheme 1). This ligand design allowed us to improve both the anti-
RI
proliferative activity against cancer cell lines and the nuclease activity of the complexes. The choice of
SC
these groups was motivated by the fact that i) putrescine is a natural polyamine whose targets at physiological pH (when the amines are protonated) are nucleic acids, [79] ii) polyamine analogues
NU
exhibit an antitumor activity in several cell lines (including breast and lung)[80-82] and iii) the pyrene is a planar and hydrophobic polyaromatic molecule, which can interact with DNA through -stacking
MA
(intercalation between the base pairs).[83] We report in this article the preparation and characterization
2. Materials and methods
PT E
D
of the copper complexes, as well as the evaluation of their anti-proliferative and nuclease activities.
CE
2.1 General methodologies
All chemicals were of reagent grade and were used without purification. NMR spectra were recorded
AC
on a Bruker AM 300 (1H at 300 MHz) spectrometer. Chemical shifts are quoted relative to tetramethylsilane (TMS). Mass spectra were recorded on a Thermofinningen (EI/DCI) or a Bruker Esquire ESI-MS apparatus. For pKa determinations UV/Vis spectra were recorded on a Cary Varian 50 spectrophotometer equipped with a Hellma immersion probe (1.000 cm path length). The temperature in the cell was controlled using a Lauda M3 circulating bath and the pH was monitored using a Methrom 716 DMS Titrino apparatus. A least square fit of the titration data was realized with the SPECFIT software. The fluorescence spectra were recorded on a Cary Eclipse spectrometer. Xband EPR spectra were recorded on a Bruker EMX Plus spectrometer equipped with a Bruker nitrogen
ACCEPTED MANUSCRIPT flow cryostat and a high sensitivity cavity. Spectra were simulated using the Bruker SIMFONIA software. Electrochemical measurements were carried out using a CHI 620 potentiostat. Experiments were performed in a standard three-electrode cell under argon atmosphere. A glassy carbon disc electrode (3 mm diameter), which was polished with 1 mm diamond paste, was used as the working electrode. The auxiliary electrode is a platinum wire, while a SCE was used as reference.
PT
2.2 Synthesis
RI
HL1, HL2, 3-tert-butyl-4-hydroxy-benzaldehyde and N1-(pyren-1-ylmethyl)butane-1,4-diamine were
SC
prepared according to literature procedures.[84-86] For HL3boc, HL3 and H2L4 (see below) the numbering used for the assignment of the NMR resonances is depicted in ESI.
NU
HL3boc. The ligand (3-tert-Butyl-4-hydroxy-5-benzaldehyde) (HL1, 300 mg, 0.74 mmol) and tertbutyl-N-(4-aminobutyl)carbamate (140 mg, 0,74 mmol) were dissolved in MeOH (30 mL) and the
MA
reaction mixture was stirred for overnight at room temperature. Sodium borohydride (84 mg, 2.22 mmol) was then added by small fractions during 3h and the reaction was further stirred for 2h. The
D
reaction mixture was extracted with AcOEt, washed with water and brine and dried over anhydrous
PT E
Na2SO4. The solution was evaporated under vacuum to afford yellow-brown oil. Column chromatography on silica gel with CH2Cl2/MeOH (95/5) with a gradient of triethylamine (0 - 5%) as
CE
the eluent afforded HL3boc as a colorless oil in a 57 % yield. NMR 1H (400 MHz, CDCl3) : δ (ppm) = 1.42 (s, 9H, t-Bu); 1.45 (s, 9H, t-Bu); 1.52 (m, 4H, f and g); 2.58 (s, 3H, Me-pyr); 2.62 (t, 2H, e or h);
AC
3.10 (m, 2H, e or h); 3.65 (s, 2H, a); 3.77 (s, 2H, b); 3.82 (s, 2H, d); 3.83 (s, 2H, c); 6.88 (d, 1H, 4JH-H = 1,9 Hz, H1 or H2); 6,99 (d, 1H, 3JH-H = 7,6 Hz, H9); 7.09 (m, 2H, H7 and H1 or H2); 7.12 (ddd, 1H, 3
JH-H = 7.5 Hz, 3JH-H = 5.0 Hz, 4JH-H = 1.3 Hz, H4); 7.38 (d, 1H, d, 3JH-H = 7.5 Hz, H6); 7.49 (t, 3JH-H =
7.6 Hz, H8); 7.60 (td, 1H, 3JH-H = 7.5 Hz, 4JH-H = 1.8 Hz, H5); 8.51 (dd, 3JH-H = 5.0 Hz, 4JH-H = 1.3 Hz, H3); 10.81 (s, 1H, OH phenol). NMR 13C (Q.DEPT, 400 MHz, CDCl3) : δ (ppm) = 24.27 (CH3, Mepyr); 27.35; 28.05 (CH2); 28.58; 29,70 (CH3, t-Bu phenol and Boc); 34.92 (C, t-Bu phenol); 40.60 (C, t-Bu Boc); 46.28; 49.12; 54.05; 57.74; 59.45; 59.59 (CH2); 120.20; 121.81; 122.23 (CHaro); 123.03 (Caro); 123.61; 126.44; 128.05 (CHaro); 129.29; 136.61 (Caro); 136.69; 136.94; 149.00 (CHaro); 155.63;
ACCEPTED MANUSCRIPT 156.18; 157.50; 158.14; 158.67 (Caro). C34H49N5O3 • CH2Cl2 • 2CH3OH: calcd. C, 60.30; H, 8.39; N, 9.25; found: C 59.98; H 8.39; N 9.52%. IR (ATR): ν (cm-1) = 3281 (broad), 2951, 2860, 2815, 1698, 1594, 1524, 1473, 1454, 1444, 1384, 1369, 1242, 1201, 1173, 1087, 1043, 989, 872, 755. HR-MS (QTOF): m/z, 576.3900; Calcd: 576.3914 for [M+H]+. HL3. The ligand HL3boc ( 202 mg, mg, 0.35 mmol) is dissolved in ethanol (20 mL) saturated with
PT
gaseous HCl beforehand. After one hour stirring at room temperature the solvent is evaporated under vacuum, affording HL3 as a white hydrochloride salt in a quantitative yield. NMR 1H (400 MHz,
RI
D2O) : δ (ppm) = 1.29 (s, 9H, t-Bu); 1.78 (m, 4H, f and g); 2.81 (s, 3H, Me-pyr); 3.07 (m, 4H, e and
SC
h); 4.01 (s, 2H, b); 4.12 (s, 2H, a); 4.39 (s, 2H, d); 4.41 (s, 2H, c); 7.16 (s, 1H, H2); 7.21 (s, 1H, H1); 7.77 (d, 1H, 3JH-H = 7.9 Hz, H7); 7.89 (m, 2H, H4 and H9); 7.97 (d, 1H, 3JH-H = 7.9 Hz, H6); 8.37 (t, 1H, JH-H = 7,9 Hz, H8); 8.46 (t, 1H, 3JH-H = 7.9 Hz, H5); 8.68 (d, 1H, 3JH-H = 5.6 Hz, H3). RMN 13C
NU
3
MA
(Q.DEPT, 400 MHz, D2O) : δ (ppm) = 19.00 (CH3, Me-pyr); 22.71; 24.05 (CH2); 28.93 (CH3, t-Bu); 34.11 (C, t-Bu); 38.82; 46.16; 50.47; 57.36; 57.43; 57.83 (CH2); 122.57; 124.59 (Caro); 124.95 ; 126.20 ; 127.14 ; 127.54 ; 129.38 ; 130.29 (CHaro); 138.99 (Caro); 141.33 ; 146.49 ; 146.70 (CHaro);
D
150.79 ; 151.90 ; 154.35 ; 154.66 (Caro). C29H49Cl4N5O3 • 4HCl • 2H2O: calcd. C 52.97; H 7.51; N
PT E
10.65%; found: C 52.94; H 7.85; N 10.55%. IR (ATR): ν (cm-1) = 3392 (broad), 2993, 2955, 2819,1650, 1597, 1477, 1458, 1429, 1366, 1309, 1242, 1207. MS (ESI): m/z (positive mode) 476.4
CE
[M+H]+; m/z (negative mode) 474.3 [M-H]-. H2L4. 3-tert-Butyl-4-hydroxy-5-benzaldehyde (HL1, 500 mg, 1.24 mmol) and 1,4-butane diamine (55
AC
mg, 0.62 mmol) were dissolved in MeOH (30 mL). Then, 6 molar equivalents of NaBH4 (141 mg, 3.72 mmol) were added in small fractions over a period of 3h under stirring at room temperature. The reaction was further stirred for 2h. The solvent was evaporated under vacuum and the reaction mixture was extracted with AcOEt, washed with water and brine and dried over anhydrous Na2SO4. The solution was evaporated under vacuum to afford yellow-brown oil that is sensitive to carbonatation. Column chromatography on silica gel with CH2Cl2/MeOH (95/5) with a gradient of triethylamine (0 5%) as the eluent afforded HL4 as a white powder in a 52 % yield. NMR 1H (400 MHz, CDCl3) : δ (ppm) = 1.44 (s, 18H, t-Bu); 1.57 (m, 4H, f); 2.57 (s, 6H, Me-pyr); 2.64 (m, 4H, e); 3.64 (s, 4H, a);
ACCEPTED MANUSCRIPT 3.76 (s, 4H, b); 3.81 (s, 4H, c); 3.82 (s, 4H, d); 6.86 (d, 2H, 4JH-H = 1.8 Hz, H2); 6.99 (d, 2H, 3JH-H = 7.6 Hz, H7); 7.10 (m, 6H, H1-H4-H9); 7.37 (d, 2H, 3JH-H = 7.7 Hz, H6); 7.49 (t, 2H, 3JH-H = 7.6 Hz, H8); 7.59 (td, 2H, 3JH-H = 7.7 Hz, 4JH-H = 1.8 Hz, H5); 8.50 (dd, 3JH-H = 4.6 Hz, 4JH-H = 0.7 Hz, H3); 10.84 (s, 1H, OHphenol). RMN 13C (Q.DEPT, 400 MHz, CDCl3) : δ (ppm) = 23.22 (CH3, Me-pyr); 27.97 (CH2 f); 29.71 (CH3, t-Bu); 34,93 (C, t-Bu); 49.22; 53.78 ; 57.72; 59.45 ; 59.59 (CH2); 120.21 ; 121.82 ; 122.23
PT
(CHaro); 123.11 (Caro); 123.63; 126.52; 128.16; 136.68; 136.94; 149.00 (CHaro); 155.73; 157.49; 158.13; 158.64 (Caro). C54H70N8O2 • H2CO3 • 0.5H2O: calcd. C 70.71; H 7.88; N 11.99%; found: C
RI
70.00; H 7.75; N 12.14%. IR (ATR): ν (cm-1) = 3400 (broad), 2949, 1594, 1578, 1457, 1432, 1366,
SC
1296, 1242, 1207, 1191, 1147, 1090, 1001. HR-MS (Q-TOF): m/z, 863.5674; Calcd: 863.5700 for [M+H]+.
NU
Copper complexes. The copper complex 1 was isolated as single crystals, as previously described,[84] and used under this form for the spectroscopic and biological studies. Complex 2 were prepared in situ
MA
by mixing equimolar amounts of CuCl2 • 2 H2O, triethylamine and the ligand in DMF, as reported in literature.[84] Complexes 3 and 4 were prepared in a similar way, except that two molar equivalent of
D
copper were used in the case of 4. The concentrated DMF solutions of the complexes were diluted in
CE
2.3 DFT calculations
PT E
water in order to get a final medium of composition (DMF:H2O) (10:90).
All density functional theory (DFT) calculations were performed with the Gaussian suite (release
AC
9.0)[87] by using the B3LYP functional[88-90] in combination with the 6-31g* basis set for all the atoms. After geometry optimization a frequency calculation was systematically run in order to ensure that the optimized structures correspond to real energy minima and not saddle points. The relative energies are the sum of electronic and thermal free energies at 298 K. The spectroscopic properties were calculated by time dependent DFT[91] at the same level of theory than geometry optimization. The solvent was taken in account through a polarized continuum model (PCM)[92] and the 30 lowest energy excitations were calculated. 2.4 Determination of the DNA binding constants
ACCEPTED MANUSCRIPT The DNA binding constants were obtained from titration of solutions of the complexes with CT DNA. DNA (type I, fibers, from Sigma Aldrich) was first purified by four successive extractions with the phenol-CHCl3-isoamyl alcohol 25-24-1 mixture. Then 0.1 volume (with respect to the aqueous phase) of a 3 M sodium acetate solution and 2 volumes of ethanol were added and the solution was stored overnight at 253 K. The precipitated DNA was solubilised and analyzed by UV-Vis to ensure that the
PT
A260/A280 ratio was ≥ 1.8. For the direct determination of the binding constants by fluorescence the complex (1 M) was mixed with various concentrations of CT DNA (5 - 200 M in base pair) in a
RI
buffered H2O:DMF 90:10 solution containing [Tris-HCl] = 0.05 M, [NaCl] = 0.02 M at pH 7. After
SC
the equilibrium was reached (10 min) the fluorescence spectrum was recorded at 298 K. The DNA binding constants K were deduced from a Scatchard–Von Hippel plot. Typically, the ratio r/c was
NU
plotted against r to give a straight line whose slope is K. r corresponds to the number of bound molecules per site, while c represents the concentration of free drug. The concentration of bound
MA
molecules was obtained from the expression C0 × (f0-f)/(f0-fb), where f0 is the fluorescence of the free drug, f the fluorescence at any DNA concentration, fb the fluorescence of the drug bound to DNA and
D
C0 the total concentration of drug. c was deduced from the mass balance. In order to avoid
PT E
inaccuracies inherent to the determination of the concentration of free and bound molecules at the end and beginning, respectively, of the titration we removed the two firsts and two lasts points. Each binding constant was obtained from a duplicate experiment. For the competition experiment both the
CE
DNA and ethidium bromide (EtBr) concentrations were kept constant at 20 M and increasing
AC
concentrations of complex were added (2 - 50 M). The fluorescence was measured with an excitation wavelength of 517 nm, which corresponds to the excitation maximum for EtBr. The data were treated as detailed in the text, by using a binding constant for EtBr of 4.9 105 M-1.[93] 2.5 Procedure for DNA cleavage experiments X174 RF1 DNA was purchased from Fermentas and was stored at -20°C. The typical reaction mixture, containing double-stranded DNA and the copper complexes in a 10 mM phosphate solution (pH 7.2) with 10% (vol) DMF, was incubated at 37°C for the required time, with or without additives. After the incubation period, the reaction was quenched at -20°C, followed by the addition of loading
ACCEPTED MANUSCRIPT buffer (6x loading dye solution, Fermentas). The reaction mixture was loaded on a 0.8 % agarose gel in Tris-Boric acid-EDTA (TBE) buffer (pH 8), (0.5 x TBE) and electrophoresis was performed at 70 V for 2h30 - 3h. After DNA migration, the gels were stained by incubation for 10 min with a 1g/ml EtBr solution then washed with distilled water. The gels were visualised and the fluorescence quantified using Imager Typhoon 9400 and Image Quant Software. The cleavage efficiency was
PT
measured by determining the ability of the complex to convert the supercoiled DNA (SC) to nicked
RI
circular form (NC) and linear form (L).
SC
2.6 Cell culture
Human bladder cancer cell line RT112 and VERO cells (derived from the kidney of a healthy green
NU
monkey) were purchased from Cell Lines Service (Eppelheim, Germany). Cisplatin resistant RT112 cells (RT112-CP) were kindly provided by B. Köberle (Institute of Toxicology, Clinical Centre of
MA
University of Mainz, Mainz, Germany). RT112 and RT112-CP cells were cultured in RPMI 1640 medium and VERO cells in DMEM. These media are supplemented with 10% (v/v) fetal calf serum
D
(FCS) and 2 mM glutamine (Invitrogen Life Technologies, Paisley, UK). Cells were maintained at
PT E
37°C in a 5% CO2-humidified atmosphere and tested to ensure freedom from mycoplasma contamination.
CE
2.7 Cell proliferation assay
Inhibition of cell proliferation by copper complexes was measured by a MTT (3-(4,5-Dimethylthiazol-
AC
2-yl)-2,5-diphenyltetrazolium bromide) assay. RT112 and RT112-CP cells were seeded into 96-well plates (5 x 103 cells / well) in 100 µl of culture medium. After 24 h, cells were treated with cisplatin (Sigma-Aldrich, Lyon France) or complexes at various concentrations. In parallel, a control with DMF (vehicle alone) at the same dilutions was done. Following incubation for 48 h, 10 µl of a MTT (Euromedex, Mundolsheim, France) stock solution in PBS at 5 mg/ml was added in each well and the plates were incubated at 37 °C for 3 h. Plates were then centrifugated 5 min at 1500 rpm before the medium was discarded and replaced with DMF (100 µl/well) to solubilize water-insoluble purple formazan crystals. After 15 min under shaking, absorbance was measured on an ELISA reader (Tecan,
ACCEPTED MANUSCRIPT Männedorf, Switzerland) at a test wavelength of 570 nm and a reference wavelength of 650 nm. Absorbance obtained by cells treated with the same dilution of the vehicle alone (DMF) was rated as 100% of cell survival. Each data point is the average triplicates of three independent experiments. 2.8 Cell cycle analysis
PT
RT112 cells were cultured in 6-wells plates (2 x 105 cells/well) for 24 h before treatment with DMF or complexes for 24 h or 48 h. Trypsinized and floating cells were then pooled, fixed with 70% ethanol,
RI
washed with PBS and stained with 20 µg/ml propidium iodide (PI) in the presence of 0.5 mg/ml of
SC
RNAse (Sigma-Aldrich, Lyon France). Data acquisitions were performed with the Accuri C6 cytometer (BD Biosciences, Le Pont de Claix, France) equipped with a 488 nm laser Argon and PI
NU
fluorescence was collected with a 585±44 nm filter. Parameters from 2x104 cells were acquired using the Cell Quest Pro software (Becton Dickinson). The percentage of cell cycle distribution in the G1, S
MA
and G2+M phases was determined using FCS Express 5 software (De Novo Sofware, Los Angeles, CA).
D
2.9 Apoptosis detection
PT E
RT112 and RT112-CP cells were treated with DMF or complexes for 24 h or 48 h. Trypsinized and floating cells were then pooled and labeled with FITC-coupled Annexin V (Miltenyi Biotec, Paris,
CE
France) and PI following the manufacturer's instructions. Data acquisitions were performed with the Accuri C6 cytometer. Parameters from 2x104 cells were acquired using the C-Flow software. For each
AC
condition, fractions of live cells (Annexin V negative / PI negative), early apoptotic cells (Annexin V positive / PI negative) and late apoptotic cells (Annexin V positive / PI positive cells) were quantified with FCS Express 5 software (De Novo Software). The experiments were repeated three times independently.
3. Results and discussion 3.1 Synthesis of the ligands
ACCEPTED MANUSCRIPT The ligand synthesis is based on the preparation of an aldehyde precursor and a reductive amination. We recently reported the synthesis of the ligands HL1 and HL2.[84] Reductive amination of HL1 by the pyrene-putrescine conjugate PyrPt[86] affords the ligand HL2.[84] Reductive amination of HL1 by the tert-butyl-N-(4-aminobutyl)carbamate affords the ligand HL3boc, which was deprotected under acidic conditions to give HL3. The reaction of 1,4-butane-diamine with two equivalents of HL1 in the
RI
PT
presence of NaBH4 affords the binucleating ligand H2L4.
SC
3.2 Formation of the complexes followed by mass spectrometry and UV-Vis spectroscopy Equimolar mixtures of the ligands HL1-3 and copper chloride were prepared in DMF and diluted in
NU
water for characterization by mass spectrometry (ESI-MS). For all the mixtures we obtained clean spectra with a single dominant peak corresponding to the expected copper complex. Indeed, a
MA
prominent peak is observed at m/z = 465.16 (monocation) in the case of HL1, which corresponds to [(L1)- + Cu2+]. Regarding HL2 the main peak is observed at m/z = 787.08 and assigned to [(HL2) + Cl-
D
+ Cu2+]. For HL3 the dominating peak is observed at m/z = 269.13 and corresponds to a dication. It is
PT E
formulated as [(HL3) + Cu2+]. Finally, the dinucleating ligand H2L4 gives, in the presence of two molar equivalents of copper, a prominent peak at m/z = 1095.31 (monocation), together with a peak of weaker intensity at m/z = 1059.32, which are assigned to [(H2L4) + 3 Cl- + 2 Cu2+] and [(HL4)- + 2 Cl-
CE
+ 2 Cu2+], respectively. Thus, ESI-MS confirms the formation of complexes with the expected (1:1)
Figure 1
AC
and (2:1) stoichiometries (M:L) for HL1-3 and H2L4, respectively.
The addition of copper chloride to a DMF solution of the ligands HL1-3 and H2L4 in the presence of NEt3 (to ensure phenol deprotonation) induces a colour change of the solution from pale yellow to violet, which is characterized by the development of a new band at around 470 nm (see below). Consistent with the ESI-MS data its intensity is maximal at one molar equivalent added for HL1-3,
ACCEPTED MANUSCRIPT disclosing the formation of complexes with a 1:1 stoichiometry (complexes denoted 1, 2 and 3, respectively, scheme 2).[94] In the case of H2L4 two molar equivalents or copper must be added in order to reach the plateau, confirming the dinucleating properties of this ligand, which forms the bimetallic
PT
complex denoted 4 (scheme 2).
Scheme 2
RI
In order to gain insight into the structure and the behaviour of the complexes in aqueous medium we
SC
investigated their spectroscopic properties as a function of the pH (Figure 2).[24] For solubility reasons (the complexes are not soluble in neat water) we formed the 1:1 (1-3) or 2:1 (4) M:L complexes in situ
NU
in a concentrated DMF solution, and next diluted them in water until reaching a water:DMF ratio of 90:10. At pH = 7 the UV-Vis spectra of all the complexes are dominated by a band at ca. 470 nm
MA
(Table 1), which corresponds to the phenolate-to-copper charge transfer (CT) transition.[78, 84, 85, 95-98][99] This assignment is corroborated by the blue shift of the max on going from 1 (electron withdrawing
D
CHO para substituent) to 2-4 (electron donating alkyl para substituent). A low intensity shoulder is in
PT E
general observed at lower energy, which is assigned to the copper(II) d-d transitions (Table 1). The intensity of the CT band is twice as large in complex 4 in comparison to complex 1, in agreement with
Figure 2
AC
CE
the chelation of two metal centers.
Upon decreasing the pH from 7 to 2 the charge transfer transition decreases in intensity, consistent with the protonation of the phenolate moiety. The evolution of the spectra as a function of the pH was fitted to give the pKa values: 2.47±0.01, 3.93±0.02, 3.77±0.01, for 1, 2 and 3, respectively (scheme 3).
Scheme 3
ACCEPTED MANUSCRIPT For complex 4 the titration data were refined by using a model involving two protons, giving a -log2 value of 7.97±0.03. For every complex the pKa values are much lower than 7, indicating that the complexes exist mainly under their deprotonated form under physiological conditions. These values deserve few comments. Firstly, complex 1 shows the smallest pKa value, in line with the electron withdrawing effect of the aldehyde para substituent. Secondly, the pKa values are much lower than
PT
those reported for complexes featuring axial phenolates.[96-98] The easier deprotonation in our case reflects a stronger Cu-O bond, suggesting an equatorial positioning of the phenolate.[84, 96-98] The
RI
formation of a 5-membered chelate ring with the -methylpyridine in equatorial position is indeed
SC
strongly unfavored for steric reasons (steric clash due to the methyl substituent). This enforces an axial positioning of the -methylpyridine, and subsequently an equatorial coordination of the phenolate.
MA
NU
This hypothesis is fully supported by the crystal structure of 1.[84]
Table 1
PT E
D
3.3 EPR spectroscopy
The EPR spectra of complexes 1, 2, 3 and 4 at 100 K display the typical pattern of mononuclear copper(II) compounds. The (S = ½) signal is axial and hyperfine splitted, as a result of the interaction
CE
of the spin with the copper nuclear spin (ICu = 3/2). From simulation of the spectra we obtained the spin Hamiltonian parameters listed in Table 2. The ordering of g is indicative with a dx2-dy2 ground spin
AC
state for the copper ion, consistent with a square planar or square pyramidal geometry.
Figure 3 At pH = 7 the spectra of 2, 3 and the boc protected 3(Boc) are superimposable, supporting a similar metal ion geometry (Figure 3). This subsequently rules out the coordination by the terminal amine in the case of 3. The EPR spectrum of 4 exhibits the same shape than that of 3, but its intensity is twice as large. It can therefore be concluded that the two copper centers are magnetically uncoupled. This
ACCEPTED MANUSCRIPT idea is reinforced by the absence of MS = 2 resonance at around g = 4, as well as similarities in the linewidth for both 3 and 4. We interpret this result by a high flexibility of the putrescine arm and the functionalization in para position, which both do not favour a spatial proximity of the metal ions. The EPR spectra recorded at pH = pKa are broader than those taken at pH = 7, pKa+1 and pKa-1. The broadening is substantial in the parallel region, consistent with the presence of an equimolar mixture
PT
of the copper-phenol and copper-phenolate complexes in solution. Examination of the spectra in the
RI
low field region at pH = pKa+1 than at pKa-1 reveals substantial changes in the hyperfine coupling constant A//, congruent with a change in the copper ion geometry. An empirical relationship based on
SC
the g///A// ratio has been often used to estimate the tetragonal distortions of the copper ion.[100] Whatever is the pH the g///A// ratio is 126-133 cm within the present series (Table 2), indicative of
NU
weak tetragonal distortions. It is worth noting that the g///A// ratio is smaller at pH = pKa-1 than at
MA
pKa+1 (Table 2). Thus the distortions are less marked at low pH, presumably due to the weak coordinating ability of the phenol, which does not impose significant distortion in the coordination
D
polyhedron.
PT E
Interestingly, the spin Hamiltonian parameters for the phenolate complexes slightly differ between the saline (0.1 M NaCl) and Tris-HCl buffers (0.05 M containing 0.02 M of NaCl). We take this as evidence that an exchangeable ligand completes the coordination sphere of the copper center, in
CE
agreement with the structures depicted in Scheme 3. In order to gain insight on the nature of this exchangeable group we prepared the four complexes by mixing the ligand and the appropriate amount
AC
of CuCl2 2 H2O in DMF:H2O (10:90) without supporting electrolyte and added to the resulting solutions either NaCl or silver trifluoromethane sulfonate in excess. In this latter case precipitation of AgCl is observed, indicating removal of chloride from the solution. The EPR spectra without additive are very similar to those taken with silver trifluoromethane sulfonate, but they differ from those recorded in the NaCl 0.1 M medium. The main difference concerns the g// value, which is systematically lower in the NaCl 0.1 M medium (g// = 0.005). We take this behaviour as evidence of a change in the coordination polyhedron, which is assigned to the binding of Cl-. The increase in g// can indeed be interpreted in terms of an increased delocalization of the spin over the exogenous
ACCEPTED MANUSCRIPT chloride ligand, in agreement with DFT calculations (Mulliken spin population of 0.686 vs. 0.691 for the phenolate complexes with chloride and water equatorially bound, respectively). Thus, these data support binding of a chloride molecule at the equatorial position of the copper in an electrolytic
PT
medium containing a large amount of NaCl, such as that used for spectroscopic titrations.
Table 2
RI
3.4 DFT calculations
SC
We investigated the structure and spectroscopic properties of the copper complexes by DFT and TDDFT calculations (see ESI). In order to minimize the computational cost we designed an archetype
NU
denoted HLarch wherein the alkylammonium chain in para position of the phenol was replaced by a
MA
methyl substituent, which exerts similar electronic effects (Figure 4). The geometry of the copper complexes under their protonated “phenol” ([(HLarch)Cu(X)]y) and deprotonated “phenolate” [(Larch)Cu(X)]y-1) forms was optimized by taking in account the solvent (water) and by considering
D
both a water molecule (X = H2O, Figure 4) and chloride ion (X = Cl) as exogenous ligand. With the
PT E
aim of gaining insight on the positioning of the phenolic moiety we used as initial guess structures wherein this group is arbitrary located either in equatorial or axial environment. For
CE
[(HLarch)Cu(H2O)]2+ the isomer showing the phenol bound in axial position lies 7.2 kcal/mol below the one that features an equatorially bound phenol. For [(Larch)Cu(H2O)]+ (deprotonated) the
AC
difference in energy between the isomers is 6.2 kcal/mol, but this is now the structure showing an equatorially bound phenolate that is lower in energy. We next considered complexes featuring an exogenous chloride ion instead of the water molecule since such structures might form in the saline buffer medium. Both the axial positioning of the phenol and the equatorial positioning of the phenolate are again favoured, but the energy differences between the conformers are smaller (2.9 and 0.7 kcal/mol for the phenol and phenolate forms, respectively).
ACCEPTED MANUSCRIPT Figure 4 The absorption spectrum of the complexes was then computed by TD-DFT methods. We considered only the most favoured isomers in our calculations. For [(HLarch)Cu(H2O)]2+ an electronic excitation is predicted at 635 nm, which corresponds to the band detected at 680-690 nm in the visible spectrum of 2-4. The oscillator strength is rather small (fosc = 0.0017), in line with the small molar extinction
PT
coefficient observed experimentally. Consistent with its low intensity its assignment is dxz or dyz dx2-y2 transition. When [(HLarch)Cu(Cl)]+ is considered instead of [(HLarch)Cu(H2O)]2+ the dxz or dyz
RI
dx2-y2 transition is predicted at lower energy, 652 nm (fosc = 0.0012), in better agreement with the
SC
experimental spectrum. We next examined the deprotonated complex [(Larch)Cu(H2O)]+ (equatorial phenolate) with a coordinated water molecule. An intense electronic excitation is computed at 576 nm
NU
(fosc = 0.0454), which is assigned to the experimental band at 457-461 nm. The oscillator strength is
MA
much larger than in the protonated complex, in agreement with experiment. The NTO analysis discloses a different origin for this band, which is phenolate-to-copper charge transfer transition. Interestingly, calculations predict a shift of this band towards 547 nm when the chloride ion is
D
considered as exogenous ligand instead of the water molecule, without significant change of oscillator
PT E
strength (fosc = 0.0433). This predicted max again better fits with the experimental values, supporting
CE
the binding of a chloride ion in electrolytic media.
AC
3.5 Electrochemical behaviour of the complexes The electrochemical behaviour of the complexes (Figure 5, Table 3) has been investigated by cyclic voltammetry (CV) in a water: DMF 90:10 medium. Before commenting on the CV of the putrescine appended complexes, we will first discuss the behaviour of the archetype 1 (Figure 4a). The CV curve of 1 displays a cathodic peak at Epc,red = -0.46 V vs. SCE, which corresponds to the reduction of the copper(II) to copper(I). On the reverse scan an oxidation peak is observed at Epa,red = -0.12 V. This peak is absent from the anodic scan, when the initial potential is set at -0.2 V. Thus, Epa,red at Epc,red are found as a pair of semi-reversible peaks, which are attributed to the copper(II)/copper(I) redox couple.
ACCEPTED MANUSCRIPT The large Ep value (Ep = 0.34 V) attests that a chemical reaction is coupled to the reduction, which we assign to rearrangement and/or protonation of the phenolate.[84] Upon scanning towards a more positive potential an irreversible oxidation wave is detected at Epa,ox = 0.96 V (Table 3). It is assigned to the oxidation of the phenolate into a phenoxyl radical.[48, 78, 84, 85, 95-97, 101] The irreversibility of the wave is indicative of a low stability of the radical species, which is not unexpected owing to the
RI
PT
electron withdrawing properties of the CHO substituent.
SC
Figure 5
We next examined the electrochemical behaviour of the conjugate PyrPt (Figure 5b). No clear
NU
reduction wave could be discerned when scanning down to -1.4 V, while an anodic signal could be observed at 0.86 V, followed by ill-defined broad oxidation wave above 1 V. The oxidation peaks
D
MA
arises from the irreversible oxidation of the amines[102] followed by the pyrene group.[103]
PT E
Table 3
The CV curves of 2 and 3 are depicted in Figure 5c-d, while that of 4 is shown in ESI. The
CE
copper(II)/copper(I) redox couple in 2, 3, and 4 is observed as a pair of well separated peaks Epa,red and Epc,red, similarly to 1. This again supports a rearrangement of the coordination sphere associated to the
AC
electron transfer. The substituent effect on the Cu(II)/Cu(I) redox couple, if any, is difficult to appreciate due to broadening of the cathodic peak (Epc,red). The characteristic features of the PyrPt substituent are observable above 0.8 V in the CV of 2, in addition to a shoulder in the range 0.5-0.7 V, which is assigned to the phenoxyl/phenolate redox couple. This latter wave is observed at a lower potential in 2, 3 and 4 when compared to 1, consistent with the electron donating properties of the alkyl para substituent.[101]
ACCEPTED MANUSCRIPT 3.6 Fluorescence study The pyrene moiety is known to form excimers in solution, which are responsible for its intense fluorescence at around 470 nm. Upon addition of DNA to 2 the fluorescence at 468 nm progressively decreases. This behaviour is consistent with the intercalation of the pyrene moiety in DNA. The spectral changes were fitted by using a Scatchard – Von Hippel model (Figure 6),[104] giving a K value
PT
of 2.6 × 106 M-1 (n = 2, Table 4). The K value is three orders of magnitude larger than that of the copper complex of HL0,[78] pointing out the beneficial influence of the pyrene-putrescine chain. It is
RI
slightly smaller than that reported for PyrPt (1.1 × 107 M-1).[86] However, this difference likely arises
SC
from the higher salinity of the medium in our case (0.07 M versus 0.0097 M in Collette’s work), which disfavours ionic interactions between the ammoniums and the sugar-phosphate backbone of
NU
DNA. Finally, the K values of 2 is much higher than that of the classical intercalator ethidium bromide
MA
(EtBr) and putrescine,[105] showing the synergetic effect of the polyamine chain and the tripodal unit.
D
Figure 6
PT E
Table 4
In order to shed some light on the role of the pyrene and the tripodal unit in the strength of the
CE
interaction with DNA we conducted competition experiments between EtBr and 2-4. [106] This
AC
methodology has been successfully used to measure the binding constant of putrescine to DNA,[107] despite the fact that this polyamine is not an intercalator. A plot of the fluorescence intensity as a function of the [complex]/[EtBr] ratio is shown in Figure 7 for 2 (those of PyrPt, 3 and 4 are shown in ESI). We calculated the concentration of the complex that induces a quenching of 50 % of fluorescence. The following equation:
gives access to the apparent binding constants Kapp. By using the reported values KEtBr = 6.9 × 105 M-1 [93]
and [EtBr] = 20 M we calculated the Kapp values reported in Table 4. The Kapp of 2, 3 and 4 are
ACCEPTED MANUSCRIPT 6 × 106, 5.5 × 105 M-1 and 2.7 × 106 M-1, respectively. We also determined the binding constant of PyrPt under our experimental conditions, which gave 1.7 × 106 M-1. It must be underlined that the agreement between the K determined by the direct method for 2 and the Kapp from competition experiment is quite reasonable. When the Kapp of 2 is compared to that of 3 it immediately comes that the pyrene moiety enhances the
PT
binding constant by ca. one order of magnitude. In addition, when the Kapp of 4 is compared to those of 2 and 3 one can conclude that an extra tripodal unit is a better anchor than the terminal amine, but it is
RI
not as much as efficient as a pyrene moiety. In line with these results, the apparent binding constant of
SC
PyrPt is four times smaller than that of 2, again showing that the copper-tripodal unit enhances the
NU
binding to DNA.
MA
Figure 6
D
3.7 DNA cleavage
PT E
The nuclease activity of the copper(II) complexes has been investigated by gel electrophoresis, by using a plasmidic X174 RF1 supercoiled DNA, in a water: DMF 90:10 phosphate buffer (10 mM) at pH 7.2. Upon cleavage, the covalently closed circular supercoiled (SC) plasmidic DNA is converted
CE
into the nicked circular (NC) form (single strand breakage) and then to the linear (L) form, which
AC
corresponds to a double strand breakage.
Figure 8 The nuclease activity was first investigated in the absence of reductant, after 1h of incubation at 37°C by using concentrations of complexes ranging between 3 and 50 M. The results are summarized in Figure 8. No significant cleavage activity was observed for complex 1 at 50 M. The gel electrophoresis shows the disappearance of the SC form between 10 and 30 M (full disappearance at
ACCEPTED MANUSCRIPT 50 M) for complex 2 (Figure 8b) and between 30 and 50 M for complex 4 (Figure 8d), without concomitant apparition of the typical pattern for single or double strand cleavage. The absence of nicked circular or linear form in the gel has prompted us to repeat the experiment over a sharper range of complex concentrations. Between 10 and 30 M (complex 2, Figure 8f) or between 20 and 50 M (complex 4, ESI), i. e. before disappearance of the band, the supercoiled plasmid DNA clearly
PT
experiences a delay in the migration. This strongly suggests that the binding of the complexes promotes condensation of the DNA. It is indeed known that the polyamine chains, protonated under
RI
physiological conditions, can interact strongly with the phosphate groups of DNA. In some cases it
SC
leads to the aggregation of DNA,[108-111] which manifests by a delay in the migration, as experimentally observed for 2 and 4. Regarding 3 (Figure 8c) and especially in the case of 3(Boc) (Figure 8e) the
MA
NU
effect is less pronounced, consistent with their weaker affinity with DNA.
Figure 9
D
We next examined the nuclease activity of the complexes in the presence of two reductants,
PT E
mercaptoethanol and ascorbic acid, after one hour of reaction at 37°C (Table 5).[112] As shown in Figure 9, all the complexes exhibit a higher nuclease activity in the presence of a reductant than
CE
without it. As the concentrations used to promote cleavage are lower in this experiment, the putative DNA condensation generally does not hamper the observation of the NC form. A common observation
AC
is that the cleavage activity is higher when the reductant is ascorbic acid than mercaptoethanol. We interpret this result by the fact that mercaptoethanol could potentially bind to the copper(I) (soft Sdonor), interfering with the reaction. We will therefore limit our discussion to the results obtained with ascorbic acid in the following section. Complex 2 is the most active nuclease, which cleaves 50 % DNA at the concentration of 1.7 M. Complex 1, which bears neither a pyrene, nor a putrescine chain is the least active compound (45 M for 50% of cleavage), while complex 3 is intermediate (14 M). Thus, the nuclease activity is correlated to the nature of the R substituent according to: pyreneputrescine > putrescine > CHO. This trend reveals that the affinity of the complex for DNA is the
ACCEPTED MANUSCRIPT main determinant for the nuclease activity of the compounds. Interestingly, at a concentration of 30 M, complex 3 also promotes very efficiently the double strand cleavage of DNA. At this concentration the cleavage produced an equimolar mixture of NC and L forms. Regarding the dicopper complex 4, gel electrophoresis displays the total disappearance of the SC form at concentrations in complex higher than 20 M, without appearance of bands characteristic of the NC and L forms. This
PT
peculiar behaviour is in line with that observed in the absence of reductant, hampering the determination of precise IC50 values. It is worth noting that both the NC and L forms could be
RI
observed with ratio of 67:33 in the case of 4 at a concentration of 15 M, while no more SC form was
SC
present. For comparison the SC form fully disappeared at a complex concentration in the range 5-10
NU
M in the case of 2. It is therefore clear that complex 4 is also a powerful nuclease.
MA
Table 5
In order to gain insight into both the binding mode and the reaction mechanism we conducted
D
experiments in the presence of various scavengers and exogenous agents. The results obtained for 1, 2
PT E
and 3 (at concentrations inducing significant cleavage) are depicted in Figure 10. By using a concentration of 2 of 8 M, it is observed that 75 % of the plasmid was converted into its NC form
CE
after 1 hour of reaction. The addition of superoxide dismutase (lane 4), catalase (lane 7) or NaN3 (lane 5) did not affect the yield of cleavage, showing that neither superoxide, nor hydrogen peroxide or
AC
singlet dioxygen is involved in the reaction. When CH3CH2OH (lane 2) or DMSO (lane 3, OH• scavengers) are added to the medium, the ratio of NC form drops to 63 and 42 %, respectively, indicative of the involvement of the hydroxyl radical in the cleavage reaction. Thus, it can be proposed that the copper(II) ion is reduced by ascorbic acid into copper(I), which reacts with molecular dioxygen to form the freely diffusible hydroxyl radical responsible for DNA strand scission.[10,15] The essential role of the copper in this mechanism could be confirmed by the inhibitory effect of EDTA (lane 6, 32 % NC form), which competes with the ligands for metal chelation.
ACCEPTED MANUSCRIPT Figure 10 The salt NaCl (350 M) has a marginal effect on the DNA cleavage (lane 9), while the addition of a minor groove binder (Hoechst 33258, lane 8) significantly alters the activity of 2 (48 % of NC form). These results indicate that 2 preferentially binds to the minor groove of DNA. They are consistent with the observation made by Ramirez et al. that free putrescine interacts primarily with the minor groove
PT
of duplex DNA[113] and suggest that the polyamine chain of 2 has the sufficient degree of freedom to
RI
interact with DNA in a similar way than the free putrescine. We can therefore propose a binding model wherein the putrescine chain interacts with the sugar-phosphate backbone in the minor groove,
SC
while the pyrene moiety is intercalated into base pairs. Consistent with this assumption, the Hoechst 33258 agent, which dramatically alters the nuclease activity of 2 has only a minor effect on the activity
NU
of 1 (46 % versus 40 % of NC form after 1 hour incubation, by using a complex concentration of 50
PT E
3.8 Anti-proliferative activity
D
MA
M).
These promising nuclease activities of the complexes in vitro encouraged us to investigate their activity in cellulo. The standard treatment for muscle-invasive and metastatic bladder cancer is a
CE
neoadjuvant cis-platin-based combination chemotherapy followed by radical cystectomy. However, chemoresistance against cis-platin and its toxicity limit its use to patients with advanced bladder
AC
cancer. So, we evaluated the cytotoxicity of copper complexes on bladder cancer cell lines sensitive (RT112) or resistant to cis-platin (RT112-CP). Cells were incubated with various concentrations of the complexes for 48 h and proliferation was then monitored by a MTT assay. The obtained IC50 values are summarized in Table 6.
Table 6
ACCEPTED MANUSCRIPT Except for 1, the anti-proliferative activity of the complexes seemed to be directly correlated with their nuclease activity and their affinity towards DNA. Complex 2, which presented the highest cleavage activity was 5 times more active than 3, which appeared as less efficient nuclease. Complex 4 was the most active compound, with an IC50 of about 1 µM. Interestingly, complexes 2 and 4 presented the same cytotoxic activity against the cis-platin-resistant cells RT112-CP and were about 20 times more
PT
active than cis-platin against this cell line. Remarkably, all compounds appended by at least a putrescine chain were more efficient than cis-platin against RT112 cells and more importantly, all the
RI
complexes (irrespective of the phenolate para substituent) overcome the resistance observed with cis-
SC
platin on RT112-CP cells. As shown in Table 6, the RT112-CP cells are 2.6 times more resistant to cis-platin than the RT112 parental cell line, as expressed by the resistance factor (RF). On the other
NU
hand, the RF for copper complexes ranged between 0.6 and 1.8.
MA
The cytotoxicity of the complexes was also evaluated on the VERO normal non-tumor cell line. All the complexes were 1.5-3.6 fold less active against VERO cells compared to tumoral cell lines. It is particularly noteworthy that complex 4, the most active compound, is about 2.5 times less cytotoxic
CE
3.9 Apoptosis induction
PT E
D
against VERO cells than on RT112 or RT112-CP cell lines.
In order to determine whether the anti-proliferative activity of the complexes results from apoptosis
AC
induction, we performed an annexin V/propidium iodide (PI) double staining, followed by flow cytometry analysis after incubating the cells with 1, 2 and 4, i.e. the compounds that show the lowest IC50. Based on the results of the MTT assay, RT112 cells were treated with the IC50 or 1.5 fold the IC50 concentration for 24 or 48 h. As shown on Figure 11, the fraction of apoptotic cells was dramatically increased in a dose and time dependent manner after incubation with all the investigated complexes. After a 48 h treatment with the IC50 concentration, the percentage of apoptotic cells were 42% for 1 and 2 and 43% for 4, compared to 12% for the control condition without treatment. These results
ACCEPTED MANUSCRIPT clearly indicate that the apoptotic pathway is implicated in the anti-proliferative activity of the complexes.
Figure 11
PT
3.10 Cell cycle arrest
RI
To further understand the mechanisms underlying proliferation inhibition and apoptosis induction, we evaluated the effect of the complexes on cell cycle progression. RT112 cells were treated with IC50
SC
concentrations of molecules 1, 2 and 4. After 24 h, the cell cycle phase distribution of treated cells was
NU
determined by flow cytometry. Our results showed that 2 and 4 caused a significant accumulation of RT112 cells at the G1 phase with a concomitant decrease of the number of cells in S and G2+M
MA
phases (Figure 12). The percentage of cells at the G1 phase after incubation with complexes 2 and 4 were 72% and 73%, respectively, compared to 51% in the non-treated cells. Thus, both 2 and 4 appeared as powerful blockers of the cell cycle, in line with their strong inhibitory effect on RT112
PT E
D
cells proliferation.
AC
4. Conclusions
CE
Figure 12
In summary, we herein report an original functionalization of two tripodal ligands with putrescine and putrescine-pyrene chains. The copper-phenolate complexes were prepared and characterized. The binding constants towards duplex DNA are greatly enhanced (up to 3 orders of magnitude) in comparison to previously described copper-tripodal complexes that do not harbour these chains. This new generation of compounds shows a high nuclease activity in the presence of ascorbic acid, with a remarkable IC50 of 1.7 M for 2. Furthermore, the most active nucleases are also very potent antiproliferative agents towards RT112 and RT112-cis-platin resistant tumour cell lines. Our complexes
ACCEPTED MANUSCRIPT present a higher activity against tumoral cells compared to normal cells, which is particularly important for limiting the toxicity against healthy tissues and side effects. They block the cell cycle at the G1 phase, inducing apoptosis, which could be the consequence of the activation of DNA damage checkpoint. These new copper complexes are promising molecules, since they had a much greater anti-proliferative activity than cis-platin, a conventional agent used for the treatment of metastatic
AC
CE
PT E
D
MA
NU
SC
RI
PT
bladder cancer.
ACCEPTED MANUSCRIPT Aknowledgements The authors wish to acknowledge the LabEx ARCANE (ANR-11-LABX-0003-01), the EPHE, the University Grenoble Alpes (AGIR grant), as well as the Region Rhone-Alpes (ARC Santé) for
AC
CE
PT E
D
MA
NU
SC
RI
PT
financial supports.
ACCEPTED MANUSCRIPT References
AC
CE
PT E
D
MA
NU
SC
RI
PT
[1] Z. Guo, P.J. Sadler, Medicinal Inorganic Chemistry, Adv. Inorg. Chem., 49 (1999) 183-306. [2] B. Lippert, Cisplatin, Chemistry and Biochemistry of a Leading Anticancer Drug, Wiley-VCH, Weinheim, 1999. [3] T. Storr, K.H. Thompson, C. Orvig, Design of targeting ligands in medicinal inorganic chemistry, Chem. Soc. Rev., 35 (2006) 534-544. [4] L. Ronconi, P.J. Sadler, Using coordination chemistry to design new medicines, Coord. Chem. Rev., 251 (2007) 1633-1648. [5] J. Reedijk, Medicinal Applications of Metal Complexes Binding to Biological Macromolecules, Macromolecular Symposia, 270 (2008) 193-201. [6] J. Reedijk, Improved understanding in platinium antitumour chemistry, Chem. Commun., (1996) 801-806. [7] D. Lebwohl, R. Canetta, Clinical development of platinum complexes in cancer therapy: an historical perspective and an update, Eur. J. Cancer, 34 (1998) 1522-1534. [8] D. Wang, S.J. Lippard, Cellular processing of platinum anticancer drugs, Nat Rev Drug Discov, 4 (2005) 307-320. [9] W. Xiaoyong, Fresh Platinum Complexes with Promising Antitumor Activity, Anti-Cancer Agents Med. Chem., 10 (2010) 396-411. [10] M. Cristina, P. Maura, T. Francesco, S. Carlo, Copper Complexes as Anticancer Agents, AntiCancer Agents Med. Chem., 9 (2009) 185-211. [11] T. Saverio, M. Luciano, Copper Compounds in Anticancer Strategies, Curr. Med. Chem., 16 (2009) 1325-1348. [12] F. Tisato, C. Marzano, M. Porchia, M. Pellei, C. Santini, Copper in diseases and treatments, and copper-based anticancer strategies, Med. Res. Rev., 30 (2010) 708-749. [13] C. Duncan, A.R. White, Copper complexes as therapeutic agents, Metallomics, 4 (2012) 127-138. [14] C. Santini, M. Pellei, V. Gandin, M. Porchia, F. Tisato, C. Marzano, Advances in Copper Complexes as Anticancer Agents, Chem. Rev., 114 (2014) 815-862. [15] W. Wong, Copper for breast cancer metastasis, Science, 344 (2014) 1238. [16] R. Galindo-Murillo, J.C. García-Ramos, L. Ruiz-Azuara, I.I.I.T.E. Cheatham, F. Cortés-Guzmán, Intercalation processes of copper complexes in DNA, Nucleic Acids Res., 43 (2015) 5364-5376. [17] K. Garber, Targeting copper to treat breast cancer, Science, 349 (2015) 128-129. [18] D. Screnci, M.J. McKeage, Platinum neurotoxicity: clinical profiles, experimental models and neuroprotective approaches, J. Inorg. Biochem., 77 (1999) 105-110. [19] A. Rajeswaran, A. Trojan, B. Burnand, M. Giannelli, Efficacy and side effects of cisplatin- and carboplatin-based doublet chemotherapeutic regimens versus non-platinum-based doublet chemotherapeutic regimens as first line treatment of metastatic non-small cell lung carcinoma: A systematic review of randomized controlled trials, Lung Cancer, 59 (2008) 1-11. [20] L.R. Kelland, S.Y. Sharp, C.F. O’Neill, F.I. Raynaud, P.J. Beale, I.R. Judson, Mini-review: discovery and development of platinum complexes designed to circumvent cisplatin resistance, J. Inorg. Biochem., 77 (1999) 111-115. [21] M.A. Fuertes, C. Alonso, J.M. Pérez, Biochemical Modulation of Cisplatin Mechanisms of Action: Enhancement of Antitumor Activity and Circumvention of Drug Resistance, Chem. Rev., 103 (2003) 645-662. [22] Z.H. Siddik, Cisplatin: mode of cytotoxic action and molecular basis of resistance, Oncogene, 22 (2003) 7265-7279. [23] M.C. Linder, Biochemistry of Copper, Plenum press, New York, 1991. [24] A. Chan, F. Wong, M. Arumanayagam, Serum Ultrafiltrable Copper, Total Copper and Caeruloplasmin Concentrations in Gynaecological Carcinomas, Ann. Clin. Biochem., 30 (1993) 545549. [25] M. Zowczak, M. Iskra, L. Torliński, S. Cofta, Analysis of serum copper and zinc concentrations in cancer patients, Biol. Trace Elem. Res., 82 (2001) 1.
ACCEPTED MANUSCRIPT
AC
CE
PT E
D
MA
NU
SC
RI
PT
[26] H. Mazdak, F. Yazdekhasti, A. Movahedian, N. Mirkheshti, M. Shafieian, The comparative study of serum iron, copper, and zinc levels between bladder cancer patients and a control group, International Urology and Nephrology, 42 (2010) 89-93. [27] D. Denoyer, S. Masaldan, S. La Fontaine, M.A. Cater, Targeting copper in cancer therapy: 'Copper That Cancer', Metallomics, 7 (2015) 1459-1476. [28] S. Majumder, S. Chatterjee, S. Pal, J. Biswas, T. Efferth, S.K. Choudhuri, The role of copper in drug-resistant murine and human tumors, BioMetals, 22 (2008) 377. [29] X. Huiqi, Y.J. Kang, Role of Copper in Angiogenesis and Its Medicinal Implications, Curr. Med. Chem., 16 (2009) 1304-1314. [30] C. Wende, C. Lüdtke, N. Kulak, Copper Complexes of N-Donor Ligands as Artificial Nucleases, Eur. J. Inorg. Chem., 2014 (2014) 2597-2612. [31] M. Scarpellini, A. Neves, R. Hörner, A.J. Bortoluzzi, B. Szpoganics, C. Zucco, R.A. Nome Silva, V. Drago, A.S. Mangrich, W.A. Ortiz, W.A.C. Passos, M.C.B. de Oliveira, H. Terenzi, Phosphate Diester Hydrolysis and DNA Damage Promoted by New cis-Aqua/Hydroxy Copper(II) Complexes Containing Tridentate Imidazole-rich Ligands, Inorg. Chem., 42 (2003) 8353-8365. [32] L. Zhu, O. dos Santos, C.W. Koo, M. Rybstein, L. Pape, J.W. Canary, Geometry-Dependent Phosphodiester Hydrolysis Catalyzed by Binuclear Copper Complexes, Inorg. Chem., 42 (2003) 79127920. [33] Pattubala A.N. Reddy, M. Nethaji, Akhil R. Chakravarty, Hydrolytic Cleavage of DNA by Ternary Amino Acid Schiff Base Copper(II) Complexes Having Planar Heterocyclic Ligands, Eur. J. Inorg. Chem., 2004 (2004) 1440-1446. [34] Y. An, S.-D. Liu, S.-Y. Deng, L.-N. Ji, Z.-W. Mao, Cleavage of double-strand DNA by linear and triangular trinuclear copper complexes, J. Inorg. Biochem., 100 (2006) 1586-1593. [35] C. Fernandes, G.L. Parrilha, J.A. Lessa, L.J.M. Santiago, M.M. Kanashiro, F.S. Boniolo, A.J. Bortoluzzi, N.V. Vugman, M.H. Herbst, A. Horn Jr, Synthesis, crystal structure, nuclease and in vitro antitumor activities of a new mononuclear copper(II) complex containing a tripodal N3O ligand, Inorg. Chim. Acta, 359 (2006) 3167-3176. [36] M.C.B. Oliveira, D. Mazera, M. Scarpellini, P.C. Severino, A. Neves, H. Terenzi, Mononuclear CuII−Phenolate Bioinspired Complex is Catalytically Promiscuous: Phosphodiester and Peptide Amide Bond Cleavage, Inorg. Chem., 48 (2009) 2711-2713. [37] L. Tjioe, A. Meininger, T. Joshi, L. Spiccia, B. Graham, Efficient Plasmid DNA Cleavage by Copper(II) Complexes of 1,4,7-Triazacyclononane Ligands Featuring Xylyl-Linked Guanidinium Groups, Inorg. Chem., 50 (2011) 4327-4339. [38] P. Jaividhya, R. Dhivya, M.A. Akbarsha, M. Palaniandavar, Efficient DNA cleavage mediated by mononuclear mixed ligand copper(II) phenolate complexes: The role of co-ligand planarity on DNA binding and cleavage and anticancer activity, J. Inorg. Biochem., 114 (2012) 94-105. [39] V.R. de Almeida, F.R. Xavier, R.E.H.M.B. Osorio, L.M. Bessa, E.L. Schilling, T.G. Costa, T. Bortolotto, A. Cavalett, F.A.V. Castro, F. Vilhena, O.C. Alves, H. Terenzi, E.C.A. Eleutherio, M.D. Pereira, W. Haase, Z. Tomkowicz, B. Szpoganicz, A.J. Bortoluzzi, A. Neves, In vitro and in vivo activity of a new unsymmetrical dinuclear copper complex containing a derivative ligand of 1,4,7triazacyclononane: catalytic promiscuity of [Cu2(L)Cl3], Dalton Trans., 42 (2013) 7059-7073. [40] P. Jaividhya, M. Ganeshpandian, R. Dhivya, M.A. Akbarsha, M. Palaniandavar, Fluorescent mixed ligand copper(ii) complexes of anthracene-appended Schiff bases: studies on DNA binding, nuclease activity and cytotoxicity, Dalton Trans., 44 (2015) 11997-12010. [41] Y. Xiao, Q. Wang, Y. Huang, X. Ma, X. Xiong, H. Li, Synthesis, structure, and biological evaluation of a copper(ii) complex with fleroxacin and 1,10-phenanthroline, Dalton Trans., 45 (2016) 10928-10935. [42] X. Zhang, X. Liu, D.L. Phillips, C. Zhao, Mechanistic Insights Into the Factors That Influence the DNA Nuclease Activity of Mononuclear Facial Copper Complexes Containing Hetero-Substituted Cyclens, ACS Catalysis, 6 (2016) 248-257. [43] S. Borah, M.S. Melvin, N. Lindquist, R.A. Manderville, Copper-Mediated Nuclease Activity of a Tambjamine Alkaloid, J. Am. Chem. Soc., 120 (1998) 4557-4562. [44] M.S. Melvin, K.E. Wooton, C.C. Rich, G.R. Saluta, G.L. Kucera, N. Lindquist, R.A. Manderville, Copper-nuclease efficiency correlates with cytotoxicity for the 4-methoxypyrrolic natural products, J. Inorg. Biochem., 87 (2001) 129-135.
ACCEPTED MANUSCRIPT
AC
CE
PT E
D
MA
NU
SC
RI
PT
[45] M. González-Álvarez, G. Alzuet, J.L. García-Giménez, B. Macías, J. Borrás, Biological Activity of Flavonoids Copper Complexes, Z. Anorg. Allg. Chem., 631 (2005) 2181-2187. [46] M.C.B. Oliveira, M.S.R. Couto, P.C. Severino, T. Foppa, G.T.S. Martins, B. Szpoganicz, R.A. Peralta, A. Neves, H. Terenzi, Nucleic acid cleavage by a Cu(II) polyaza macrocyclic complex, Polyhedron, 24 (2005) 495-499. [47] C. Sissi, F. Mancin, M. Gatos, M. Palumbo, P. Tecilla, U. Tonellato, Efficient Plasmid DNA Cleavage by a Mononuclear Copper(II) Complex, Inorg. Chem., 44 (2005) 2310-2317. [48] P.U. Maheswari, S. Roy, H. den Dulk, S. Barends, G. van Wezel, B. Kozlevčar, P. Gamez, J. Reedijk, The Square-Planar Cytotoxic [CuII(pyrimol)Cl] Complex Acts as an Efficient DNA Cleaver without Reductant, J. Am. Chem. Soc., 128 (2006) 710-711. [49] P.U. Maheswari, K. Lappalainen, M. Sfregola, S. Barends, P. Gamez, U. Turpeinen, I. Mutikainen, G.P. van Wezel, J. Reedijk, Structure and DNA cleavage properties of two copper(ii) complexes of the pyridine-pyrazole-containing ligands mbpzbpy and Hmpzbpya, Dalton Trans., (2007) 3676-3683. [50] S.S. Bhat, A.A. Kumbhar, H. Heptullah, A.A. Khan, V.V. Gobre, S.P. Gejji, V.G. Puranik, Synthesis, Electronic Structure, DNA and Protein Binding, DNA Cleavage, and Anticancer Activity of Fluorophore-Labeled Copper(II) Complexes, Inorg. Chem., 50 (2011) 545-558. [51] T. Bortolotto, P.P. Silva, A. Neves, E.C. Pereira-Maia, H. Terenzi, Photoinduced DNA Cleavage Promoted by Two Copper(II) Complexes of Tetracyclines and 1,10-Phenanthroline, Inorg. Chem., 50 (2011) 10519-10521. [52] P.P. Silva, W. Guerra, J.N. Silveira, A.M.d.C. Ferreira, T. Bortolotto, F.L. Fischer, H. Terenzi, A. Neves, E.C. Pereira-Maia, Two New Ternary Complexes of Copper(II) with Tetracycline or Doxycycline and 1,10-Phenanthroline and Their Potential as Antitumoral: Cytotoxicity and DNA Cleavage, Inorg. Chem., 50 (2011) 6414-6424. [53] K. Ghosh, P. Kumar, V. Mohan, U.P. Singh, S. Kasiri, S.S. Mandal, Nuclease Activity via SelfActivation and Anticancer Activity of a Mononuclear Copper(II) Complex: Novel Role of the Tertiary Butyl Group in the Ligand Frame, Inorg. Chem., 51 (2012) 3343-3345. [54] P.P. Silva, W. Guerra, G.C. dos Santos, N.G. Fernandes, J.N. Silveira, A.M. da Costa Ferreira, T. Bortolotto, H. Terenzi, A.J. Bortoluzzi, A. Neves, E.C. Pereira-Maia, Correlation between DNA interactions and cytotoxic activity of four new ternary compounds of copper(II) with N-donor heterocyclic ligands, J. Inorg. Biochem., 132 (2014) 67-76. [55] B. Meunier, Metalloporphyrins as versatile catalysts for oxidation reactions and oxidative DNA cleavage, Chem. Rev., 92 (1992) 1411-1456. [56] L. Li, N.N. Murthy, J. Telser, L.N. Zakharov, G.P.A. Yap, A.L. Rheingold, K.D. Karlin, S.E. Rokita, Targeted Guanine Oxidation by a Dinuclear Copper(II) Complex at Single Stranded/Double Stranded DNA Junctions, Inorg. Chem., 45 (2006) 7144-7159. [57] S. Thyagarajan, N.N. Murthy, A.A. Narducci Sarjeant, K.D. Karlin, S.E. Rokita, Selective DNA Strand Scission with Binuclear Copper Complexes: Implications for an Active Cu2−O2 Species, J. Am. Chem. Soc., 128 (2006) 7003-7008. [58] J.L. García-Giménez, G. Alzuet, M. González-Álvarez, A. Castiñeiras, M. Liu-González, J. Borrás, A Dinuclear Copper(II) Complex with Adeninate Bridge Ligands and Prominent DNA Cleavage Activity. Structural and Spectroscopic Characterization and Magnetic Properties, Inorg. Chem., 46 (2007) 7178-7188. [59] V. Rajendiran, R. Karthik, M. Palaniandavar, V.S. Periasamy, M.A. Akbarsha, B.S. Srinag, H. Krishnamurthy, Mixed-Ligand Copper(II)-phenolate Complexes: Effect of Coligand on Enhanced DNA and Protein Binding, DNA Cleavage, and Anticancer Activity, Inorg. Chem., 46 (2007) 82088221. [60] M. González-Álvarez, G. Alzuet, J. Borrás, L. del Castillo-Agudo, J.M. Montejo-Bernardo, A. Gutiérrez-Rodríguez, S. García-Granda, Evaluation of antiproliferative activities and apoptosis induction caused by copper(II)–benzothiazolesulfonamide complexes in Jurkat T lymphocytes and Caco-2 cells, JBIC Journal of Biological Inorganic Chemistry, 13 (2008) 1249-1265. [61] S. Ramakrishnan, M. Palaniandavar, Interaction of rac-[Cu(diimine)3]2+ and rac[Zn(diimine)3]2+ complexes with CT DNA: effect of fluxional Cu(ii) geometry on DNA binding, ligand-promoted exciton coupling and prominent DNA cleavage, Dalton Trans., (2008) 3866-3878.
ACCEPTED MANUSCRIPT
AC
CE
PT E
D
MA
NU
SC
RI
PT
[62] A. Arbuse, M. Font, M.A. Martínez, X. Fontrodona, M.J. Prieto, V. Moreno, X. Sala, A. Llobet, DNA-Cleavage Induced by New Macrocyclic Schiff base Dinuclear Cu(I) Complexes Containing Pyridyl Pendant Arms, Inorg. Chem., 48 (2009) 11098-11107. [63] Y. Kou, J. Tian, D. Li, W. Gu, X. Liu, S. Yan, D. Liao, P. Cheng, Synthesis, structure, magnetic properties and DNA cleavage of binuclear Cu(ii) Schiff-base complexes, Dalton Trans., (2009) 23742382. [64] S. Ramakrishnan, V. Rajendiran, M. Palaniandavar, V.S. Periasamy, B.S. Srinag, H. Krishnamurthy, M.A. Akbarsha, Induction of Cell Death by Ternary Copper(II) Complexes of lTyrosine and Diimines: Role of Coligands on DNA Binding and Cleavage and Anticancer Activity, Inorg. Chem., 48 (2009) 1309-1322. [65] X. Dong, X. Wang, M. Lin, H. Sun, X. Yang, Z. Guo, Promotive Effect of the Platinum Moiety on the DNA Cleavage Activity of Copper-Based Artificial Nucleases, Inorg. Chem., 49 (2010) 25412549. [66] K. Ghosh, P. Kumar, N. Tyagi, U.P. Singh, V. Aggarwal, M.C. Baratto, Synthesis and reactivity studies on new copper(II) complexes: DNA binding, generation of phenoxyl radical, SOD and nuclease activities, European Journal of Medicinal Chemistry, 45 (2010) 3770-3779. [67] S. Ramakrishnan, D. Shakthipriya, E. Suresh, V.S. Periasamy, M.A. Akbarsha, M. Palaniandavar, Ternary Dinuclear Copper(II) Complexes of a Hydroxybenzamide Ligand with Diimine Coligands: the 5,6-dmp Ligand Enhances DNA Binding and Cleavage and Induces Apoptosis, Inorg. Chem., 50 (2011) 6458-6471. [68] R. Loganathan, S. Ramakrishnan, E. Suresh, A. Riyasdeen, M.A. Akbarsha, M. Palaniandavar, Mixed Ligand Copper(II) Complexes of N,N-Bis(benzimidazol-2-ylmethyl)amine (BBA) with Diimine Co-Ligands: Efficient Chemical Nuclease and Protease Activities and Cytotoxicity, Inorg. Chem., 51 (2012) 5512-5532. [69] C. Rajarajeswari, R. Loganathan, M. Palaniandavar, E. Suresh, A. Riyasdeen, M.A. Akbarsha, Copper(ii) complexes with 2NO and 3N donor ligands: synthesis, structures and chemical nuclease and anticancer activities, Dalton Trans., 42 (2013) 8347-8363. [70] M.-J. Li, T.-Y. Lan, X.-H. Cao, H.-H. Yang, Y. Shi, C. Yi, G.-N. Chen, Synthesis, characterization, DNA binding, cleavage activity and cytotoxicity of copper(ii) complexes, Dalton Trans., 43 (2014) 2789-2798. [71] R. Loganathan, S. Ramakrishnan, E. Suresh, M. Palaniandavar, A. Riyasdeen, M.A. Akbarsha, Mixed ligand [small mu ]-phenoxo-bridged dinuclear copper(ii) complexes with diimine co-ligands: efficient chemical nuclease and protease activities and cytotoxicity, Dalton Trans., 43 (2014) 61776194. [72] R. Loganathan, S. Ramakrishnan, M. Ganeshpandian, N.S.P. Bhuvanesh, M. Palaniandavar, A. Riyasdeen, M.A. Akbarsha, Mixed ligand copper(ii) dicarboxylate complexes: the role of co-ligand hydrophobicity in DNA binding, double-strand DNA cleavage, protein binding and cytotoxicity, Dalton Trans., 44 (2015) 10210-10227. [73] N. Morellato, C. Musetti, M. Palumbo, A. Dolmella, C. Bolzati, C. Sissi, Biologically relevant Cu2+ complexes with a tridentate phenanthroline analogue: Structural properties and DNA damage, Inorg. Chim. Acta, 434 (2015) 127-134. [74] Q. Gan, C.-L. Zhang, B.-F. Wang, Y.-H. Xiong, Y.-L. Fu, Z.-W. Mao, X.-Y. Le, Two new mixed copper(ii)-dipeptide complexes of N,N-donor heterocycle ligands: studies on their non-covalent DNA binding, chemical nuclease, antioxidant and anticancer activities, RSC Advances, 6 (2016) 3595235965. [75] U. Golla, A. Adhikary, A.K. Mondal, R.S. Tomar, S. Konar, Synthesis, structure, magnetic and biological activity studies of bis-hydrazone derived Cu(ii) and Co(ii) coordination compounds, Dalton Trans., 45 (2016) 11849-11863. [76] S. Kathiresan, S. Mugesh, M. Murugan, F. Ahamed, J. Annaraj, Mixed-ligand copper(ii)phenolate complexes: structure and studies on DNA/protein binding profiles, DNA cleavage, molecular docking and cytotoxicity, RSC Advances, 6 (2016) 1810-1825. [77] W.-J. Lian, X.-T. Wang, C.-Z. Xie, H. Tian, X.-Q. Song, H.-T. Pan, X. Qiao, J.-Y. Xu, Mixedligand copper(ii) Schiff base complexes: the role of the co-ligand in DNA binding, DNA cleavage, protein binding and cytotoxicity, Dalton Trans., 45 (2016) 9073-9087.
ACCEPTED MANUSCRIPT
AC
CE
PT E
D
MA
NU
SC
RI
PT
[78] N. Berthet, V. Martel-Frachet, F. Michel, C. Philouze, S. Hamman, X. Ronot, F. Thomas, Nuclease and anti-proliferative activities of copper(ii) complexes of N3O tripodal ligands involving a sterically hindered phenolate, Dalton Trans., 42 (2013) 8468-8483. [79] G. Iacomino, G. Picariello, L. D'Agostino, DNA and nuclear aggregates of polyamines, Biochimica et Biophysica Acta (BBA) - Molecular Cell Research, 1823 (2012) 1745-1755. [80] E.W. Gerner, F.L. Meyskens, Polyamines and cancer: old molecules, new understanding, Nat Rev Cancer, 4 (2004) 781-792. [81] Y. Huang, J.C. Keen, E. Hager, R. Smith, A. Hacker, B. Frydman, A.L. Valasinas, V.K. Reddy, L.J. Marton, R.A. Casero, N.E. Davidson, Regulation of Polyamine Analogue Cytotoxicity by c-Jun in Human MDA-MB-435 Cancer Cells, Mol. Cancer Res., 2 (2004) 81-88. [82] C.N. N’soukpoé-Kossi, A.A. Ouameur, T. Thomas, A. Shirahata, T.J. Thomas, H.A. TajmirRiahi, DNA Interaction with Antitumor Polyamine Analogues: A Comparison with Biogenic Polyamines, Biomacromolecules, 9 (2008) 2712-2718. [83] F. Secco, M. Venturini, T. Biver, F. Sánchez, R. Prado-Gotor, E. Grueso, Solvent Effects on the Kinetics of the Interaction of 1-Pyrenecarboxaldehyde with Calf Thymus DNA, J. Phys. Chem. B, 114 (2010) 4686-4691. [84] S. Gentil, D. Serre, C. Philouze, M. Holzinger, F. Thomas, A. Le Goff, Electrocatalytic O2 Reduction at a Bio-inspired Mononuclear Copper Phenolato Complex Immobilized on a Carbon Nanotube Electrode, Angew. Chem. Int. Ed., 55 (2016) 2517-2520. [85] F. Michel, F. Thomas, S. Hamman, E. Saint-Aman, C. Bucher, J.-L. Pierre, Galactose Oxidase Models: Solution Chemistry, and Phenoxyl Radical Generation Mediated by the Copper Status, Chem. – Eur. J., 10 (2004) 4115-4125. [86] K. Ohara, M. Smietana, A. Restouin, S. Mollard, J.-P. Borg, Y. Collette, J.-J. Vasseur, Amine−Guanidine Switch: A Promising Approach to Improve DNA Binding and Antiproliferative Activities, J. Med. Chem., 50 (2007) 6465-6475. [87] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. J. A. Montgomery, J.E. Peralta, F. Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, Ö. Farkas, J.B. Foresman, J. V.Ortiz, J. Cioslowski, D.J. Fox, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2009. [88] J.P. Perdew, Density-functional approximation for the correlation energy of the inhomogeneous electron gas, Phys. Rev. B, 33 (1986) 8822-8824. [89] C. Lee, W. Yang, R.G. Parr, Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density, Phys. Rev. B: Condens. Matter Mater. Phys., 37 (1988) 785-789. [90] A.D. Becke, Density‐functional thermochemistry. III. The role of exact exchange, J. Chem. Phys., 98 (1993) 5648-5652. [91] M.E. Casida, in: D.P. Chong (Ed.) Recent Advances in Density Functional Methods, World Scientific, Singapore, 1995. [92] J. Tomasi, B. Mennucci, R. Cammi, Quantum Mechanical Continuum Solvation Models, Chem. Rev., 105 (2005) 2999-3094. [93] D.M. Hinton, V.C. Bode, Ethidium binding affinity of circular lambda deoxyribonucleic acid determined fluorometrically, J. Biol. Chem., 250 (1975) 1061-1067. [94] 23 Despite that 1 could be isolated as single crystals for structural characterization (see ref. 84) it was not possible to crystallize the complexes 2-4, presumably because of the large flexibility of the putrescine arm. It must however be emphasized that the geometry observed in 1, i.e. square pyramidal copper ion with an equatorial phenolate and an exogenous ligand (chloride) that completes the coordination sphere, provides a good model for the coordination polyhedron in the other ligands. [95] A. Philibert, F. Thomas, C. Philouze, S. Hamman, E. Saint-Aman, J.-L. Pierre, Galactose Oxidase Models: Tuning the Properties of CuII–Phenoxyl Radicals, Chem. – Eur. J., 9 (2003) 3803-3812.
ACCEPTED MANUSCRIPT
AC
CE
PT E
D
MA
NU
SC
RI
PT
[96] F. Michel, S. Hamman, F. Thomas, C. Philouze, I. Gautier-Luneau, J.-L. Pierre, Galactose Oxidase models: 19F NMR as a powerful tool to study the solution chemistry of tripodal ligands in the presence of copper(ii), Chem. Commun., (2006) 4122-4124. [97] F. Michel, F. Thomas, S. Hamman, C. Philouze, E. Saint-Aman, J.-L. Pierre, Galactose Oxidase Models: Creation and Modification of Proton Transfer Coupled to Copper(II) Coordination Processes in Pro-Phenoxyl Ligands, Eur. J. Inorg. Chem., 2006 (2006) 3684-3696. [98] F. Michel, S. Hamman, C. Philouze, C.P. Del Valle, E. Saint-Aman, F. Thomas, Galactose oxidase models: insights from 19F NMR spectroscopy, Dalton Trans., (2009) 832-842. [99] The relative affinity of the copper ion for the ligands was tentatively assessed by a competition experiment with nitrilotriacetic acid (NTA). At one molar equivalent of NTA added the intensity of the CT transition remains >90% of the initial one, disclosing a much higher affinity of the tripodal ligands for the copper ion. For higher NTA concentration a rapid precipitation of the tripodal ligand is observed, precluding any accurate determination of the binding constant. [100] A.W. Addison, T.N. Rao, J. Reedijk, J. van Rijn, G.C. Verschoor, Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen-sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2[prime or minute]-yl)-2,6dithiaheptane]copper(II) perchlorate, J. Chem. Soc., Dalton Trans., (1984) 1349-1356. [101] F. Thomas, Ten Years of a Biomimetic Approach to the Copper(II) Radical Site of Galactose Oxidase, Eur. J. Inorg. Chem., (2007) 2379-2404. [102] M. Masui, H. Sayo, Y. Tsuda, Anodic oxidation of amines. Part I. Cyclic voltammetry of aliphatic amines at a stationary glassy-carbon electrode, J. Chem. Soc. B: Phys. Org., (1968) 973-976. [103] M. Hasegawa, S. Iwata, Y. Sone, J. Endo, H. Matsuzawa, Y. Mazaki, Chiroptical Properties and the Racemization of Pyrene and Tetrathiafulvalene-Substituted Allene: Substitution and Solvent Effects on Racemization in Tetrathiafulvalenylallene, Molecules, 19 (2014) 2829. [104] J.D. McGhee, P.H. von Hippel, Theoretical aspects of DNA-protein interactions: Co-operative and non-co-operative binding of large ligands to a one-dimensional homogeneous lattice, J. Mol. Biol., 86 (1974) 469-489. [105] A.A. Ouameur, H.-A. Tajmir-Riahi, Structural Analysis of DNA Interactions with Biogenic Polyamines and Cobalt(III)hexamine Studied by Fourier Transform Infrared and Capillary Electrophoresis, J. Biol. Chem., 279 (2004) 42041-42054. [106] The fluorescence of EtBr is not altered by the addition of the complexes in the absence of DNA. [107] A. Kabir, G. Suresh Kumar, Binding of the Biogenic Polyamines to Deoxyribonucleic Acids of Varying Base Composition: Base Specificity and the Associated Energetics of the Interaction, PLOS ONE, 8 (2013) e70510. [108] L.C. Gosule, J.A. Schellman, Compact form of DNA induced by spermidine, Nature, 259 (1976) 333-335. [109] J. Pelta Jr, D. Durand, J. Doucet, F. Livolant, DNA mesophases induced by spermidine: structural properties and biological implications, Biophys. J., 71 (1996) 48-63. [110] V.A. Bloomfield, DNA condensation by multivalent cations, Biopolymers, 44 (1997) 269-282. [111] M.M. Patel, T.J. Anchordoquy, Contribution of Hydrophobicity to Thermodynamics of LigandDNA Binding and DNA Collapse, Biophys. J., 88 (2005) 2089-2103. [112] In order to ensure that the reaction was complete within 1 hour we investigated the kinetics of the DNA cleavage in the presence of ascorbate (see ESI) for four representative complexes. The reaction was found to be extremely rapid, since it is complete within 2 minutes for both 1 and 2. [113] J. Ruiz-Chica, M.A. Medina, F. Sánchez-Jiménez, F.J. Ramírez, Fourier Transform Raman Study of the Structural Specificities on the Interaction between DNA and Biogenic Polyamines, Biophys. J., 80 (2001) 443-454.
ACCEPTED MANUSCRIPT Figure captions Figure 1 Electrospray mass spectra of the in situ generated copper complexes (a) 1; (b) 2; (c) 3; (d) 4. The complexes were obtained by mixing the ligand with the appropriate amount of copper(II) chloride in DMF and further dilution with water at pH 7 for analysis. Figure 2 pH-dependence of the electronic spectrum of 2 in a H2O: DMF 90: 10 solution. [NaCl] = 0.1
PT
M, T = 298 K, l = cm. The pH varies from 2.56 to 6.94 and the arrow indicates the spectral change
RI
upon increasing the pH.
SC
Figure 3 X-Band EPR spectra in a water: DMF 90:10 medium at pH 7. (a) 1; (b) 2; (c) 3; (d) simulation by using the parameters given in Table 2 for 2; (e) 4. [Tris-HCl] = 0.05 M, [NaCl] = 0.02
NU
M; T, 100 K; microwave frequency, 9.44 GHz; microwave power, 4 mW; mod. Freq.,100 KHz; mod. Amp. 0.4 mT.
MA
Figure 4 Geometry optimized structures for the archetypes: (a) [(HLarch)Cu(H2O)]2+ with an axial phenol; (b) [(HLarch)Cu(H2O)]2+ with an equatorial phenol; (c) [(Larch)Cu(H2O)]+ with an axial
D
phenolate (as initial guess); (d) [(Larch)Cu(H2O)]+ with an equatorial phenolate.
PT E
Figure 5 Cyclic voltammetry curves in a water:DMF 90:10 medium at pH 7. (a) 1; (b) PyrPt; (c) 2; (d) 3. [Tris-HCl] = 0.05 M, [NaCl] = 0.02 M; T, 298 K; scan rate, 0.1 V sec-1; The potentials are given
CE
versus the SCE reference.
AC
Figure 6 Scatchard – Von Hippel plot for the binding of complex 2 to CT DNA (from fluorescence data). Water: DMF 90:10 medium,[Tris-HCl] = 0.05 M, [NaCl] = 0.02 M; T, 298 K. Figure 7 Variation in the ethidium bromide (EtBr) fluorescence upon addition of complex 2; (a) Fluorescence spectra, the arrow indicates changes upon addition of 0-50 M of complex 2 in the presence of 20 M EtBr. Water: DMF 90:10 medium at pH 7; [Tris-HCl] = 0.05 M, [NaCl] = 0.02 M; T, 298 K;exc = 527 nm. (b) Plot of the fluorescence intensity as a function of the [complex]/[EtBr] ratio.
ACCEPTED MANUSCRIPT Figure 8 Agarose gel electrophoresis patterns of the cleavage reaction of X174 supercoiled DNA (20 M base pairs) mediated by the copper complexes in a phosphate buffer 10 mM pH 7.2 (+ 10 % DMF) at 37°C for 1 h, in the absence of reductant. Abbreviations: NC: nicked circular, SC: supercoiled. The concentrations of the complexes indicated on the top are in M. (a) 1; (b) 2; (c) 3; (d) 4; (e) 3(Boc); (f) Expended view of the region 10-30 M in the case of 2.
PT
Figure 9 Agarose gel electrophoresis patterns of the cleavage of X174 supercoiled DNA (20 M
RI
base pairs) mediated by the copper complexes in a phosphate buffer 10 mM pH 7.2 (+ 10 % DMF) at 37°C for 1 h in the presence of reductant. [ascorbate] = 0.8 mM. Abbreviations: NC: nicked circular,
SC
SC: supercoiled, L: linear. The concentrations of the complexes indicated on the top are in M. (a) 2;
NU
(b) 3; (c) 4.
Figure 10 Agarose gel electrophoresis patterns of supercoiled X174 DNA (20 M base pairs)
MA
incubated with the copper complexes in a phosphate buffer 10 mM pH 7.2 (+ 10 % DMF) at 37°C for 1 h; [ascorbate] = 0.8 mM; (a) 1, 50 M; (b) 2, 8 M; (c) 3, 30 M. Lane 0, DNA control; lane 1,
D
DNA + complex (without scavenger); lanes 2-10, DNA + complex in the presence of agents: lane 2,
PT E
ethanol; lane 3, DMSO (2 L); lane 4, Superoxide Dismutase (0.5 unit); lane 5, NaN3 (100 M); lane 6, EDTA (10 mM); lane 7, Catalase (0.1 unit); lane 8, Hoechst 33258 (100 M); lane 9, NaCl (350
CE
M).
Figure 11 Flow cytometry analysis after 48 h treatment with the copper complexes. Colour code:
AC
Green, late apoptosis; red, early apoptosis; blue, viable cells (necrosis corresponds to less than 0.3 % of the cells). (a) Control; (b) 2, 2.5 M; (c) 2, 3.75 M; (d) 4, 2 M; (e) 4, 3 M; (f) 1, 5 M; (g) 1, 7.5 M. Figure 12 Cell distributions after 24h treatment of RT112 cells with the copper complexes at a concentration corresponding to the IC50. (a) Control without complex; (b) treatment with 1; (c) treatment with 2; (d) treatment with 4. Colour code: Green, phases G2, M; Red, phase S; Blue, phases G0, G1.
ACCEPTED MANUSCRIPT Tables:
Complex pKa
phenolate formb
phenol formc
max
max
d
-d
1
2.47±0.01
481
790
-
2
3.93±0.02
460
1020
680
33
3
3.77±0.01
461
780
690
35
4
7.97±0.03e 457
2000
684
172
PT
Table 1. pKa values and electronic spectra of the copper complexesa
RI
a
SC
Fitted with the SPECFIT software. Data obtained in a H2O:DMF 90:10 solution, [NaCl] = 0.1 M. The pH is adjusted by addition of HClO4 or NaOH. The number in bracket corresponds to the standard deviation in the pKa value for the equilibrium depicted in Scheme 3. b
NU
Phenolate-to-copper charge transfer transition. A broad shoulder is observed at lower energy (ca. 700 nm) with a much lower intensity, which corresponds to the d-d transitions (not indicated in the table). Copper d-d transition.
d
At the lowest pH investigated (1.8) the phenolate-to-copper charge transfer was still observable.
MA
c
The value indicated corresponds to the -logβ2 value associated to the equilibrium: [(H2L4)Cu2]4+ H+ + [(L4)Cu2]2+. The fit was not improved by considering two stepwise protonation steps.
AC
CE
PT E
D
e
2
ACCEPTED MANUSCRIPT
4
g/A e
2.053 2.250
1.4
17.0
126
2.056 2.261
2.0
17.0
127
3.60 d
2.055 2.262
1.4
16.4
132
7b
2.051 2.248
1.4
17.0
126
c
2.055 2.262
1.8
17.1
126
4.99 d
2.051 2.258
1.5
16.2
133
7b
2.053 2.247
1.7
17.0
126
2.78 c
2.055 2.264
1.8
16.9
128
4.78 d
2.054 2.260
1.4
16.4
132
7b
2.053 2.247
1.7
17.0
126
3.07 c
2.054 2.262
1.8
17.1
126
5.06 d
2.052 2.258
1.5
16.2
2.97
3
A//
c
1.62
2
A_|_
RI
7b
g//
NU
1
g_|_
SC
Complex pH
PT
Table 2. Spin Hamiltonian parameters for the copper complexesa
133
In H2O:DMF 90:10 frozen solution. A for Cu (I = 3/2) in mT.
b
[Tris-HCl] = 0.05 M, [NaCl] = 0.02 M.
c
[NaCl] = 0.1 M. The pH is adjusted by addition of HClO4 to the value corresponding to ca. pKa-1.
d
[NaCl] = 0.1 M. The pH is adjusted by addition of NaOH to the value corresponding to ca. pKa+1.
e
ratio expressed in cm.
D
PT E
CE AC
MA
a
ACCEPTED MANUSCRIPT Table 3. Electrochemical properties of the compounds at pH 7 a Epc,red
Epa,red
Epa,ox
1
-0.46
-0.12
0.96
2
-0.5br
0.07
0.57, 0.69,0.8b
3
-0.46
-0.17
0.68,0.9b
4
-0.5br
0.08
0.57, 0.68, 0.85b
PyrPt
-
-
0.86, 1.0b
PT
Complex
a
CE
PT E
D
MA
NU
SC
Ill-defined.
AC
b
RI
In a H2O:DMF 90:10 solution, [Tris-HCl] = 0.05 M, [NaCl] = 0.02 M. Reference, SCE; T, 298 K; Br: broad.
ACCEPTED MANUSCRIPT Table 4. Binding constants at pH 7 Compound
K (M-1)
Method
Cu-N3O tripod
a
Direct
Kapp (M-1)
Ref
2.0×10
3
[78]
2.6×10
6
This work b
Direct
3
Competition Competition
6×106 c 5.5×105 c
This work b This work b
4
Competition
2.7×106
This work b
Putrescine
Mobility
1.0×105 c
PyrPt
Competition Direct 1.1×107
[101] [86] d
1.7×106 c
EtBr
Competition Direct 6.9×105
This work b [87]
1.0×105
SC
RI
[99]
PT
2
Cu-N3O tripod: copper(II) complex of the 2-tert-butyl-4-methoxy-6-phenol.
b
In a buffered H2O:DMF 90:10 solution, [Tris-HCl] = 0.05 M, [NaCl] = 0.02 M, T = 298 K, pH = 7.
c
Calculated by using a K value for EtBr of 4.9 105 (reference 87).
d
10 mM sodium cacodylate buffer, pH 7.
AC
CE
PT E
D
MA
NU
a
ACCEPTED MANUSCRIPT Table 5. Nuclease activity at pH 7.2 a Complex
Mercaptan Ascorbic acid
1
> 50
45
2
14
1.7
3
32
14
4
b
-b
-
Expressed as the concentration of complex (in M) that produces 50 % of cleavage of X174 supercoiled DNA. [DNA] = 20 M base pairs; T = 37°C; t = 1 h; [reductant] =0.8 mM; phosphate buffer 10 mM; pH = 7.2; water:DMF (90:10).
RI
PT
a
b
AC
CE
PT E
D
MA
NU
SC
The gel electrophoresis displays only the disappearance of the SC form, due to significant fragmentation or condensation of DNA (see the text). The IC50 value cannot be determined under these conditions.
ACCEPTED MANUSCRIPT Table 6. Antiproliferative activities of the complexesa
VERO c
RT112
RT112-CP
Cis-platin
ND
9.1 ± 1.2
23.8 ± 1.2
2.6
1
11.1 ± 1.5
3.1 ± 0.3
5.5 ± 0.2
1.8
2
3.1 ± 1.0
2.0 ± 0.2
1.5 ± 0.1
0.75
3
14.7 ± 1.4
9.7 ± 1.1
9.7 ± 0.6
1.0
4
3.2 ± 0.6
1.4 ± 0.3
1.3 ± 0.1
0.9
The concentration resulting in 50% loss of cell viability relative to untreated cells (IC50) was
RI
a
PT
RFb
IC50 (µM)
Complexes
SC
determined from dose-response curves. Results represent means ± SD of three independent experiments. RF is the relative ratio of IC50 in both the cis-platin-resistant and the sensitive cell lines. RF is the relative ratio of IC50 in both the cis-platin-resistant and the sensitive cell lines.
c
Not determined.
AC
CE
PT E
D
MA
NU
b
ACCEPTED MANUSCRIPT
D
MA
NU
SC
RI
PT
Schemes and figures:
AC
CE
PT E
Scheme 1 Formula of the organic compounds of interest in this study.
ACCEPTED MANUSCRIPT
AC
CE
PT E
D
MA
NU
SC
RI
PT
Scheme 2 in situ generated complexes.
ACCEPTED MANUSCRIPT
AC
CE
PT E
D
MA
NU
SC
RI
PT
Scheme 3 General structures and protonation equilibrium of the complexes (Ex: H2O or Cl-).
ACCEPTED MANUSCRIPT
2x106
3x107
(b) +
1x106
5x105
+
2x107
1x107
0 458 460 462 464 466 468 470 472 474
0 780
785
m/z
SC
6x106
NU
Intensity (a. u.)
+
8x106
2x105
4x106 2x106
0 266
268
270
272
0 1085
1090
1095
1100
1105
1110
m/z
D
m/z
MA
Intensity (a. u.)
4x10
5
(d)
2+
RI
1x107
(c) 6x10
795
m/z
8x105
5
790
PT
Intensity (a. u.)
Intensity (a. u.)
(a)
PT E
Figure 1 Electrospray mass spectra of the in situ generated copper complexes (a) 1; (b) 2; (c) 3; (d) 4. The complexes were obtained by mixing the ligand with the appropriate amount of copper(II) chloride
AC
CE
in DMF and further dilution with water at pH 7 for analysis.
ACCEPTED MANUSCRIPT
1.4 1.2 1.0
A
0.8 0.6
PT
0.4 0.2 0.0 500
600
700
800
RI
400
SC
/ nm
Figure 2 pH-dependence of the electronic spectrum of 2 in a H2O: DMF 90: 10 solution. [NaCl] = 0.1
NU
M, T = 298 K, l = cm. The pH varies from 2.56 to 6.94 and the arrow indicates the spectral change
AC
CE
PT E
D
MA
upon increasing the pH.
ACCEPTED MANUSCRIPT
(a) (b) (c) d'' dB
280
300
320
340
360
SC
B / mT
RI
260
PT
(d) (e)
NU
Figure 3 X-Band EPR spectra in a water: DMF 90:10 medium at pH 7. (a) 1; (b) 2; (c) 3; (d) simulation by using the parameters given in Table 2 for 2; (e) 4. [Tris-HCl] = 0.05 M, [NaCl] = 0.02
MA
M; T, 100 K; microwave frequency, 9.44 GHz; microwave power, 4 mW; mod. Freq.,100 KHz; mod.
AC
CE
PT E
D
Amp. 0.4 mT.
MA
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
Figure 4 Formula of the archetype [(HLarch)Cu(H2O)]2+ and geometry optimized structures for the archetypes: (a) [(HLarch)Cu(H2O)]2+ with an axial phenol; (b) [(HLarch)Cu(H2O)]2+ with an equatorial
AC
CE
equatorial phenolate.
PT E
D
phenol; (c) [(Larch)Cu(H2O)]+ with an axial phenolate (as initial guess); (d) [(Larch)Cu(H2O)]+ with an
ACCEPTED MANUSCRIPT
(a) (b)
PT
I/A (c)
-1.0
-0.5
0.0
0.5
1.0
SC
-1.5
RI
(d)
E/V
NU
Figure 5 Cyclic voltammetry curves in a water:DMF 90:10 medium at pH 7. (a) 1; (b) PyrPt; (c) 2; (d)
MA
3. [Tris-HCl] = 0.05 M, [NaCl] = 0.02 M; T, 298 K; scan rate, 0.1 V sec-1; The potentials are given
AC
CE
PT E
D
versus the SCE reference.
ACCEPTED MANUSCRIPT
36000
32000
r/c
28000
24000
0.010
0.012
0.014
PT
20000
0.016
SC
RI
r
Figure 6 Scatchard – Von Hippel plot for the binding of complex 2 to CT DNA (from fluorescence
AC
CE
PT E
D
MA
NU
data). Water: DMF 90:10 medium, [Tris-HCl] = 0.05 M, [NaCl] = 0.02 M; T, 298 K.
450
600
(a)
500
Fluorescence / a.u.
Fluorescence (a. u.)
ACCEPTED MANUSCRIPT
400 300 200 100
(b) 400 350 300 250
600
650
700
750
0.0
800
(nm)
0.2
0.4
PT
200 0 550
0.6
0.8
1.0
RI
[complex]/[EtBr]
Figure 7 Variation in the ethidium bromide (EtBr) fluorescence upon addition of complex 2; (a)
SC
Fluorescence spectra, the arrow indicates changes upon addition of 0-50 M of complex 2 in the
NU
presence of 20 M EtBr. Water: DMF 90:10 medium at pH 7; [Tris-HCl] = 0.05 M, [NaCl] = 0.02 M; T, 298 K;exc = 527 nm. (b) Plot of the fluorescence intensity as a function of the [complex]/[EtBr]
AC
CE
PT E
D
MA
ratio.
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
MA
Figure 8 Agarose gel electrophoresis patterns of the cleavage reaction of X174 supercoiled DNA (20 M base pairs) mediated by the copper complexes in a phosphate buffer 10 mM pH 7.2 (+ 10 %
D
DMF) at 37°C for 1 h, in the absence of reductant. Abbreviations: NC: nicked circular, SC:
PT E
supercoiled. The concentrations of the complexes indicated on the top are in M. (a) 1; (b) 2; (c) 3; (d)
AC
CE
4; (e) 3(Boc); (f) Expended view of the region 10-30 M in the case of 2.
SC
RI
PT
ACCEPTED MANUSCRIPT
Figure 9 Agarose gel electrophoresis patterns of the cleavage of X174 supercoiled DNA (20 M
NU
base pairs) mediated by the copper complexes in a phosphate buffer 10 mM pH 7.2 (+ 10 % DMF) at 37°C for 1 h in the presence of reductant. [ascorbate] = 0.8 mM. Abbreviations: NC: nicked circular,
MA
SC: supercoiled, L: linear. The concentrations of the complexes indicated on the top are in M. (a) 2;
AC
CE
PT E
D
(b) 3; (c) 4.
PT
ACCEPTED MANUSCRIPT
RI
Figure 10 Agarose gel electrophoresis patterns of supercoiled X174 DNA (20 M base pairs) incubated with the copper complexes in a phosphate buffer 10 mM pH 7.2 (+ 10 % DMF) at 37°C for
SC
1 h; [ascorbate] = 0.8 mM; (a) 1, 50 M; (b) 2, 8 M; (c) 3, 30 M. Lane 0, DNA control; lane 1, DNA + complex (without scavenger); lanes 2-10, DNA + complex in the presence of agents: lane 2,
NU
ethanol; lane 3, DMSO (2 L); lane 4, Superoxide Dismutase (0.5 unit); lane 5, NaN3 (100 M); lane
MA
6, EDTA (10 mM); lane 7, Catalase (0.1 unit); lane 8, Hoechst 33258 (100 M); lane 9, NaCl (350
AC
CE
PT E
D
M).
ACCEPTED MANUSCRIPT
4% 8% 11% 20% 15% 19% 15%
80
31%
27%
28%
60
46%
25%
57% 40
88% 59%
58%
57%
20
23%
PT
%
32%
35%
44%
(f) (g)
NU
(a) (b) (c) (d) (e)
SC
0
RI
100
Figure 11 Flow cytometry analysis after 48 h treatment with the copper complexes. Colour code:
MA
Green, late apoptosis; red, early apoptosis; blue, viable cells (necrosis corresponds to less than 0.3 % of the cells). (a) Control; (b) 2, 2.5 M; (c) 2, 3.75 M; (d) 4, 2 M; (e) 4, 3 M; (f) 1, 5 M; (g) 1,
AC
CE
PT E
D
7.5 M.
ACCEPTED MANUSCRIPT
100
25% 80 60
24%
15%
15%
13%
12%
72%
73%
(c)
(d)
29%
23%
40
51%
47%
(a)
(b)
RI
20
PT
%
NU
SC
0
Figure 12 Cell distributions after 24h treatment of RT112 cells with the copper complexes at a
MA
concentration corresponding to the IC50. (a) Control without complex; (b) treatment with 1; (c) treatment with 2; (d) treatment with 4. Colour code: Green, phases G2, M; Red, phase S; Blue, phases
AC
CE
PT E
D
G0, G1.
ACCEPTED MANUSCRIPT
Copper(II) complexes of N3O tripodal ligands appended with pyrene and polyamine groups: Antiproliferative and nuclease activities Doti Serre, Sule Erbek, Nathalie Berthet, Christian Philouze, Xavier Ronot, Véronique Martel-Frachet and Fabrice Thomas*
PT E
D
MA
NU
SC
RI
PT
The pyrene and polyamine groups enhance the affinity of the copper complexes towards DNA. The complexes exhibit a high nuclease activity in the presence of reductant, and show a much greater anti-proliferative activity than cis-platin towards bladder cancer cells.
AC
CE
Graphical abstract