Open Access Article
Feng
Chen
a,
Joan J.
Soldevila-Barreda
a,
Isolda
Romero-Canelón
ab,
James P. C.
Coverdale
a,
Ji-Inn
Song
a,
Guy J.
Clarkson
a,
Jana
Kasparkova
c,
Abraha
Habtemariam
a,
Viktor
Brabec
c,
Juliusz A.
Wolny
d,
Volker
Schünemann
d and
Peter J.
Sadler
*a
aDepartment of Chemistry, University of Warwick, Gibbet Hill Road, Coventry CV4 7AL, UK. E-mail: P.J.Sadler@warwick.ac.uk
bSchool of Pharmacy, University of Birmingham, Birmingham B15 2TT, UK
cDepartment of Biophysics, Faculty of Science, Palacky University, 17. listopadu 12, CZ-77146 Olomouc, Czech Republic
dDepartment of Physics, University of Kaiserslautern, Erwin-Schrödinger-Str. 46, 67663 Kaiserslautern, Germany
First published on 13th April 2018
A series of neutral pseudo-octahedral RuII sulfonamidoethylenediamine complexes [(η6-p-cym)Ru(N,N′)Cl] where N,N′ is N-(2-(R1,R2-amino)ethyl)-4-toluenesulfonamide (TsEn(R1,R2)) R1,R2 = Me,H (1); Me,Me (2); Et,H (3); benzyl,H (Bz, 4); 4-fluorobenzyl,H (4-F-Bz, 5) or naphthalen-2-ylmethyl,H (Naph, 6), were synthesised and characterised including the X-ray crystal structure of 3. These complexes catalyse the reduction of NAD+ regioselectively to 1,4-NADH by using formate as the hydride source. The catalytic efficiency depends markedly on the steric and electronic effects of the N-substitutent, with turnover frequencies (TOFs) increasing in the order: 1 < 2 < 3, 6 < 4, 5, achieving a TOF of 7.7 h−1 for 4 with a 95% yield of 1,4-NADH. The reduction rate was highest between pH* (deuterated solvent) 6 and 7.5 and improved with an increase in formate concentration (TOF of 18.8 h−1, 140 mM formate). The calculations suggested initial substitution of an aqua ligand by formate, followed by hydride transfer to RuII and then to NAD+, and indicated specific interactions between the aqua complex and both NAD+ and NADH, the former allowing a preorganisation involving interaction between the aqua ligand, formate anion and the pyridine ring of NAD+. The complexes exhibited antiproliferative activity towards A2780 human ovarian cancer cells with IC50 values ranging from 1 to 31 μM, the most potent complex, [(η6-p-cym)Ru(TsEn(Bz,H))Cl] (4, IC50 = 1.0 ± 0.1 μM), having a potency similar to the anticancer drug cisplatin. Co-administration with sodium formate (2 mM), increased the potency of all complexes towards A2780 cells by 20–36%, with the greatest effect seen for complex 6.
The pathways of hydride transfer between pyridinium salts and dihydropyridine compounds are also of interest. The first mechanistic study of the TH reduction of BNA+ (1-benzylnicotinamide), as a model for NAD+ was reported by Steckhan and Fish et al. using [(η5-Cp*)Rh(bipy)Cl] as the catalyst and sodium formate as a hydride source in aqueous media in the 1990s.10–12,16,17 They proposed a catalytic cycle involving a ring-slippage η4-Cp* intermediate with Rh coordinated to the amide of the pyridine ring.18 Knör et al. reported a Rh-coordinated poly(arylene-ethynylene)-alt-poly(arylene-vinylene) polymer as photocatalyst for the reduction of NAD+; involving a possible photoexcited polymer chain being quenched and transferring an electron to the RhIII active centre.19 More recently, Yoon et al. described a mechanism involving hydride transfer to Cp* and formation of the RhI intermediate [(η4-Cp*-H)Rh((CH2OH)2-bipy)]+ followed by hydride transfer from the endo orientation of the C–H bond to maintain the 1,4-regioselectivity of NADH.20
The half-sandwich ruthenium complex [(η6-p-cym)Ru(TsDPEN)Cl] (TsDPEN: N-((1S,2S)-2-amino-1,2-diphenylethyl)-4-methylbenzenesulfonamide) was first reported by Noyori and coworkers in 1995.21,22 Potent catalytic activity has been shown in asymmetric TH reduction of aromatic ketones. Most recently, the 16-electron Os analogues [(η6-arene)Os(TsDPEN)] of Noyori type complexes were reported to reduce pyruvate enantioselectively to (D- or L-) lactate via asymmetric transfer hydrogenation in human cancer cells.23 Nonetheless, the hydrophobic nature of the two phenyl groups on the ethylene backbone limits its application as a possible catalyst for TH reduction of NAD+ under biologically relevant conditions. Complexes with chelating diamine ligands such as complex 7 in Fig. 1, display good aqueous solubility but poor catalytic activity in TH reduction of NAD+.24 However, p-cymene (p-cym) complexes with functional sulfonyl substituents such as [(η6-p-cym)Ru(TsEn)Cl] (e.g. complex 8 in Fig. 1),25 exhibit good solubility in water and improved catalysis for NAD+ reduction to NADH in aqueous media. Moreover, co-administration of [(η6-p-cym)Ru(TsEn)Cl] with low non-cytotoxic doses of sodium formate led to an enhancement of the antiproliferative activity against A2780 human ovarian cancer cells by up to 50×.15,26
Here we investigate the effect on catalytic reduction of NAD+ using formate as a hydride source upon variation of substituents on the amino group of the N,N-chelating TsEn ligand in RuII complexes [(η6-p-cym)Ru(N,N′)Cl] where N,N′ is N-(2-(methylamino)ethyl)-4-toluenesulfonamide (TsEn(Me,H), 1), N-(2-(dimethylamino)ethyl)-4-toluenesulfonamide (TsEn(Me,Me), 2), N-(2-(ethylamino)ethyl)-4-toluene sulfonamide (TsEn(Et,H), 3), N-(2-(benzylamino)ethyl)-4-toluenesulfonamide (TsEn(Bz,H), 4), N-(2-((4-fluorobenzyl)amino) ethyl)-4-toluenesulfonamide (TsEn(4-F-Bz,H), 5) and N-(2-((naphthalen-2-ylmethyl)amino) ethyl)-4-toluenesulfonamide (TsEn(Naph,H), 6) (Table 1). In addition, the catalytic mechanism was investigated both experimentally and by density functional theory (DFT) calculations. We also explored the effect of co-administration of formate on the antiproliferative activity of these complexes against A2780 human ovarian cancer cells.
A2780 human ovarian carcinoma cells were obtained from the European Collection of Cell Cultures. The cell line was grown in Roswell Park Memorial Institute medium (RPMI-1640) supplemented with 10% of foetal calf serum, 1% v/v of 2 mM glutamine and 1% v/v penicillin/streptomycin (10
000 units). All cells were grown as adherent monolayers at 310 K in a 5% CO2-humidified atmosphere and passaged at ca. 70–80% confluency.
Elemental analysis were performed by Warwick Analytical using an Exeter Analytical elemental analyzer (CE440).
Positive ion electrospray mass spectra were obtained on an Agilent 6130B ion mass spectrometer. High resolution mass spectrometry data were obtained on a Bruker Maxis Plus Q-TOF instrument.
X-ray crystallographic diffraction data were collected on an Oxford Diffraction Gemini four-circle system with a Ruby CCD area detector. The structure was refined by full-matrix least-squares against F2 using SHELXL 9735 and solved by direct methods using SHELXS36 (TREF) with additional light atoms found by Fourier methods. The atoms from the sulfonamide nitrogen to the end of the chain (C10 C11 N12 C13) were modelled as disordered over two positions related by a small ruffle in the chain. The occupancy of the two positions was linked to a free variable which refined to 86
:
14. The minor component was refined isotropically. X-ray crystallographic data for complex 3 has been deposited in the Cambridge Crystallographic Data Center (CCDC) under the accession number CCDC 1571331.†
ICP-OES analysis were carried out on a PerkinElmer Optima 5300 DV series Optical Emission Spectrophotometer. The water used for ICP-OES analysis was doubly deionized (DDW) using a Millipore Milli-Q water purification system and a USF Elga UHQ water deionizer. The ruthenium Specupure plasma standard (ruthenium chloride, 1004 ± 5 μg mL−1 in 10% v/v hydrochloric acid) was diluted with 3.6% v/v HNO3 to freshly prepare calibrants at concentrations of 50–700 ppb. Calibration standards were adjusted to match the sample matrix by standard addition of sodium chloride (TraceSELECT®). Total dissolved solids did not exceed 0.2% w/v. Data were acquired and processed using WinLab32 V3.4.1 for Windows.
ICP-MS analysis were carried out on an Agilent Technologies 7500 series ICP-MS instrument. The water used for ICP-MS analysis was double-deionized (DDW) using a Millipore Milli-Q water purification system and a USF Elga UHQ water deionizer. The Ruthenium Specpure plasma standard (ruthenium chloride, 1004 ± 5 μg mL−1 in 10% v/v hydrochloric acid) was diluted with 3.6% v/v HNO3 to prepare calibrants freshly at concentrations of 0.1–1000 ppb. The ICP-MS instrument was set to detect 101Ru in no gas mode. Total dissolved solids did not exceed 0.1% w/v. An internal standard of 166Er (50 ppb) was used. Data were acquired using ICP-MS-TOP and proceeded using Offline Data Analysis (ChemStation version B.03.05, Agilent Technologies, Inc.).
pH values were measured using a Minilab IQ125 pH meter equipped with a ISFET silicon chip pH sensor and referenced in KCl gel. pH* values (pH meter reading without correction for the effect of deuterium on the sensor) of NMR samples in D2O were measured at 310 K. Relative hydrophobicity measurements were performed utilising the Agilent 1200 HPLC system with a VWD and 50 μL loop. The column was an Agilent Zorbax 300SB C18, 150 × 4.6 mm with a 5 μm pore size. The mobile phase was H2O (50 mM NaCl)/H2O/CH3CN 1
:
1 (50 mM NaCl), with a flow of 1 mL min−1. The detection wavelength was set at 254 nm with the reference wavelength at 360 nm.
:
9 v/v, 102 mM sodium formate and 510 μM NAD+ in H2O) was added to a 1 mL cuvette, and the pH adjusted to 7.2, bringing the total volume to 1 mL (final concentrations: Ru complex 28 μM; NAD+ 170 μM; NaHCO2 34 mM; molar ratio 1
:
6
:
1200). UV spectra were recorded and the absorbance at 340 nm was monitored every 5 min until completion of the reaction.
:
4, v/v) (1.4 mM) in a glass vial. Solutions of sodium formate (35 mM) and NAD+ (5.6 mM) in D2O were also prepared and then incubated at 310 K, pH* 7.2 ± 0.1. An aliquot of 200 μL from each solution was added to a 5 mm NMR tube, giving a final volume of 0.64 mL (Ru complex 0.44 mM; NAD+ 1.75 mM; NaHCO2 10.94 mM; molar ratio 1
:
4
:
25). A 1H NMR spectrum was recorded at 310 K every 162 s until the completion of the reaction. Further experiments under similar conditions using different concentrations of sodium formate (complex 4, NAD+ and sodium formate in ratio of 1: 4: X, where X = 10, 25, 50, and 100 mol equiv.) and different concentrations of NAD+ (complex 4, NAD+ and sodium formate in ratio of 1
:
Y
:
25, where Y = 2, 4, 6 and 10) were also studied. Another series of experiments using different pH* values of the reaction solutions (5, 6, 7, 8 and 9) were also performed. Molar ratios of NAD+ and NADH were determined by integrating 1H NMR peaks corresponding to NAD+ (9.33 ppm) and 1,4-NADH (6.96 ppm). The turnover number (TON) for the reaction was calculated as follows:
:
14. Compared to reported ruthenium ethylenediamine complexes (either neutral or +1 charge),25,36,37 the Ru–N− bond length (N9, 2.126(9)) is within the expected range of 2.11–2.14 Å,37 but the Ru–N12 length (2.1702(11) Å) is longer than the neutral analogue [(η6-biph)Ru(TsEn)Cl] (2.122(3) Å),25 suggesting that the presence of N-ethyl substituent causes a slight weakening of this Ru–N bond. The remaining bond length and angles show no significant difference.
![]() | ||
| Fig. 2 ORTEP diagrams for complex 3. Ellipsoids are shown at the 50% probability level. All hydrogen atoms have been omitted for clarity. | ||
| Bonds | Length/angle |
|---|---|
| Ru1–N9 | 2.1256(9) |
| Ru1–N12 | 2.1702(11) |
| Ru1–N12A | 2.157(8) |
| Ru1–Cl1 | 2.4173(3) |
| Ru1–arene (centroid) | 1.664 |
| N9–Ru1–N12 | 78.74(4) |
| N9–Ru1–N12A | 76.1(2) |
| N9–Ru1–Cl1 | 89.47(3) |
| N12–Ru1–Cl1 | 87.55(4) |
determination
:
9 (v/v)). The 1H NMR spectrum remained unchanged after 24 h and the hydrolysis was assumed to be rapid since the peaks could be assigned to the aqua RuII species (4a) by comparison to those from the aqua species generated in a reaction with silver nitrate in D2O (1 mol equiv.). The
(pKa value determined in deuterated solvent) of complex 4a was determined by a pH* (meter reading) titration ranging from 2 to 12 by addition of NaOD or DNO3 solutions as appropriate. Changes in the chemical shift of a tosyl 1H NMR resonance were followed and the data were fitted to the Henderson–Hasselbalch equation, giving a
value of 9.73 ± 0.06 (Fig. 3).
:
9, v/v, Table 3); in all the cases, an increase in intensity of the band at 340 nm was observed, which is assignable to formation of NADH (Fig. S1, ESI†). The kinetics of conversion were also monitored by 1H NMR at 310 K and pH* 7.2 ± 0.1. The reactions were performed in a mixed solvent d4-MeOD/D2O (1
:
4, v/v), due to the poor aqueous solubility of complexes 5 and 6, although the presence of methanol in such reactions is known to enhance the reaction rate.25
In general, the introduction of substituents on the terminal nitrogen improved the catalytic activity. The bulkier the substituents on the terminal nitrogen, the higher the TH reaction rate becomes. The turnover frequency reaches a maximum (ca. 7.54 h−1) when the substituent on the terminal N is benzyl (complex 4), making it as efficient as the RhIII complex [(η5-Cp*)Rh(bipy)Cl]PF6.16 Interestingly, the TOF decreases when the substituent is para-fluoro-benzyl (complex 5) or naphthalene (complex 6), probably, because these ligands hamper the approach of NAD+ to the Ru centre. Compared to the en complex with unsubstituted nitrogens [(η6-biph)Ru(en)Cl]PF6, the turnover frequency of complex 4 is 41× higher,24 and 2.7× higher compared to [(η6-p-cym)Ru(TsEn)Cl].25
The NH proton of the chelated diamine ligand appears to be essential for the TH reduction of ketones to alcohols;39 normally, RuII catalysts for TH of ketones form 16-e intermediates.40,41 It has been reported that a RuII complex with two N-alkyl groups (R,R)-[(η6-benzene)Ru(TsDPEN-Me2)Cl] exhibited poor catalytic reactivity in TH reaction of ketones.40 However, complex 2 [(η6-p-cym)Ru(TsEn(Me,Me))Cl] exhibited good catalytic activity towards the TH reduction of NAD+ to NADH (TOF = 4.1 h−1, Table 3), despite not having an NH proton, which suggests, as expected, that an N–H is not essential in the transfer reduction of NAD+ to NADH.
The dependence of the rate of catalysis on pH was determined. Six pH* values ranging from 5 to 9 were studied for complex 4 at a mol ratio complex 4
:
NAD+
:
formate of 1
:
4
:
25, in the same mixed solvent at 310 K (Fig. S2, ESI†). The TOF was relatively insensitive to pH* over the range pH* 6–8 (ca. 7.5 h−1), but decreased slightly at lower and higher pH* (5.6 h−1 at pH* 5, 6.6 h−1 at pH* 9).
The dependence of turnover frequency on the concentrations of sodium formate and NAD+ was also investigated for complex 4 in d4-MeOD/D2O (1
:
4) at 310 K. The mol ratio of complex 4
:
NAD+
:
formate was 1
:
4
:
X, where X = 5, 10, 25, 50 and 100 (Fig. S3, ESI†). The TOF increased steadily from 2.2 h−1 to 18.8 h−1 as the concentration of formate was increased from 7 mM to 140 mM. Next the dependence of TOF on the NAD+ concentration was studied for mol ratio complex 4
:
NAD+
:
formate = 1
:
Y
:
25, where Y = 2, 6 and 10. The TOF was found to be independent of NAD+ concentration (7.7 ± 0.5 h−1).
The Michaelis–Menten kinetic behaviour is apparent from a plot of turnover frequency versus formate concentration. A reciprocal plot of turnover frequency versus formate concentration gave a Michaelis constant of KM = 0.086 mM (Fig. S3 and S4, ESI†). The maximum turnover frequency TOFmax for complex 4 (30.3 h−1) is ca. 5× higher than for [(η6-p-cym)Ru(TsEn)Cl] (complex 8, TOFmax = 6.4 h−1)25 and 20× higher than for the complex [(η6-hmb)Ru(en)Cl]PF6 (TOFmax = 1.46 h−1).24 The much lower Michaelis–Menten constant (KM = 0.086 mM) for the N-benzyl complex 4 indicates a stronger affinity of the complex for formate compared to [(η6-p-cym)Ru(TsEn)Cl] (KM = 27.8 mM)25 and [(η6-hmb)Ru(en)Cl]PF6 (KM = 58 mM).24
The maximum turnover frequency was observed at pH* 6 (TOFmax = 7.7 h−1) (Fig. S2, ESI†). The TOF for complex 4 gradually decreased when the pH* was raised above 6. Transfer hydrogenation was halted below pH* 4 because of the decomposition of the complex.
![]() | ||
| Fig. 4 Antiproliferative activity of RuII complexes 1–6 and cisplatin towards A2780 human ovarian cancer cells. | ||
Combination treatment with formate can greatly increase the antiproliferative activity of RuII arene sulfonyl diamine complexes, which offers a potential new strategy for cancer treatment.15 In this work, the antiproliferative activity of RuII complexes in A2780 human ovarian cancer cells in the presence of sodium formate was studied (Fig. 5). Firstly, the cytotoxicity of sodium formate alone towards A2780 human ovarian cancer cells was investigated. No significant toxicity was found up to formate concentrations of 2 mM which is in agreement with the previous report.15 Then, A2780 human ovarian cancer cells were coincubated with equipotent concentrations of complexes 1–6 (1/3 × IC50) and three different concentrations of sodium formate (0.5, 1 and 2 mM) in order to observe the formate-concentration dependence of the cell viability. The antiproliferative activity of complexes 1–6 increased significantly upon coincubation with 2 mM formate. The formate-induced decrease in viability of A2780 cells ranged from 20% to 36% in the presence of complexes 1–6. Interestingly for complex 6, a 28% decrease in cell viability was observed with only 0.5 mM formate present (Fig. 5, for percentage of viability decrease see Table S3, ESI†). The largest decrease of cell survival was 31% for complex 6 in the presence of 2 mM sodium formate, followed by 29% and 32% for the other two complexes with aromatic substituents, complexes 4 and 5, respectively. Complexes 1–3 with aliphatic functional groups showed an increase in potency of 18%, 21% and 22%, respectively.
Complex 4 gave the lowest cellular accumulation (0.52 ± 0.08 ng of Ru per 106 cells), while complex 6 with moderate anticancer activity, exhibited the highest extent of cell accumulation with 4.5 ± 0.2 ng of Ru per 106 cells at IC50 concentration, 8.6× higher than complex 4. Complexes 1–3 and 5, gave rise to similar cell uptake 2.4 ± 0.3 ng, 1.2 ± 0.2 ng, 3.0 ± 0.2 ng and 1.3 ± 0.2 ng per 106 cells, respectively, following the order: 4 < 2, 5 < 1 < 3 < 6.
The relative hydrophobicity of complexes 1–6 was determined by RP-HPLC. The more hydrophobic complexes have longer retention times on a reverse-phase C18 column.44 To ensure solubility of the RuII complexes in water, methanol was used as co-solvent (MeOH/H2O, 1
:
9 v/v) together with NaCl (50 mM) to suppress hydrolysis of the complexes. The HPLC solvents were also prepared with 50 mM NaCl (measurements see Fig. S5, ESI†). The resulting retention times are shown in Table 4, and follow the order: 1, 2, 3 < 4, 5 < 6. Complex 3 shows the shortest retention time (least hydrophobic) of 14.0 min, while complex 6 shows the longest retention time (most hydrophobic), 20.9 min.
| Complex | t R (min) | Cellular-Ru (ng per 106 cells) |
|---|---|---|
| 1 | 15.4 ± 0.9 | 2.4 ± 0.3 |
| 2 | 14.5 ± 0.3 | 1.2 ± 0.2 |
| 3 | 14.0 ± 0.3 | 3.0 ± 0.2 |
| 4 | 17.4 ± 0.2 | 0.52 ± 0.08 |
| 5 | 17.27 ± 0.08 | 1.3 ± 0.2 |
| 6 | 20 ± 1 | 4.5 ± 0.2 |
It is evident from Table 4 that the RuII complexes with aromatic substituents (complexes 4–6) exhibit higher hydrophobicity than complexes with aliphatic substituents (complexes 1–3). The most hydrophobic complex (6) shows the highest cell accumulation. Nonetheless, there is no linear correlation between the hydrophobicity of complexes 1–6 and their cellular accumulation. This has been observed before.45 In these cases, the chemistry and the mechanism of action of each particular complex has a higher impact on the compound's anticancer activity than cellular accumulation per se. However, complex 4 has the lowest extent of cell uptake, but the most potent antiproliferative activity, suggesting that it is the chemical properties of the intracellular drug that are more important for activity than the total amount of Ru entering the cell. In general, a high hydrophobicity could facilitate interaction between the organometallic complex and cell membranes, and also correlate with the potency of the complex, but that is not always the case.43,45
:
1 mol ratio. The formation of adduct 4-9-EtG was confirmed by following the new set of peaks, and up to 90% yield of adduct was obtained when 1.5 mol equiv. nucleobase solution was added. However, no adduct was found when 1.5 mol equiv. of 5′-AMP was added to complex 4, even after 24 h incubation at 310 K. Reactions of double-helical calf thymus DNA (ct-DNA, 32 μg mL−1) and plasmid DNA pBR322 (28 μg mL−1) with complex 4 in various molar ratios (ri = 0.05–1, ri = the molar ratio of free Ru complex to nucleotide phosphates at the onset of incubation with DNA) were studied. Very low amounts of ruthenium (5–7% of initial Ru) were found in the samples of DNA treated with complex 4 for 24 h. No significant changes in the mobilities of supercoiled (sc) or open circular (oc) form of plasmid DNA were observed even when incubated with high concentration of complex 4 (ri = 1, Fig. S7, ESI†). DNA is thought to be a cellular target for the en complex 7 (Fig. 1).52,53 However, for the substituted-en complex studied here, no obvious unwinding of DNA was observed after coincubation of ct-DNA with complex 4, suggesting that binding is weak, nor changes in the ratio of sc and oc forms of plasmid DNA, suggesting that complex 4 does not cleave DNA.
![]() | ||
| Fig. 8 (Top) Reduction cycle for conversion of NAD+ to 1,4-NADH via transfer hydrogenation with formate as the hydride donor. (Bottom) DFT energy profile for the formation of Ru formate species, Ru hydride complex and hydride transfer from ruthenium; brown line, complex 1; red line, complex 2; blue line, complex 3; green line, complex 4; purple line, complex 5; black line, complex 6. Sets of calculated structures of states 1–7 are supplied in the ESI and illustrated graphically in Fig. S8† for complex 2. To calculate the energy of states 1 and 7, the energies of the states represented in pdb files 1 and 7 were added to the energies calculated for NAD+ and NADH. | ||
For state 5, with ring-slipped coordinated η2-p-cymene, the introduction of water into coordination sphere was necessary, while highly distorted complexes without coordinated water were found ca. 100 kJ mol−1 higher in energy. It is notable that the Ru atoms of all complexes in state 5 are coordinated to the amide oxygen atom of NADH, while only weakly bound to the (hydridic) CH2 of NADH, giving a Ru–H distance of 3.11–3.12 Å for R1,R2 = Me,H (1); Et,H (3); Naph,H (6) and 3.06–3.07 Å for R1,R2 = Bz,H (4) and 4-F-Bz,H (5). For R1,R2 = Me,Me (2) the calculations revealed a true bonding of the (hydridic) CH2, with a Ru–H distance of 1.99 Å. The results obtained are shown in Fig. 8. Four general conclusions can be drawn from these data: (a) there is a strong interaction between the [(η6-p-cym)Ru(TsEn(R1,R2))(OH2)]+ cation and NAD+ and NADH molecules, leading to a stabilisation of the cationic form by 60–70 kJ mol−1 for NAD+ (41 kJ mol−1 for R1,R2 = 4-F-Bz,H (5)) and 130–150 kJ mol−1 for NADH; (b) depending on the N-substituent, the species of the lowest energy is either [(η6-p-cym)Ru(TsEn(R1,R2))(HCOO−)]·NAD+ (R1,R2 = Me,H (1); Bz,H (4); 4-F-Bz,H (5) and Naph,H (6)) or [(η6-p-cym)Ru(TsEn(R1,R2))(OH2)]·NAD+ (R1,R2 = Et,H (3) and Me,Me (2)), the difference between them being only 6–12 kJ mol−1; (c) the effective NADH-hydride coordination for bulky R1,R2 = Me,Me (2) lowers the energy, relative to the state of lowest energy, of the species with coordinating NADH by 30–40 kJ mol−1, compared to other complexes; (d) the formation of the state with the Ru–H hydride bond, including the twist of formate and the elimination of carbon dioxide, corresponds to the highest energy step. These four factors seem all to influence the turnover. The energy barriers and optimized structures for the seven states of complex 2 in the cycle with NAD+ are listed in Table S5 and illustrated graphically in Fig. S8.† The structure files for the remaining complexes are supplied as ESI.†
To investigate the possibility of achieving transfer hydrogenation mediated by formate in cells, we investigated the effect of formate on the antiproliferative activity of these complexes towards human ovarian cancer cells. In each case a dose-dependent increase in potency of the complexes (20–36%) was observed with increasing formate concentration over a range of non-toxic formate concentrations (0–2 mM). The complexes with aromatic substituents were the most potent, the benzyl complex 4 being as potent as the anticancer drug cisplatin (Fig. 6). In general, the most hydrophobic complexes were found to be the most biologically active. However, the activity does not correlate closely with total cell accumulation of Ru or with hydrophobicity (Table 3). Although DNA can be a target for related arene RuII diamine complexes, it does not appear to be a target for these sulfonyl-en RuII catalysts since we observe very weak binding to both calf thymus and plasmid DNA (Fig. S7, ESI†).
We showed that complexes 1 and 4 can generate high levels of ROS in A2780 human ovarian cancer cells, especially 4, the most potent complex. This is consistent with interference in cellular redox pathways and possible attack on NAD+ when sodium formate is present. The enhancement of anticancer activity by low non-toxic dose of formate might be clinically useful since it introduces a new mechanism of activity which does not involve DNA attack, unlike the clinical drug cisplatin. Such a regime might therefore avoid some unwanted side-effects. Formate itself is a natural biochemical molecule enriched in some cancer cells.54 However, more work remains to be done to investigate possible intracellular catalysis, especially since a range of metabolites might readily poison these catalysts in cells.
Feng Chen carried out synthesis and characterisation of ligands and complexes, investigated hydrolysis, determined the pKa value, and TH turnover frequencies.
Isolda Romero-Canelón and Ji-Inn Song carried out the cell antiproliferative screening and related biochemical assays.
Guy J. Clarkson carried out the X-ray crystallography.
Juliusz A. Wolny and Volker Schünemann carried out all the DFT calculations.
Jana Kasparkova and Viktor Brabec carried out DNA binding studies.
James P. C. Coverdale carried out metal analyses by ICP-OES and ICP-MS, and related biological and biochemical assays.
All authors contributed to the writing of the paper.
We also thank Dr Ivan Prokes, Dr Lijiang Song, and Mr Philip Aston (University of Warwick) for their excellent assistance with the NMR and MS measurements.
Footnote |
| † Electronic supplementary information (ESI) available. CCDC 1571331. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c8dt00438b |
| This journal is © The Royal Society of Chemistry 2018 |