Next Article in Journal
Therapeutic Targeting the Cell Division Cycle 25 (CDC25) Phosphatases in Human Acute Myeloid Leukemia — The Possibility to Target Several Kinases through Inhibition of the Various CDC25 Isoforms
Next Article in Special Issue
Synthetic Applications of Intramolecular Thiol-Ene “Click” Reactions
Previous Article in Journal
Data-Driven Techniques for Detecting Dynamical State Changes in Noisily Measured 3D Single-Molecule Trajectories
Previous Article in Special Issue
The Dynamical Behavior of the s-Trioxane Radical Cation—A Low-Temperature EPR and Theoretical Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermal Reactivity of Neutral and Oxidized Ferrocenyl-Substituted Enediynes

by
Mehmet Emin Cinar
,
Guido Morbach
and
Michael Schmittel
*
Department Chemie-Biologie, Universität Siegen, Adolf-Reichwein-Str., Siegen D-57068, Germany
*
Author to whom correspondence should be addressed.
Molecules 2014, 19(11), 18399-18413; https://doi.org/10.3390/molecules191118399
Submission received: 18 September 2014 / Revised: 1 November 2014 / Accepted: 3 November 2014 / Published: 12 November 2014
(This article belongs to the Special Issue Free Radicals and Radical Ions)

Abstract

:
The coupling of two equivalents of ethynylferrocene (2) with one equivalent of 1,2-diiodocyclohexene (1) and 1,2-diiodobenzene (4) using Sonogashira cross-coupling conditions led to 1,2-bis(ferrocenylethynyl)cyclohexene (3) and 1,2-bis(ferrocenylethy-nyl)benzene (5), respectively. At high temperatures enediynes 3 and 5 showed exothermic signals in differential scanning calorimetry (DSC) measurements, suggestive of intramolecular diradicaloid ring formation (Bergman (C1−C6) or Schreiner-Pascal (C1−C5) cyclizations). The oxidation of 3 and 5 to the mono-oxidized enediynes 3+ and 5+ decreased the onset temperatures drastically. Equally, 1-ferrocenylethynyl-2-(p-nitro-phenyl)ethynylbenzene (8) displayed a significant decrease in the onset temperature after oxidation to 8+. Because the insoluble nature of the polymeric material formed in the thermolysis of the oxidized enediynes prevented characterization, the origin of this drastic effect was studied by DFT. Contrary to expectations, one-electron oxidation does not lower the barrier for intramolecular cyclization. Rather, the computations suggest that the polymerization is initiated by a bimolecular process.

Graphical Abstract

1. Introduction

The discovery of the antibacterial and antitumor properties of the naturally occurring enediynes [1], such as calicheamicin, dynemicin and kedarcidin, has attracted the interest of many scientists over the last three decades [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16]. Because of their ability to afford (under physiological conditions!) reactive benzenoid σ,σ-diradicals [17,18,19,20] that induce single and double strand DNA cleavage, a significant amount of interest has focused on the design of simple acyclic enediynes with similar properties. Notably, chelate complexation in acyclic enediynes bringing closer together the C1 and the C6 termini [21,22,23,24,25,26,27] lowered the cyclization temperature as did the use of smaller strained cyclic enediynes [28,29]. Moreover, ligand field effects were exploited to influence the reactivity of acyclic metalloenediynes [30]. While it has been widely documented that the temperature of the Bergman cyclization depends on the C1–C6 distance, the effect of varying electronic properties has not been fully understood yet [31,32,33,34,35,36,37]. Since the activation of neutral enediynes has been accomplished by photoinduced electron transfer (PET) [38], we evaluate herein the possibility to activate ferrocenyl enediynes by thermal one-electron oxidation, a methodology that has been successful in our hands over many years [39,40].

2. Results and Discussion

2.1. Synthesis of Enediynes 3, 5 and 8, and Their Oxidation to 3+, 5+ and 8+

The synthesis of enediynes 3 and 5 was accomplished by reacting two equivalents of ethynylferrocene (2) [41] in Sonogashira cross-couplings [42,43,44,45] with one equivalent of 1,2-diiodocyclohexene (1) or 1,2-diiodobenzene (4) [46], furnishing 3 and 5 in 61% and 58% yield, respectively. Two sequential Sonogashira couplings of 1,2-diiodobenzene (4) [47], first with ethynylferrocene (2) and then p-nitrophenylacetylene (7) [48,49] provided 8 (Scheme 1). The structures of 3 and 5 were unambiguously established by X-ray crystallography (Figure 1). Their oxidation using tris(4-bromophenyl)aminium hexachloroantimonate or tris(4-tolyl)aminium hexafluoroantimonate as one-electron oxidant furnished 3+, 5+ and 8+ in high yield.
Figure 1. Crystal structures of 3 (left) and 8 (right). Critical distances are given in pm.
Figure 1. Crystal structures of 3 (left) and 8 (right). Critical distances are given in pm.
Molecules 19 18399 g001
Scheme 1. Synthesis of enediynes 3, 5 and 8, and their oxidation to 3+, 5+ and 8+.
Scheme 1. Synthesis of enediynes 3, 5 and 8, and their oxidation to 3+, 5+ and 8+.
Molecules 19 18399 g008
Electrospray ionization mass spectra of 3, 5 (Figure 2) and 8 (Figure S7) showed in all cases the signal of the mono-oxidized enediynes 3+ (498.4), 5+ (494.2) and 8+ (431.4). The spectrum of 3+ displayed additionally a very weak signal, which belongs to the doubly oxidized enediyne 32+ (249.2).
Figure 2. Electrospray ionization mass spectra of 3 (left) (insets displaying 3+ (a) and 32+ (b)) and 5 (right) along with their isotopic distributions (red: calculated, black: experimental distribution).
Figure 2. Electrospray ionization mass spectra of 3 (left) (insets displaying 3+ (a) and 32+ (b)) and 5 (right) along with their isotopic distributions (red: calculated, black: experimental distribution).
Molecules 19 18399 g002

2.2. Electrochemical Properties of 3, 5 and 8

Electrochemical investigations of compounds 3, 5 and 8 in CH3CN were carried out using cyclic voltammetry at scan rates of 20, 50, 100, 200 and 500 mV·s−1. While the symmetrical enediynes exhibited reversible oxidation waves at E1/2ox (3) = +0.90 VTPP and E1/2ox (5) = +0.89 VTPP, the push–pull system 8 displayed reversible oxidation and reduction waves at E1/2ox (8) = +0.90 VTPP and E1/2red (8) = −0.63 VTPP, respectively. The amount of charge transfer in the oxidation was gauged against 1,1'-dimethylferrocene (1 equiv.) showing a ratio of current ratio of 0.98 for 3, 0.96 for 5 and 1.0 for 8 (Figure 3). Because compound 8 contains only one ferrocenyl unit, but shows the same current ratio against the internal standard as 3 and 5, the results clearly support a one-electron oxidation of 3, 5 and 8 yielding the monocations 3+, 5+ and 8+. This finding is supported by the computational results demonstrating equal spin density at both ferrocene units in the cations, but contrasts other interpretations in the literature claiming no electronic communication between the ferrocene groups [50]. The latter studies, however, were done in CH2Cl2 as a strongly ion-pairing solvent.
Figure 3. Cyclic voltammograms of (a) 3; (b) 5 and (c) 8 in CH3CN at 100 mV·s−1. In each case, the cyclic voltammogram shows the wave of 1,1'-dimethylferrocene (lower potential) in presence of the enediyne (higher potential). (Potentials are given with respect to TPP = 2,4,6-triphenylpyrylium tetrafluoroborate).
Figure 3. Cyclic voltammograms of (a) 3; (b) 5 and (c) 8 in CH3CN at 100 mV·s−1. In each case, the cyclic voltammogram shows the wave of 1,1'-dimethylferrocene (lower potential) in presence of the enediyne (higher potential). (Potentials are given with respect to TPP = 2,4,6-triphenylpyrylium tetrafluoroborate).
Molecules 19 18399 g003

2.3. Thermal Properties

To evaluate the thermal stability of both the neutral and mono-oxidized enediynes we used differential scanning calorimetry (DSC) at a heating rate of 10 °C·min−1 (Figure 4). For the neutral enediynes 3, 5 and 8 the Tonset of the exothermic signals were determined to be 262, 300 and 216 °C. These onset temperatures are in good agreement with those of structurally related diaryl enediynes undergoing Bergman cyclization [35]. Equally, the trend is the same. With increasing electron density pushed into the enediyne by the substituents, the temperature of cyclization is raised. In contrast, the Tonset temperatures of the oxidized enediynes 3+, 5+ and 8+ were recorded at 77, 128 and 76 °C, respectively. Unfortunately, the analysis of the thermolyzed samples showed insoluble polymeric material that excluded any routine organic characterization.
Figure 4. DSC graphs of 3, 5 and 8 and their monocations 3+, 5+ and 8+ (heating rate: 10 °C·min−1).
Figure 4. DSC graphs of 3, 5 and 8 and their monocations 3+, 5+ and 8+ (heating rate: 10 °C·min−1).
Molecules 19 18399 g004

2.4. Computations

To shed light on the different thermal stability of the neutral and mono-oxidized enediynes, we conducted DFT calculations on both 3 and 3+. By comparing the most common DFT methods in regard to accuracy and computational costs for the Bergman and Myers-Saito reactions, Schreiner et al. [51] established the pure DFT functionals (MPWLYP, G96LYP, and BLYP) as the most suitable methods. Therefore, the BLYP method with a 6-31G(d,p) basis set combined with LANL2DZ was chosen for computing the stationary points in our study. Unless otherwise stated, full geometry optimization of all gas-phase stationary points was performed with 6-31G(d,p) and LANL2DZ basis sets using Becke’s pure gradient-corrected exchange functional [52] in conjunction with the Lee-Yang-Parr non-local correlation functional (BLYP) [53] as implemented in Gaussian 09 [54]. For all open-shell singlet state transition state (TS) structures, intermediates and cation radicals we used unrestricted calculations with a broken spin symmetry approach (BS-UBLYP) involving the mixing of the frontier molecular orbitals (HOMO and LUMO) to break the spin and spatial symmetries. The restricted method was applied for all closed-shell molecules. It is well known that the broken spin approach in combination with DFT provides good results for the cyclization of unsaturated systems [2,51,55,56,57]. The minima and transition states were verified by analyzing the harmonic vibrational frequencies, using analytical second derivatives, which have NIMAG = 0 and 1, respectively.
Based on the computational results, mono-oxidation of the enediyne 3 did not decrease the activation barrier for the Bergman or Pascal-Schreiner [51,58] cyclizations (Scheme 2). This phenomenon can be rationalized by the fact that a decrease of electron density occurs in the out-of-plane π-orbitals rather than diminishing electron repulsion between the in-plane π-orbitals (Figure 5). Interestingly, when we rotated the ferrocenyl group in 8+ to probe for an in-plane distonic cation radical, the spin density nevertheless remained in the out-of-plane orbitals. A minimum was located at 68° that was only 2.6 kcal·mol−1 higher in energy (Figure S9).
Scheme 2. (U)BLYP/6-31G(d,p)/LANL2DZ computational results on the Bergman (B) and Schreiner-Pascal (SP) cyclizations of enediynes 3 and 3+ (electronic energies related to the starting enediynes are given in kcal·mol−1).
Scheme 2. (U)BLYP/6-31G(d,p)/LANL2DZ computational results on the Bergman (B) and Schreiner-Pascal (SP) cyclizations of enediynes 3 and 3+ (electronic energies related to the starting enediynes are given in kcal·mol−1).
Molecules 19 18399 g009
Figure 5. SOMO (a); LUMO (b) and spin density (c) of 3+.
Figure 5. SOMO (a); LUMO (b) and spin density (c) of 3+.
Molecules 19 18399 g005
To investigate the reason for the drastic change in the thermal reactivity of the monocations, we had a closer look on the parent enediyne 9. The cation radical 9+• exhibits an activation barrier of 28.9 kcal·mol−1 for the Bergman−analogous cyclization, which is higher than the TS of the neutral enediyne cyclization by 3.7 kcal·mol−1. In contrast, for the Schreiner-Pascal reaction, the cation radical exhibits a lower barrier (by 4.6 kcal·mol−1) than the neutral enediyne (Scheme 3). However, this rather small decrease cannot explain the low temperatures for cyclization as found experimentally. Hence, a plausible explanation for the experimental finding is to postulate an intermolecular dimerization initiating the polymerization. Notably, the DSC graphs display sharp exothermic signals in line with a chain reaction. Indeed, bringing neutral 9 and cation radical 9+• molecules close together results in the corresponding product formation almost without any activation barrier (Figure 6). Alike two cation radicals 9+• showed an almost barrierless (<0.7 kcal·mol−1) bond formation (Figure 7).
Scheme 3. (U)BLYP/6-31+G(d) calculations on the Bergman (B) and Schreiner-Pascal (SP) cyclizations of cation radical 9+• (Calculations on 9 at (UBS)-BLYP/6-31G(d) level from reference [56]. Relative energies with respect to enediynes are given in kcal·mol−1).
Scheme 3. (U)BLYP/6-31+G(d) calculations on the Bergman (B) and Schreiner-Pascal (SP) cyclizations of cation radical 9+• (Calculations on 9 at (UBS)-BLYP/6-31G(d) level from reference [56]. Relative energies with respect to enediynes are given in kcal·mol−1).
Molecules 19 18399 g010
Figure 6. Analysis of the bond formation between C7 and C19 by performing the relaxed scan method at UBLYP/6-31+G(d) level on the reaction of 9+• (left) and the neutral 9 (right) molecule (Relative energies are in kcal·mol−1 and distances are given in Å).
Figure 6. Analysis of the bond formation between C7 and C19 by performing the relaxed scan method at UBLYP/6-31+G(d) level on the reaction of 9+• (left) and the neutral 9 (right) molecule (Relative energies are in kcal·mol−1 and distances are given in Å).
Molecules 19 18399 g006
Figure 7. Analysis of the bond formation between C7 and C19 by performing the relaxed scan method at (UBS)-BLYP/6-31+G(d) level on the intermolecular reaction between two cation radicals 9+• (Relative energies are in kcal·mol−1 and distances are given in Å).
Figure 7. Analysis of the bond formation between C7 and C19 by performing the relaxed scan method at (UBS)-BLYP/6-31+G(d) level on the intermolecular reaction between two cation radicals 9+• (Relative energies are in kcal·mol−1 and distances are given in Å).
Molecules 19 18399 g007

3. Experimental Section

3.1. General Information

Melting points were determined with a Büchi melting point apparatus and are uncorrected. 1H-NMR (200 MHz) and 13C-NMR spectra were recorded on a Bruker AC 200 instrument and chemical shifts are reported as δ values using the solvent CDCl3H = 7.25/δC = 77.00) as internal reference. IR spectra were recorded on a Perkin Elmer 1750 FTIR. Column chromatography was carried out over silica gel (Merck, Darmstadt, Germany, SiO2, 0.063–0.200 mm, 70–230 mesh). TLC was performed using pre-coated silica gel SIL G/UV254 plates (Macherey-Nagel GmbH & Co. KG, Düren, Germany). ESI investigations were conducted on the LCQ Deca from Thermo Quest. Solvents for extraction and chromatography were technical grade and distilled prior to use. NEt3 was distilled over CaH2. All Sonogashira cross coupling reactions were carried out under dry nitrogen or argon in oven dried glassware. Cyclic voltammetry (CV) measurements were performed with a potentiostat (model 362, Fa. Princeton Applied Research) using a standard three electrode set-up (Pt working and Pt auxiliary electrode, silver wire as reference electrode). The experiments were carried out using a 1 mM solution of the enediyne in acetonitrile with 0.1 M tetra-n-butylammonium hexafluorophosphate as supporting electrolyte. All potentials were measured at a scan rate of 100 mV·s−1 and are referenced initially to triphenylpyrylium tetrafluoroborate (TPP, E1/2red = −0.38 VSCE, E1/2red = −0.75 VFc) as internal standard. Ethynylferrocene (2) was synthesized according to the literature [41].

3.2. 1,2-Bis(ferrocenylethynyl)cyclohexene (3)

Under nitrogen, 1,2-diiodocyclohexene (1) [46] (397 mg, 1.19 mmol), bis(triphenylphosphine)palla-dium(II) chloride (167 mg, 2.38 µmol) and copper(I)iodide (90.7 mg, 476 µmol) were suspended in triethylamine (40 mL). Then, ethynylferrocene (2) [41] (500 mg, 2.38 mmol) was added and the mixture was stirred for 5 h at 55–60 °C. The solvent was evaporated under reduced pressure and the remaining residue was dissolved in dichloromethane (30 mL). The organic layer was washed three times with a 2.5% potassium cyanide solution (25 mL) and three times with water (25 mL). The combined organic layers were dried over anhydrous magnesium sulfate and concentrated under atmospheric pressure. The crude product was purified by column chromatography over silica gel with n-hexane/dichloromethane (4:1, Rf = 0.18) as eluent furnishing an orange solid in 61% yield (359 mg, 721 µmol) with a melting point of 131 °C (DSC). IR (KBr): ν = 3097 (m), 2930 (s), 2858 (m), 2830 (m), 2134 (s, C≡C), 1636 (m), 1474 (m), 1433 (m), 1410 (m), 1351 (m), 1261 (w), 1202 (w), 1176 (w), 1135 (w), 1106 (s), 1024 (s), 1001 (s), 931 (w), 915 (m), 817 (s), 476 (s) cm−1. 1H-NMR (CDCl3): δ = 1.68 (m, 4H), 2.33 (m, 4H), 4.22 (t, J = 2.0 Hz, 4H), 4.23 (s, 10H), 4.49 (t, J = 2.0 Hz, 4H). 13C-NMR (CDCl3): δ = 22.0, 30.4, 65.8, 68.7, 70.1, 71.4, 87.0, 92.0, 125.6. C30H26Fe2 (498.2): Calcd. C 72.32, H 5.26; found C 71.81, H 5.27.

3.3. Oxidation of 1,2-Bis(ferrocenylethynyl)cyclohexene 3 to the Monocation 3+

1,2-Bis(ferrocenylethynyl)cyclohexene (3, 30.0 mg, 60.2 µmol) was dissolved in dry chloroform (3 mL) under nitrogen. Thereafter, tris(4-bromophenyl)aminium hexachloroantimonate (99.1 mg, 121 μmol) was added until the reaction mixture remained blue. The mixture was stirred for another 30 min resulting in the dark green solution. After the addition of dry n-hexane (5 mL), a precipitate formed that was filtered under nitrogen and washed with dry n-hexane (50 mL). Drying under vacuum afforded 50.0 mg (60.0 µmol, quantitative yield) of an olive green solid (C30H26Fe2SbCl6). IR (KBr): ν = 3106 (m, CFc-H), 2931 (m), 2859 (w), 2201 (m, C≡C), 1467 (s), 1418 (m, C=C), 1397 (w), 1352 (w), 1263 (w), 1181 (w), 1090 (w), 1030 (w), 1008 (w), 917 (w), 854 (s, ferrocene), 651 (w) cm−1. ESI (C30H26Fe2+): Calcd. 498.2; found 498.1.

3.4. 1,2-Bis(ferrocenylethynyl)benzene (5)

Under nitrogen, 1,2-diiodobenzene (4, 148 mg, 449 µmol), bis(triphenylphosphine)palladium(II) chloride (15.8 mg, 22.5 µmol) and copper(I)iodide (9.00 mg, 47.3 µmol) were suspended in triethylamine (5 mL) at ambient temperature. Then, ethynylferrocene (2) [41] (200 mg, 952 μmol) in anhydrous triethylamine (8 mL) was added to the yellow suspension. The reaction mixture was stirred at 80 °C for 3 h. After removal of the solvent, the remaining residue was dissolved in dichloromethane. The organic layer was washed three times with a 2.5% potassium cyanide solution (30 mL) and three times with water (30 mL). The combined organic layers were dried over anhydrous magnesium sulfate and concentrated under atmospheric pressure. The crude product was purified by column chromatography over silica gel with n-hexane/dichloromethane (4:1, Rf = 0.1) as eluent furnishing an orange solid in 58% yield (128 mg, 259 µmol). M.p.: 166 °C (DSC). IR (KBr): ν = 3037 (m, arom. =C-H), 2202 (m, C≡C), 1555 (m, C=C), 1478 (m, C=C), 1431 (s, ferrocene), 1002 (s, ferrocene), 814 (m, ferrocene) cm−1. 1H-NMR (CDCl3): δ = 4.25 (s, 10H), 4.27 (t, J = 1.8 Hz, 4H), 4.57 (t, J = 1.8 Hz, 4H), 7.25 (dd, J = 5.9, 3.4 Hz, 2H), 7.50 (dd, J = 5.9, 3.4 Hz, 2H). 13C-NMR (CDCl3): δ = 64.3, 67.9, 69.1, 70.5, 83.9, 91.3, 125.0, 126.3, 130.7. C30H22Fe2 (494.2): Calcd. C 72.91, H 4.49; found C 72.72, H 4.43.

3.5. Oxidation of 1,2-Bis(ferrocenylethynyl)benzene (5) to the Monocation 5+

1,2-Bis(ferrocenylethynyl)benzene (5, 64.3 mg, 130 µmol) was dissolved in anhydrous chloroform (5 mL) under nitrogen atmosphere. Thereafter, tris(4-tolyl)aminiumhexafluoroantimonate (136 mg, 260 mol) was added until the reaction mixture remained blue. The mixture was stirred for another 30 min, while the solution turned to dark green. After the addition of anhydrous n-hexane (5 mL), a precipitate formed, which was filtered under nitrogen and washed with anhydrous n-hexane (50 mL). Drying under vacuum yielded 90.0 mg (123 µmol, 95%) of an olive green solid (C30H22Fe2SbF6). IR (KBr): ν = 3105 (s), 2210 (m, C≡C), 1601 (s), 1561 (m), 1476 (s), 1418 (m), 1322 (m), 1191 (m), 1070 (s), 1007 (s), 986 (m), 853 (s), 824 (m), 756 (s) cm−1. ESI (C30H22Fe2+): Calcd. 494.2; found 494.2.

3.6. 2-Ferrocenylethynyl-1-iodobenzene (6)

A mixture of ethynylferrocene (2) [41] (200 mg, 952 μmol), bis(triphenylphosphine)palladium(II) chloride (67.0 mg, 95.7 μmol), copper(I)iodide (36.3 mg, 191 μmol), and 1,2-diiodobenzene (4, 330 mg, 1.00 mmol, 131 μL) in anhydrous triethylamine (10 mL) was prepared under nitrogen and stirred at 50 °C for 5 h. After removal of the solvent under reduced pressure, 100 mL of dichloromethane was added. The organic phase was extracted with 2.5% aqueous KCN solution (3 × 70 mL) and washed with water (3 × 70 mL). The combined organic layers were dried over anhydrous magnesium sulfate, thereafter the solvent was evaporated under atmospheric pressure. The crude product was purified by column chromatography over silica gel with n-hexane/dichloromethane (4:1, Rf = 0.30) as eluent resulting in the title compound in 32% yield (127 mg, 308 μmol). IR (film): ν = 3104 (m, C-H), 2925 (m, C-H), 2223 (s, C≡C), 1628 (m), 1551 (w), 1483 (s), 1443 (s), 1429 (s), 1412 (s), 1293 (w), 1200 (w), 1171 (w), 1104 (s), 1016 (s), 1000 (s), 922 (s), 816 (s), 759 (s), 710 (w), 649 (w), 549 (m), 520 (s), 498 (s), 485 cm−1 (s). 1H-NMR (CDCl3): δ = 4.28 (t, J = 1.8 Hz, 2H), 4.32 (s, 5H), 4.60 (t, J = 1.8 Hz, 2H), 7.00 (td, J = 7.4, 1.5 Hz, 1H), 7.31 (td, J = 7.4, 1.0 Hz, 1H), 7.50 (dd, J = 7.4, 1.5 Hz, 1H), 7.88 (dd, J = 7.4, 1.0 Hz, 1H). 13C-NMR (CDCl3): δ = 64.5, 69.9, 70.0, 71.5, 87.7, 92.6, 100.1, 127.7, 128.7, 130.3, 132.2, 138.6. HRMS (C18H13FeI): Calcd.: 411.9412; found: 411.9410.

3.7. 2-(4-Nitrophenylethynyl)-1-ferrocenylethynylbenzene (8)

2-Ferrocenylethynyl-1-iodobenzene (6, 93.0 mg, 226 μmol), bis(triphenylphosphine)palladium(II) chloride (16.0 mg, 22.9 μmol) and copper(I)iodide (9.00 mg, 47.3 μmol) in anhydrous triethylamine (25 mL) were combined under nitrogen at ambient temperature. Then p-nitrophenylacetylene (7) (33.1 mg, 225 μmol) was added and the reaction temperature was increased up to 50 °C for 3 h. After removal of the solvent under reduced pressure, the residue was dissolved in dichloromethane. The organic phase was extracted with 2.5% aqueous KCN solution (3 × 70 mL) and washed with water (3 × 70 mL). The combined organic layers were dried over anhydrous magnesium sulfate and the solvent was evaporated under atmospheric pressure. The crude product was purified by column chromatography over silica gel with n-hexane/chloroform (1:1, Rf = 0.10) as eluent resulting in 8 in 66% yield (64.0 mg, 148 μmol) as a red solid with a melting point of 161 °C. IR (KBr): ν = 2975 (s), 2599 (s), 2496 (m), 2370 (m), 2202 (w, C≡C), 1590 (m), 1509 (s), 1481 (m), 1438 (m), 1397 (m), 1339 (s), 1172 (w), 1106 (s), 1034 (s) cm−1. 1H-NMR (CDCl3): δ = 4.20 (s, 5H), 4.29 (t, J = 1.8 Hz, 2H), 4.51 (t, J = 1.8 Hz, 2H), 7.30–7.36 (m, 2H), 7.53–7.58 (m, 2H), 7.74 (d, J = 8.9 Hz, 2H), 8.25 (d, J = 8.9 Hz, 2H). 13C-NMR (CDCl3): δ = 64.7, 69.2, 70.1, 71.5, 84.3, 91.1, 93.5, 94.0, 123.7, 124.1, 127.1, 127.4, 129.0, 130.4, 131.8, 132.0, 132.3, 147.7. HRMS (C26H17NO2Fe): Calcd. 431.0609; found: 431.0609.

3.8. Oxidation of 2-(4-Nitrophenylethynyl)-1-ferrocenylethynylbenzene 8 to the Monocation 8+

2-(4-Nitrophenylethynyl)-1-ferrocenylethynylbenzene (8, 35 mg, 81.2 μmol) was dissolved in anhydrous chloroform (3 mL) under nitrogen. Then tris(4-bromophenyl)aminium hexachloroantimonate (66.3 mg, 81.2 μmol) was added until the reaction mixture remained blue. The reaction mixture was stirred for 1 h at room temperature. After the addition of anhydrous n-hexane (5 mL), a green precipitate formed that was filtered under nitrogen and washed with anhydrous n-hexane (50 mL) providing 50.9 mg (66.5 µmol, 82%) of an olive green solid (C26H17NO2FeSbCl6). IR (KBr): ν = 3102 (m), 2212 (m, C≡C), 1595 (m), 1515 (s), 1500 (m), 1420 (w, ferrocene), 1349 (s), 1286 (w), 1173 (m), 1108 (w), 1064 (w), 1002 (w), 853 (s), 761 (m), 749 (m), 688 (w) cm−1. ESI (C26H17NO2Fe+): Calcd. 431.3; found 431.4.

4. Conclusions

While neutral enediynes undergo thermal cyclization at high temperature in full agreement with the computed high activation barriers, the corresponding monocationic enediynes show thermal reactivity at much reduced temperatures. The DFT computations suggest, however, that monocations 3+, 5+ and 8+ neither underwent a cation radical Bergman nor a cation radical Pascal-Schreiner cyclization. Rather they suggest that polymerization was initiated by an intermolecular bond formation, alike as computed for 9+•.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/19/11/18399/s1.

Supplementary Files

Supplementary File 1

Acknowledgments

Financial support from the Deutsche Forschungsgemeinschaft is gratefully acknowledged. We thank Jens Schlirf, Dieter Schmitz and Hans-Jörg Deiseroth (all University of Siegen) for providing us with some data on the X-ray structures of 3 and 8. We are indebted to the High-Performance-Computing (HPC) Linux Cluster HorUS of University of Siegen for computational support.

Author Contributions

The synthesis and DSC investigations were performed by Guido Morbach, while all computations were done by M. Emin Cinar and the manuscript drafted by both. The research project was conceived by Michael Schmittel, who also perfected the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Joshi, M.C.; Rawat, D.S. Recent developments in enediyne chemistry. Chem. Biodivers. 2012, 9, 459–498. [Google Scholar]
  2. Schreiner, P.R.; Navarro-Vazquez, A.; Prall, M. Computational studies on the cyclizations of enediynes, enyne-allenes, and related polyunsaturated systems. Acc. Chem. Res. 2005, 38, 29–37. [Google Scholar]
  3. Poloukhtine, A.; Popik, V.V. Photoswitchable enediynes: Use of cyclopropenone as photocleavable masking group for the enediyne triple bond. Chem. Commun. 2005, 5, 617–619. [Google Scholar]
  4. Odedra, A.; Wu, C.J.; Pratap, T.B.; Huang, C.W.; Ran, Y.F.; Liu, R.S. Ruthenium-catalyzed aromatization of enediynes via highly regioselective nucleophilic additions on a π-alkyne functionality. A useful method for the synthesis of functionalized benzene derivatives. J. Am. Chem. Soc. 2005, 127, 3406–3412. [Google Scholar]
  5. Nath, M.; Pink, M.; Zaleski, J.M. Controlling both ground- and excited-State thermal barriers to Bergman cyclization with alkyne termini substitution. J. Am. Chem. Soc. 2005, 127, 478–479. [Google Scholar]
  6. Lewis, K.D.; Matzger, A.J. Bergman cyclization of sterically hindered substrates and observation of phenyl-shifted products. J. Am. Chem. Soc. 2005, 127, 9968–9969. [Google Scholar]
  7. Kovalenko, S.V.; Alabugin, I.V. Lysine–enediyne conjugates as photochemically triggered DNA double-strand cleavage agents. Chem. Commun. 2005, 1444–1446. [Google Scholar]
  8. Fouad, F.S.; Wright, J.M.; Plourde, G.; Purohit, A.D.; Wyatt, J.K.; El-Shafey, A.; Hynd, G.; Crasto, C.F.; Lin, Y.Q.; Jones, G.B. Synthesis and protein degradation capacity of photoactivated enediynes. J. Org. Chem. 2005, 70, 9789–9797. [Google Scholar]
  9. Bhattacharyya, S.; Pink, M.; Baik, M.H.; Zaleski, J.M. A unique approach to metal-induced Bergman cyclization: Long-range enediyne activation by ligand-to-metal charge transfer. Angew. Chem. Int. Ed. Engl. 2005, 44, 592–595. [Google Scholar]
  10. Basak, A.; Roy, S.K.; Mandal, S. Activation of macrocyclic enediynes by transannular cyclization. Angew. Chem. Int. Ed. Engl. 2005, 44, 132–135. [Google Scholar]
  11. Banfi, L.; Basso, A.; Guanti, G.; Riva, R. Design and synthesis of heterocycle fused enediyne prodrugs activable at will. ARKIVOC 2006, 2006, 261–275. [Google Scholar] [CrossRef]
  12. Kar, M.; Basak, A. Design, synthesis, and biological activity of unnatural enediynes and related analogues equipped with pH-dependent or phototriggering devices. Chem. Rev. 2007, 107, 2861–2890. [Google Scholar]
  13. Gredičak, M.; Jerić, I. Enediyne compounds—New promises in anticancer therapy. Acta Pharm. 2007, 57, 133–150. [Google Scholar]
  14. Liang, Z.-X. Complexity and simplicity in the biosynthesis of enediyne natural products. Nat. Prod. Rep. 2010, 27, 499–528. [Google Scholar]
  15. Perrin, C.L.; Reyes-Rodríguez, G.J. Reactivity of nucleophiles toward a p-benzyne derived from an enediyne. J. Phys. Org. Chem. 2013, 26, 206–210. [Google Scholar]
  16. Krupička, M.; Sander, W.; Marx, D. Mechanical manipulation of chemical reactions: Reactivity switching of Bergman cyclizations. J. Phys. Chem. Lett. 2014, 5, 905–909. [Google Scholar]
  17. Jones, R.R.; Bergman, R.G. p-Benzyne. Generation as an intermediate in a thermal isomerization reaction and trapping evidence for the 1,4-benzenediyl structure. J. Am. Chem. Soc. 1972, 94, 660–661. [Google Scholar]
  18. Lockhart, T.P.; Comita, P.B.; Bergman, R.G. Kinetic evidence for the formation of discrete 1,4-dehydrobenzene intermediates. Trapping by inter- and intramolecular hydrogen atom transfer and observation of high-temperature CIDNP. J. Am. Chem. Soc. 1981, 103, 4082–4090. [Google Scholar]
  19. Lockhart, T.P.; Bergman, R.G. Evidence for the reactive spin state of 1,4-dehydrobenzenes. J. Am. Chem. Soc. 1981, 103, 4091–4096. [Google Scholar]
  20. Lindh, R.; Persson, B.J. Ab initio study of the Bergman reaction: The autoaromatization of hex-3-ene-1,5-diyne. J. Am. Chem. Soc. 1994, 116, 4963–4969. [Google Scholar]
  21. Warner, B.P.; Millar, S.P.; Broene, R.D.; Buchwald, S.L. Controlled acceleration and inhibition of Bergman cyclization by metal chlorides. Science 1995, 269, 814–816. [Google Scholar]
  22. König, B.; Hollnagel, H.; Ahrens, B.; Jones, P.G. Activation of macrocyclicbiaryl-enediynes by metal ion coordination. Angew. Chem. Int. Ed. Engl. 1995, 34, 2538–2540. [Google Scholar]
  23. König, B.; Pitsch, W.; Thondorf, I. Synthesis, structure, and reactivity of enediyne macrocycles. J. Org. Chem. 1996, 61, 4258–4261. [Google Scholar]
  24. McPhee, M.M.; Kerwin, S.M. Synthesis and metal ion binding studies of enediyne—Containing crown ethers. J. Org. Chem. 1996, 61, 9385–9393. [Google Scholar]
  25. Basak, A.; Shain, J.C. Synthesis and thermal behavior of a novel diazaenediyne and its copper(II) complex. Tetrahedron Lett. 1998, 39, 1623–1624. [Google Scholar]
  26. Basak, A.; Shain, J.C. Synthesis and thermal reactivity of a novel macrocyclicenediyne and its copper(II) complex. Tetrahedron Lett. 1998, 39, 3029–3030. [Google Scholar]
  27. Basak, A.; Shain, J.C.; Khamrai, U.K.; Rudra, K.R.; Basak, A. Nitrogen substituted cyclic enediynes: Synthesis, thermal reactivity and complexation with metal ions. J. Chem. Soc. Perkin Trans. 1 2000, 1955–1964. [Google Scholar]
  28. Maier, M.E. Design of enediyne prodrugs. Synlett 1995, 13–26. [Google Scholar]
  29. Grissom, J.W.; Gunawardena, G.U.; Klingberg, D.; Huang, D. The chemistry of enediynes, enyne-allenes and related compounds. Tetrahedron 1996, 52, 6453–6513. [Google Scholar]
  30. Benites, P.J.; Rawat, D.S.; Zaleski, J.M. Metalloenediynes: Ligand field control of thermal Bergman cyclization reactions. J. Am. Chem. Soc. 2000, 122, 7208–7217. [Google Scholar]
  31. Maier, M.E.; Greiner, B. Synthesis and reactivity of a p-methoxyphenyl-substituted enediyne. A case of electronic influence on the rate of the Bergman cycloaromatization. Liebigs Ann. Chem. 1992, 855–861. [Google Scholar]
  32. Nicolaou, K.C.; Liu, A.; Zeng, Z.; McComb, S. Redox-controlled Bergman cycloaromatizations. Designed enediynes with DNA-cleaving properties and antitumor activity. J. Am. Chem. Soc. 1992, 114, 9279–9282. [Google Scholar]
  33. Semmelhack, M.F.; Neu, T.; Foubelo, F. Arene 1,4-diradical formation from o-dialkynylarenes. Tetrahedron Lett. 1992, 33, 3277–3280. [Google Scholar]
  34. Semmelhack, M.F.; Neu, T. Arene 1,4-diradical formation from o-dialkynylarenes. J. Org. Chem. 1994, 59, 5038–5047. [Google Scholar]
  35. Schmittel, M.; Kiau, S. Polar effects in the transition state of the Bergman cyclization. Chem. Lett. 1995, 953–954. [Google Scholar]
  36. Alabugin, I.V.; Manoharan, M.; Kovalenko, S.V. Tuning rate of the Bergman cyclization of benzannelated enediynes with ortho substituents. Org. Lett. 2002, 4, 1119–1122. [Google Scholar]
  37. Alabugin, I.V.; Manoharan, M. Reactant destabilization in the Bergman cyclization and rational design of light- and pH-activated enediynes. J. Phys. Chem. A 2003, 107, 3363–3371. [Google Scholar]
  38. Schmittel, M.; Viola, G.; Dall’Acqua, F.; Morbach, G. A novel concept to activate enediynes for DNA cleavage. Chem. Commun. 2003, 646–647. [Google Scholar]
  39. Schmittel, M.; Röck, M. Reaction of enol cation radicals in the presence of nucleophiles. Chem. Ber. 1992, 125, 1611–1620. [Google Scholar]
  40. Schmittel, M.; von Seggern, H. Electron-transfer catalyzed reactions 3. Ketene diene [4+2] cycloaddition products via cation radical initiated Diels-Alder reaction or vinylcyclobutanone rearrangement. J. Am. Chem. Soc. 1993, 115, 2165–2177. [Google Scholar]
  41. Rodriguez, J.G.; Onate, A.; Martin-Villamil, R.M.; Fonseca, J. A practical synthesis of ethynylferrocene from ferrocenecarboxaldehyde: Structure of 1,4-diferrocenyl-1,3-butadiyne. J. Organomet. Chem. 1996, 513, 71–76. [Google Scholar]
  42. Sonogashira, K.; Tohda, Y.; Hagihara, N. Convenient synthesis of acetylenes. Catalytic substitutions of acetylenic hydrogen with bromo alkenes, iodoarenes, and bromopyridines. Tetrahedron Lett. 1975, 50, 4467–4470. [Google Scholar]
  43. De Meijere, A.; Meyer, F.E. Fine feathers make fine birds: The Heck reaction in modern garb. Angew. Chem. Int. Ed. Engl. 1994, 33, 2379–2411. [Google Scholar]
  44. Rossi, R.; Carpita, A.; Bellina, F. Palladium- and/or copper-mediated cross-coupling reactions between 1-alkynes and vinyl, aryl, 1 alkynyl, 1,2-propadienyl, propargyl and allylic halides or related compounds. A review. Org. Prep. Proc. Int. 1995, 27, 127–160. [Google Scholar]
  45. Heck, R.F. Palladium Reagents in Organic Syntheses; Academic Press: London, UK, 1985. [Google Scholar]
  46. Voigt, K.; von Zezschwitz, P.; Rosauer, K.; Lansky, A.; Adams, A.; Reiser, O.; de Meijere, A. The twofold Heck reaction on 1,2-dihalocycloalkenes and subsequent 6π-electrocyclization of the resulting (E,Z,E)-1,3,5-hexatrienes. A new formal {2+2+2}-assembly of six-membered rings. Eur. J. Org. Chem. 1998, 1521–1534. [Google Scholar]
  47. Plevyak, J.E.; Dickerson, J.E.; Heck, R.F. Selective palladium-catalyzed vinylic substitutions with bromoiodo aromatics. J. Org. Chem. 1979, 44, 4078–4080. [Google Scholar]
  48. Takahashi, S.; Kuroyama, Y.; Sonogashira, K.; Hagihara, N. A convenient synthesis of ethynylarenes and diethynylarenes. Synthesis 1980, 627–630. [Google Scholar]
  49. Thorand, S.; Krause, N. Improved procedures for the palladium-catalyzed coupling of terminal alkynes with aryl bromides (Sonogashira coupling). J. Org. Chem. 1998, 63, 8551–8553. [Google Scholar]
  50. Diallo, A.K.; Absalon, C.; Ruiz, J.; Astruc, D. Ferrocenyl-terminated redox stars: Synthesis and electrostatic effects in mixed-valence stabilization. J. Am. Chem. Soc. 2011, 133, 629–641. [Google Scholar]
  51. Prall, M.; Wittkopp, A.; Schreiner, P.R. Can Fulvenes Form from Enediynes? A Systematic high-level computational study on parent and benzannelated enediyne and enyne–allene cyclizations. J. Phys. Chem. A 2001, 105, 9265–9274. [Google Scholar]
  52. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar]
  53. Lee, C.; Yang, W.; Parr, R.G. Development of the Colic-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar]
  54. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; Gaussian Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  55. Engels, B.; Lennartz, C.; Hanrath, M.; Schmittel, M.; Strittmatter, M. A theoretical and experimental study on the regioselectivity of biradical cyclizations in enyne-allenes: Influence of substituents on the switch from the Myers-Saito to the novel C2-C6 cyclization. Angew. Chem. Int. Ed. Engl. 1998, 37, 1960–1963. [Google Scholar]
  56. Stahl, F.; Moran, D.; Schleyer, P.V.; Prall, M.; Schreiner, P.R. Aromaticity of the Bergman, Myers-Saito, Schmittel, and directly related cyclizations of enediynes. J. Org. Chem. 2002, 67, 1453–1461. [Google Scholar]
  57. Wenthold, P.G.; Lipton, M.A. A density functional molecular orbital study of the C2–C7 and C2–C6 cyclization pathways of 1,2,4-heptatrien-6-ynes. The role of benzannulation. J. Am. Chem. Soc. 2000, 122, 9265–9270. [Google Scholar]
  58. Vavilala, C.; Byrne, N.; Kraml, C.M.; Ho, D.M.; Pascal, R.A. Thermal C1–C5 diradical cyclization of enediynes. J. Am. Chem. Soc. 2008, 130, 13549–13551. [Google Scholar]
  • Sample Availability: Samples of the compounds are not available from the authors.

Share and Cite

MDPI and ACS Style

Cinar, M.E.; Morbach, G.; Schmittel, M. Thermal Reactivity of Neutral and Oxidized Ferrocenyl-Substituted Enediynes. Molecules 2014, 19, 18399-18413. https://doi.org/10.3390/molecules191118399

AMA Style

Cinar ME, Morbach G, Schmittel M. Thermal Reactivity of Neutral and Oxidized Ferrocenyl-Substituted Enediynes. Molecules. 2014; 19(11):18399-18413. https://doi.org/10.3390/molecules191118399

Chicago/Turabian Style

Cinar, Mehmet Emin, Guido Morbach, and Michael Schmittel. 2014. "Thermal Reactivity of Neutral and Oxidized Ferrocenyl-Substituted Enediynes" Molecules 19, no. 11: 18399-18413. https://doi.org/10.3390/molecules191118399

Article Metrics

Back to TopTop