Konrad
Kowalski
*a,
Rafał
Karpowicz
a,
Grzegorz
Mlostoń
b,
Dominique
Miesel
c,
Alexander
Hildebrandt
c,
Heinrich
Lang
c,
Rafał
Czerwieniec
d and
Bruno
Therrien
e
aFaculty of Chemistry, Department of Organic Chemistry, University of Łódź, Tamka 12, PL-91403 Łódź, Poland. E-mail: kondor15@wp.pl
bFaculty of Chemistry, Department of Organic and Applied Chemistry, University of Łódź, Tamka 12, PL-91403 Łódź, Poland
cTechnische Universität Chemnitz, Faculty of Natural Sciences, Institute of Chemistry, Department of Inorganic Chemistry, D-09107 Chemnitz, Germany
dUniversität Regensburg, Institut für Physikalische und Theoretische Chemie, Universitätsstraße 31, D-93040 Regensburg, Germany
eInstitute of Chemistry, University of Neuchatel, Avenue de Bellevaux 51, CH-2000 Neuchatel, Switzerland
First published on 24th February 2015
Three novel diferrocenyl complexes were prepared and characterised. 2,2-Diferrocenyl-4,5-dimethyl-3,6-dihydro-2H-thiopyran (1, sulphide) was accessible by the hetero-Diels–Alder reaction of diferrocenyl thioketone with 2,3-dimethyl-1,3-butadiene. Stepwise oxidation of 1 gave the respective oxides 2,2-diferrocenyl-4,5-dimethyl-3,6-dihydro-2H-thiopyran-1-oxide (2, sulfoxide) and 2,2-diferrocenyl-4,5-dimethyl-3,6-dihydro-2H-thiopyran-1,1-dioxide (3, sulfone), respectively. The molecular structures of 1 and 3 in the solid state were determined by single crystal X-ray crystallography. The oxidation of sulphide 1 to sulfone 3, plays only a minor role on the overall structure of the two compounds. Electrochemical (cyclic voltammetry (= CV), square wave voltammetry (= SWV)) and spectroelectrochemical (in situ UV-Vis/NIR spectroscopy) studies were carried out. The CV and SWV measurements showed that an increase of the sulphur atom oxidation from −2 in 1 to +2 in 3 causes an anodic shift of the ferrocenyl-based oxidation potentials of about 100 mV. The electrochemical oxidation of 1–3 generates mixed-valent cations 1+–3+. These monooxidised species display low-energy electronic absorption bands between 1000 and 3000 nm assigned to IVCT (= Inter-Valence Charge Transfer) electronic transitions. Accordingly, the mixed-valent cations 1+–3+ are classified as weakly coupled class II systems according to Robin and Day.
Ferrocenyl groups are often used in organometallic chemistry as redox-active terminal units, because ferrocene is thermally stable in its neutral and oxidised form.21 In addition, ferrocenyl/ferrocenium groups in mixed-valent species show an excellent electrochemical reversibility of the Fe(II)/Fe(III) redox couple, i.e. 2,5-diferrocenyl five-membered heterocycles.8,22–32
The electron transfer between the Fe(II)/Fe(III) centres, i.e. from the donor (Fe(II)) to the acceptor (Fe(III)) ion, manifests itself by the appearance of characteristic absorptions, i.e. IVCT (= Inter-Valence Charge Transfer) bands in the near infrared (NIR) spectral range.8,33
Two distinct modes of the electronic communication between the donor and acceptor metal centres in the MV state exist: “through bond” and “through space”.5,6,34 The “through bond” mechanism is characteristic for molecules in which π-conjugated connectivities are linking the two redox-active metal termini,5,6 while the “through space” mechanism requires a close proximity of the interacting centres.34 Depending on the degree of the electronic communication, three classes of MV complexes are distinguished according to the Robin and Day classification: non-coupled (class I), weakly-coupled (class II) and fully delocalized (class III) systems.35 Based on the linking group constitution, five structural types of dinuclear transition-metal compounds are known (types A–E), as schematically shown in Fig. 1.
![]() | ||
Fig. 1 Classification of homobimetallic mixed-valence (= MV) metal complexes. M represents a metal containing redox centre. In type A molecules the centres are connected by a conjugated bridge consisting of C,C double or triple bonds, or aromatic moieties.5,6 In type B molecules the M termini are directly connected to each other,36 while in type D and E species the metal centres are separated by an aliphatic37,38 or heteroatom fragment, respectively.37–39 Type C molecules represent a special variety of type A compounds, in which the two M termini in the mixed-valent species interact most likely “through space”.34 |
Among the type A and type B molecules (Fig. 1), the respective diferrocenyl-functionalized systems have been extensively studied,5,6,36,40–42 whereas compounds of structural type D remain almost unexplored.37,43 The simplest representative of the Fc-based class D molecules is diferrocenylmethane (Fc2CH2, Fc = Fe(η5-C5H4)(η5-C5H5)), however, its mono-oxidised form (Fc2CH2+) does not show any detectable IVCT band.44 On the contrary, such IVCT absorptions were observed for other derivatives, in which the methylene hydrogen atoms were replaced by methyl or ferrocenyl groups,37,44 indicating that the appearance of IVCT transitions is related to the increased bulkiness of the linker unit due to the presence of sterically demanding groups in the latter compounds as compared to the Fc2CH2+ cation. In addition, the strength of the metal–metal interactions in type D compounds featuring CMe2, SiMe2 or GeMe2 bridges, depend on the effective distance between the metal centres.37 It was found that the extent of the electronic coupling in the MV species decreases with an increase of the atomic radius of the bridging atom (e.g. in the order CMe2 > SiMe2 > GeMe2).
In continuation of our works in the area of electron-transfer studies in (multi)ferrocenyl-functionalised organic and organometallic compounds,8,11,22–32,45–49 we herein present the synthesis and characterisation of 2,2-diferrocenyl-4,5-dimethyl-3,6-dihydro-2H-thiopyran, 2,2-diferrocenyl-4,5-dimethyl-3,6-dihydro-2H-thiopyran-1-oxide and 2,2-diferrocenyl-4,5-dimethyl-3,6-dihydro-2H-thiopyran-1,1-dioxide. The influence of the electronic and steric effects on the electron transfer between the two ferrocenyl moieties in the respective mixed-valent species is reported. The 3,6-dihydro-2H-thiopyran bridge was chosen, due to its bulkiness and the presence of the oxidisable sulphur atom in a position adjacent to the redox-active ferrocenyl groups. This S-atom reactivity allows us to synthesise three closely related derivatives 1–3 and investigate the influence of the different electron-withdrawing character of the aliphatic bridge on the IVCT properties of the mixed valence species 1+–3+.
Sulphide 1 was prepared via the hetero-Diels–Alder cycloaddition of diferrocenylthioketone50,51 with 2,3-dimethyl-1,3-butadiene in a sealed glass-tube at 75 °C (Experimental section).52 After appropriate work-up, compound 1 was obtained as a red solid in 65% yield. Treatment of sulphide 1 with hydrogen peroxide (30%) and selenium dioxide as oxidising reagents53 in methanol gave the respective sulfoxide 2, which was purified by column chromatography in 87% yield (Experimental section). Further oxidation of 2 with m-chloroperoxybenzoic acid (= MCPBA) in dichloromethane produced sulfone 3 at ambient temperature in 21% yield (Scheme 1, Experimental section).54
Compounds 1–3 are red solids soluble in common organic solvents. All three compounds are stable towards air and moisture in the solid state and in solution.
Compounds 1–3 were characterised by IR and NMR (1H, 13C{1H}; for more details see Fig. S1–S3, ESI†) spectroscopy and high-resolution mass spectrometry. The molecular structures of 1 and 3 in the solid state were determined by single crystal X-ray diffraction analysis. Electrochemical investigations were carried out by using cyclic voltammetry (= CV), square wave voltammetry (= SWV) and in situ UV-Vis/NIR spectroelectrochemistry.
![]() | ||
Fig. 2 ORTEP diagrams of 1 and 3 (two independent molecules in the crystal, 3A and 3B) at 50% probability level. |
1 | 3A | 3B | |
---|---|---|---|
Bond distances | |||
Fe⋯Fe | 5.6377(6) | 5.6824(9) | 5.6785(9) |
S–CFc | 1.850(2) | 1.834(3) | 1.832(3) |
S–CH2 | 1.799(2) | 1.767(3) | 1.771(3) |
CFc–CH2 | 1.532(3) | 1.530(4) | 1.542(4) |
C![]() |
1.339(3) | 1.344(5) | 1.331(5) |
CFc–CCp | 1.515(3) | 1.527(4) | 1.518(5) |
CFc–CCp | 1.517(3) | 1.527(4) | 1.526(4) |
Fe–centroid (Cp–CFc) | 1.657 | 1.656 | 1.653 |
Fe–centroid (Cp–CFc) | 1.658 | 1.656 | 1.643 |
Fe–centroid (Cp) | 1.660 | 1.659 | 1.646 |
Fe–centroid (Cp) | 1.658 | 1.658 | 1.650 |
Bond angles | |||
S–CFc–CH2 | 107.91(13) | 104.9(2) | 105.3(2) |
CFc–S–CH2 | 97.30(10) | 100.65(16) | 101.20(16) |
CCp–CFc–CCp | 111.20(16) | 112.1(2) | 110.0(3) |
O![]() ![]() |
117.81(15) | 117.57(16) |
Compound 1 crystallizes in the orthorhombic space group Pcab, while complex 3 crystallizes in the triclinic space group P. In the crystal packing of 3, two independent molecules (A and B) are observed. The structure analysis of both molecules confirmed the expected structures in which the two ferrocenyl groups are bonded to a sulphide (1) or a sulfone (3) moiety (Fig. 2). All ferrocenyl units show an eclipsed conformation with nearly equivalent distances between Fe and the centroid of the cyclopentadienyl rings (Table 1).
The oxidation of sulphide 1 to sulfone 3, however, plays only a minor role on the overall structure of the two compounds. A similar behaviour was found for a series of sulphur-containing heterocycles in which the oxidation of the sulphur atom has almost no impact on the geometrical parameters of the heterocycles.54,55 The only bond distances which are somewhat influenced by oxidation are S–CFc and S–CH2 (Table 1), whereby the corresponding bonds in 3, as compared to 1, are shortened by 0.02 Å, as previously observed by Petrov.54,55 In both compounds, the thiopyran ring adopts a half-chair conformation to limit the steric repulsion between the different substituents (Fig. 2). In these carbon-bridged diferrocenyl complexes, the Fe⋯Fe distances are 5.6377(6) (1), 5.6824(9) (3A) and 5.6785(9) Å (3B), respectively. These values are comparable to those found in analogous carbon-bridged diferrocenyl compounds.56 Interestingly, the dihedral angles between the two planes of the covalently bonded cyclopentadienyl rings are quite acute in 1 (61.4°) and molecule 3A (63.1°), while in 3B this dihedral angle is normal at 76.6°.56
Compounds 1–3 show relatively weak absorptions in the visible region and stronger absorptions at higher energies (Fig. 3) as it is characteristic for ferrocene derivatives.57,58 TD-DFT (= Time Dependent Density-Functional Theory) calculations (Fig. 3) performed for 1 predict eight lowest-energy transitions between 450 and 530 nm in which mainly Fe-centred molecular orbitals of 3d-character are involved.58 The low energy spectral features are, thus, assigned to an unresolved series of broad overlapping bands resulting from ferrocene-centred d–d transitions.40,59,60
The electrochemical measurements (CV and SWV) were performed under an argon atmosphere in dichloromethane solutions containing [Bu4N][B(C6F5)4] (0.1 mol L−1) as supporting electrolyte at a scan rate of 100 mV s−1 at 25 °C. The data of the cyclic voltammetry experiments are summarised in Table 2. All potentials are referenced to the FcH/FcH+ redox couple.61 The voltammograms of 1–3 are shown in Fig. 5.
Compd. | E °′1 [mV] (ΔEpc [mV]) | E °′2 [mV] (ΔEpc [mV]) | ΔE°′![]() |
---|---|---|---|
a Potentials vs. FcH/FcH+, scan rate 100 mV s−1 at glassy carbon electrode of a 1.0 mmol L−1 solution in dry dichloromethane; 0.1 mol L−1 [NnBu4][B(C6F5)4] as supporting electrolyte at 25 °C. b E°′ = Formal potential. c ΔEp = difference between the oxidation and the reduction potential. d ΔE°′ = potential difference between the two ferrocenyl-related redox processes (E°′2–E°′1). | |||
1 | −90 (73) | 275 (75) | 365 |
2 | −15 (75) | 355 (76) | 370 |
3 | 5 (73) | 375 (83) | 370 |
As it can be seen from Fig. 5, both ferrocenyl groups in 1–3 can be oxidised separately showing two well-resolved reversible one-electron oxidation steps. The chemical oxidation of the sulphur atom in the neighbouring position to the ferrocenyl units leads to an anodic shift of the both Fc-based oxidation processes, E°′1 from 90 mV (1) over −15 mV (2) to 5 mV (3) and E°2 from 275 mV (1) over 355 mV (2) to 375 (3), respectively. This chemical oxidation process increases the group-electronegativity at the sulfur (for example, see group electronegativity: –SMe = 2.592; –SOMe = 2.841; –SO2Me = 2.998)62 and hence, reduces the electron density at the ferrocenyl groups due to the increased electron withdrawing effect. This trend is also reproduced in the results of the DFT calculations (see below). The Fe-centred occupied frontier orbitals undergo substantial stabilisation upon oxidation of the sulphur atom. For example, the HOMO energy in 3 is about 60 mV smaller than that in non-oxidised 1 (Fig. 4).
![]() | ||
Fig. 4 DFT energies of the four lowest occupied orbitals of 1–3 and contour plots of the corresponding Kohn–Sham orbitals for 3. |
The redox separation between the 1st and the 2nd oxidation processes, however, is not affected by the different degree of the sulphur oxidation and is ca. 370 mV throughout the series. Due to the use of [Bu4N][B(C6F5)4]25,63–69 as weakly coordination counter-ion within the electrolyte, the ion-pairing effects are minimised70–72 and thus, the electrostatic stabilisation forces between the ferrocenyl groups are increased, when compared with diferrocenylmethane measured in [Bu4N][ClO4] (ΔE°′ = 120 mV).73
The UV-Vis/NIR spectroelectrochemical measurements were performed in an OTTLE (= Optically Transparent Thin-Layer Electrochemistry) cell74 using an analyte concentration of 2.0 mmol L−1 and [Bu4N][B(C6F5)4] (0.1 mol L−1) as supporting electrolyte in dichloromethane (1–3) or acetonitrile (2). The UV-Vis/NIR spectra are depicted in Fig. 6 (1), 7 (2), and 8 (3). The spectrum measured in acetonitrile (2) is shown in Fig. S5.†
The appropriate compounds were oxidised by stepwise increasing the potentials (step width 25, 50 and 100 mV). Thus, the studied compounds 1–3 underwent oxidation to the mono-cationic 1+–3+ and di-cationic 12+–32+ species, respectively. After complete oxidation, each sample was reduced at −200 mV to prove the reversibility of the redox processes. The resulting UV-Vis/NIR spectra were identical to those of the starting molecules. During the oxidation of 1–3 a broad band with a very low intensity (εmax = 100 L mol−1 cm−1) between 1000 and 3000 nm appeared (Fig. 6–8). A further increase of the potential resulted in the decrease of this band. Such a behaviour is typically observed for intervalence charge transfer (= IVCT) absorptions.2,13 The experimental spectra can be deconvoluted into three Gaussian-shape bands assigned to an IVCT, a ligand field transition, and a band representing the edge to the higher energy absorptions. The sum of these three Gaussian-shaped bands fits almost exactly with the experimental spectra. The deconvolution reveals the intensity εmax, the full width-at-half-height Δν1/2 and the νmax values for the IVCT component. The solvent polarity change from P = 3.1 (dichloromethane) to P = 5.8 (acetonitrile),75 resulting in a shift of the νmax value from 5250 cm−1 to 7525 cm−1. It is remarked that strong solvatochromic shifts are expected for IVCT absorption bands being of distinct charge transfer character. Thus, the IVCT assignment of the observed NIR absorption features (Fig. 6–8 and Table 3) is further substantiated. The appearance of low energy ligand field transitions is characteristic for ferrocenyl containing compounds as for example demonstrated by UV/Vis-NIR measurements of mono ferrocenyl thiophenes.27 The numerical data derived from the deconvolution procedure is summarised in Table 3. However, the data should be handled with care since the very small extinctions of these bands cause them to be susceptible to errors.
Compd. | ν max (cm−1) (εmax (L mol−1 cm−1)) | Δν1/2 (cm−1) | (Δν1/2)theob (cm−1) |
---|---|---|---|
a Measured in dry dichloromethane (DCM) or acetonitrile (ACN) using [Bu4N][B(C6F5)4] (0.1 mol dm−3) as supporting electrolyte at 25 °C. b Values calculated as (Δν1/2)theo = (2310 νmax)1/2 according to the Hush relationships for weakly coupled systems.76 | |||
1 + (DCM) | 5200 (100) | 4950 | 3468 |
2 + (DCM) | 5250 (100) | 4900 | 3478 |
2 + (ACN) | 7525 (60) | 7850 | 4169 |
3 + (DCM) | 5300 (95) | 4900 | 3512 |
The Δν1/2 values exceed the theoretical values according to the Hush model, which is a common behaviour for weakly coupled class II systems according to Robin and Day35 and is caused by interactions with the respective solvent. The absorption energy is slightly blue shifted with increasing oxidation state of the sulphur atom (the νmax value increases from 5200 cm−1 for 1+ to 5300 cm−1 determined for 3+). However, these differences are small and lie within the margin of the experimental error of our measurements. Thus, the sulphur oxidation state does not have any significant influence on the electron transfer properties of diferrocenes 1+–3+. Remarkably, for the diferrocenylmethane cation (Fc2CH2+) no IVCT absorptions could be detected, while for the mixed-valent triferrocenylmethane (Fc3CH+) weak charge transfer excitations (εmax = 165 L mol−1 cm−1; Δν1/2 = 3750 cm−1; νmax = 5900 cm−1) were found, possessing similar characteristics as mixed-valent 1+–3+.44
Electrochemical measurements of 1.0 mmol L−1 dichloromethane solutions of 1–3 were performed in a dried, argon purged cell at 25 °C with a Radiometer Voltalab PGZ 100 electrochemical workstation interfaced with a personal computer. Dichloromethane solutions (0.1 mol L−1) containing [Bu4N][B(C6F5)4] were used as supporting electrolyte. For the measurements a three electrode cell containing a Pt auxiliary electrode, a glassy carbon working electrode (surface area 0.031 cm2) and an Ag/Ag+ (0.01 mmol L−1 [AgNO3]) reference electrode fixed on a Luggin capillary was applied. The working electrode was pretreated by polishing on a Buehler microcloth first with a 1 micron and then with a 1/4 micron diamond paste. The reference electrode was constructed from a silver wire inserted into a 0.01 mmol L−1 [AgNO3] and 0.1 mol L−1 [Bu4N][B(C6F5)4] acetonitrile solution in a Luggin capillary with a Vycor tip. This Luggin capillary was inserted into a second Luggin capillary containing a 0.1 mol L−1 [Bu4N][B(C6F5)4] dichloromethane solution and a Vycor tip. Experiments under the same conditions showed that all reduction and oxidation potentials were reproducible within 5 mV. Experimental potentials were referenced against an Ag/Ag+ reference electrode but the presented results are referenced against ferrocene as an internal standard as required by IUPAC.61 To achieve this, each experiment was repeated in the presence of 1 mmol L−1 decamethylferrocene (= Fc*). Data were processed on a Microsoft Excel worksheet to set the formal reduction potentials of the FcH/FcH+ couple to 0.0 V. Under our conditions the Fc*/Fc*+ couple was at −619 mV vs. FcH/FcH+ (ΔEp = 60 mV), while the FcH/FcH+ couple itself was at 220 mV vs. Ag/Ag+ (ΔEp = 61 mV).81 Spectroelectrochemical UV-Vis/NIR measurements of 2.0 mmol L−1 solutions of 1–3 in dichloromethane (1–3) or acetonitrile (2) containing 0.1 mol L−1 of [Bu4N][B(C6F5)4] as the supporting electrolyte were performed in an OTTLE (OTTLE = Optically Transparent Thin-Layer Electrochemistry)74 cell with a Varian Cary 5000 spectrophotometer at 25 °C. The values obtained by deconvolution could be reproduced within εmax, 100 L mol−1 cm−1; νmax, 50 cm−1 and Δν1/2, 50 cm−1.
1H NMR (600 MHz, CDCl3): δ = 4.26 (bs, 2H, C5H4), 4.20 (bs, 2H, C5H4), 4.13 (bs, 2H, C5H4), 4.11 (bs, 2H, C5H4), 4.07 (s, 10H, C5H5), 3.00 (s, 2H, CH2), 2.64 (s, 2H, CH2), 1.85 (s, 3H, CH3), 1.74 (s, 3H, CH3). 13C NMR (150 MHz, CDCl3): δ = 126.6, 124.3, 97.1, 69.0, 66.9, 66.8, 66.7, 66.4, 44.5, 43.3, 31.3, 20.4, 19.4. FTIR (KBr): 3088, 2989, 2911, 2872, 1628, 1443, 1409, 1264, 1106, 1032, 1000, 824, 483 cm−1. MS (ESI): m/z = 496 (M+). MS (EI, 70 eV): m/z = 496 (M+). HRMS: m/z = 496.0614 (Calc. for C27H28SFe2: 496.0610). Anal. Calcd for: C27H28SFe2: C, 65.35; H, 5.69; S, 6.46%. Found: C, 65.10; H, 5.57; S, 6.72%.
1H NMR (600 MHz, CDCl3): δ = 4.90–4.89 (m, 1H, C5H4), 4.36–4.35 (m, 1H, C5H4), 4.32–4.31 (m, 1H, C5H4), 4.30–4.29 (m, 1H, C5H4), 4.28 (s, 5H, C5H5), 4.27–4.26 (m, 1H, C5H4), 4.22–4.21 (m, 1H, C5H4), 4.05–4.04 (m, 1H, C5H4), 4.00–3.99 (m, 1H, C5H4), 3.88 (s, 5H, C5H5), 3.18 (d, JH,H = 18.6 Hz, 1H, CH2), 3.01 (d, JH,H = 18.0 Hz, 1H, CH2), 2.97 (d, JH,H = 15.6 Hz, 1H, CH2), 2.50 (d, JH,H = 15.6 Hz, 1H, CH2), 1.84 (s, 3H, CH3), 1.68 (s, 3H, CH3). 13C NMR (150 MHz, CDCl3): δ = 127.0, 118.6, 94.3, 87.1, 70.1, 69.8, 69.2, 68.9, 68.7, 68.4, 66.9, 66.7, 66.2, 66.1, 59.4, 51.1, 41.6, 20.0, 19.4. FTIR (KBr): 3089, 2917, 2891, 2857, 1629, 1406, 1104, 1053 (s, SO), 1031, 1000, 824, 487 cm−1. MS (ESI): m/z = 535 (MNa+), 513 (MH+). MS (EI, 70 eV): m/z = 512 (M+). HRMS: m/z = 512.0558 (Calc. for C27H28OSFe2: 512.0560). Anal. Calcd for: C27H28OSFe2: C, 63.30; H, 5.51; S, 6.26%. Found: C, 63.23; H, 5.72; S, 6.27%.
1H NMR (600 MHz, CDCl3): δ = 4.47 (bs, 2H, C5H4), 4.24 (s, 4H, C5H4), 4.21–4.20 (pq, JH,H = 1.98 Hz, 1.74 Hz, 2H, C5H4), 4.13 (s, 10H, C5H5), 3.32 (s, 2H, CH2), 3.06 (s, 2H, CH2), 1.91 (s, 3H, CH3), 1.69 (s, 3H, CH3). 13C NMR (150 MHz, CDCl3): δ = 127.0, 119.9, 89.0, 69.6, 68.7, 67.6, 67.5, 67.2, 62.5, 51.6, 44.5, 19.9, 19.6. FTIR (KBr): 3103, 2918, 2857, 1629, 1305 (s, SO2), 1121(s, SO2), 820, 480 cm−1. MS (ESI): m/z = 551 (MNa+), 528 (M+). MS (EI, 70 eV): m/z = 528 (M+), 464 (M+ − SO2), 462 (M+ − H2SO2). HRMS: m/z = 528.0510 (Calc. for C27H28O2SFe2: 528.0509). Anal. Calcd for: C27H28O2SFe2: C, 61.39; H, 5.34; S, 6.07%. Found: C, 61.20; H, 5.48; S, 5.91%.
CCDC 1031233 (1) and 1031234 (3) contain the supplementary crystallographic data for this paper.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 1031233 and 1031234. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c5dt00246j |
This journal is © The Royal Society of Chemistry 2015 |