Aleksandra Mitrovića,
Jelena Stevanovića,
Milos Milčića,
Andrijana Žekićb,
Dalibor Stankovićc,
Shigui Chend,
Jovica D. Badjićd,
Dragana Milića and
Veselin Maslak*a
aFaculty of Chemistry, University of Belgrade, Studentski trg 16, P. O. B. 51, 11158 Belgrade, Serbia. E-mail: vmaslak@chem.bg.ac.rs
bFaculty of Physics, University of Belgrade, Studentski trg 12, 11000 Belgrade, Serbia
cInnovation Center of the Faculty of Chemistry, Studentski trg 12, 11000 Belgrade, Serbia
dDepartment of Chemistry and Biochemistry, The Ohio State University, 100 West 18th Avenue, Columbus, Ohio 43210, USA
First published on 12th October 2015
We synthesized three dumbbell-like compounds 2a–c, each containing two C60 groups at the periphery and pyromellitic diimide (PMDI) in the middle, and examined their electronic as well as assembly characteristics with both experimental and computational methods. Cyclic voltammetry (CV) measurements revealed that each of three electron-accepting (AAA) triads could accommodate up to eight electrons. Computational studies (density functional theory, DFT) of 2a–c at PBEPBE/6-311G(d,p) level of theory, with B3LYP/6-31G(d) optimized geometries, revealed that HOMO–LUMO energy gaps are similar to those of the model compound [6,6]-phenyl-C61-butyric acid methyl ester (PCBM). Compounds 2a–c were also found to assemble into vesicles and nanoparticles on the copper grid (100–300 nm, TEM), while giving more sizeable aggregates after a deposition on the glass (SEM, >5 μm). Understanding the packing of 2a–c on various solid substrates, as well as the assembly characteristics in general, is important for tuning the properties and fabrication of electronic/optical devices. On the basis of the results of conformational analysis (MM and DFT calculations), we deduced that different alkyl spacers in 2a–c ought to play a role in π–π interactions between the aromatic components of the triad to guide the packing and therefore morphology of the material.
Still, fullerene-building blocks are yet to become essential components of practical electrical conductors, photovoltaic cells, electroluminescence devices and field-effect transistors.16 It appears that the key challenge for constructing fullerene-based devices, having a desired performance, is in achieving a high level of control over the material's morphology at the nano scale.17 In line with it, directional noncovalent forces, such as hydrogen bonding and aromatic interactions, could enable a control of the self-assembly process for creating more ordered structures to perhaps permit the bottom-up engineering of fullerene-based devices.18–20
In accord with the discussion, dumbbell-like molecules (Fig. 1) consisting of two π-bridged fullerenes (C60–π system–C60) have been of an interest to both chemists and materials scientists. Due to the presence of two spheroid units, at the periphery, these triads are expected to have an improved solubility, such enabling more extensive examination and potential application. In addition, the fullerene groups can at periphery act as “molecular alligator clips” for anchoring to give a low spread of low-bias conductance.21 As examples of symmetric triads prepared recently, the following electron-donating units have been used to comprise the central core: oligomers (OPVs, OTVs, OPEs),22–25 triphenylamines (TPhAs),26 phenyl/fluorene,21 dithienosilole-dibenzothiadiazole (DTSDBT),27 tetrathiafulvalene (TTF),28 6b,10b-dihydrobenzo[j]cyclobut[a]acenaphthylene (DBCA).29 Interestingly, only few systems have been made to carry an electron-accepting central unit. In one instance, Champness, Khlobystov et al. examined the utility of the triad having 3,4,9,10-perylene tetracarboxylic diimide (PTCDI) group bridging two fullerenes. This system, with multiple redox sites, was capable of accommodating up to six electrons in a predictable and controllable manner.30 Naphtalenetetracarboxylic diimide (NTCDI), as another representative, have been employed to act as a component of fullerene-based rotaxanes.27 These compounds contained two C60 fullerenes, acting as stoppers at both termini, while the NTCDI unit would form noncovalent contacts with the rotaxane's macrocycle.27 In line with these efforts, we hereby describe a study of three symmetric triads 2a–c containing two fulleropyrrolidines covalently attached to pyromellitic diimide (PMDI) via alkyl chains (Scheme 1). We first prepared and characterized compounds 2a–c (C60–PMDI–C60, Scheme 1) and then examined their electronic, electrochemical and assembly characteristics.
:
1 mixture (1.2 mL) was irradiated in a microwave reactor for 30 min, with an inner temperature of 130 °C and applied pulse of 300 W. Crude product was precipitated by the addition of methanol and then purified by column chromatography (SiO2) with CH3CO2CH2CH3/toluene = 1/9 mixture as an eluent. Subsequent precipitation with methanol, from a highly concentrated solution in CHCl3, gave pure 2a–c as brown powders in 42–55% yields.
: 2930, 1719, 1370, 1093, 760, 481 cm−1. UV-VIS (CHCl3) λ (ε, M−1 cm−1): 430 (4200), 704 (380). 1H (CDCl3, 500 MHz) δ 8.37 (s, 4H), 4.48 (s, 8H), 4.47 (s, 4H), 4.34–4.28 (m, 4H), 3.50–3.44 (m, 4H). 13C (CDCl3, 125 MHz) δ 166.2 (4C); 154.8 (Cf-12); 147.5 (Cf-17); 146.4 (Cf-7); 146.1 (Cf-11); 145.6 (Cf-16); 145.5 (Cf-5); 144.7 (Cf-9); 143.3 (Cf-15); 142.8 (Cf-8); 142.2 (Cf-6); 142.1 (Cf-14); 142.0 (Cf-4); 141.8 (Cf-12,13); 141.1 (Cf-10); 140.3 (Cf-3); 138.8 (4C), 136.2 (4C); 118.9 (2CH); 70.8 (4C); 67.5 (4CH2); 37.5 (2CH2); 33.8 (2CH2). MALDI/TOF: m/z calcd for [C138H18N4O4+H]+ 1796, measured 1796.
: 2925, 1719, 1368, 1118, 728, 526 cm−1. UV-VIS (CHCl3) λ (ε, M−1 cm−1): 430 (5200), 704 (380). 1H (CDCl3, 500 MHz) δ 8.24 (s, 4H); 4.36 (s, 8H); 3.94 (t, 4H, J = 6.5); 3.13 (t, 4H, J = 7.5); 2.12–2.07 (m, 4H); 1.99–1.94 (m, 4H). 13C (CDCl3, 125 MHz): δ 165.6 (4C); 154.5 (Cf-12); 147.0 (Cf-17); 146.0 (Cf-7); 145.8 (Cf-11); 145.7 (Cf-16); 145.4 (Cf-5); 145.2 (Cf-9); 145.0 (Cf-15); 144.3 (Cf-8); 142.9 (Cf-6); 142.4 (Cf-14); 142.0 (Cf-4); 141.8 (Cf-12,13); 141.7 (Cf-10); 140.0 (Cf-3); 137.0 (4C); 136.0 (4C); 118.0 (2CH); 70.3 (4C); 67.7 (4CH2); 53.8 (2CH2); 38.3 (2CH2); 26.4 (2CH2); 26.1 (2CH2). MALDI/TOF: m/z calcd for [C142H26N4O4+H]+ 1853, measured 1853.
: 2933, 1721, 1365, 1118, 731, 525 cm−1. UV-VIS (CHCl3) λ (ε, M−1 cm−1): 430 (6600), 704 (380). 1H (CDCl3, 500 MHz) δ 8.21 (s, 4H); 4.37 (s, 8H); 3.79 (t, 4H, J = 7); 3.07 (t, 4H, J = 7.5); 1.96–1.90 (m, 4H); 1.86–1.79 (m, 4H); 1.74–1.68 (m, 4H); 1.58–1.52 (m, 4H). 13C (CDCl3, 125 MHz) δ 165.6 (4C); 154.7 (Cf-12); 147.0 (Cf-17); 146.0 (Cf-7); 145.8 (Cf-11); 145.8 (Cf-16); 145.4 (Cf-5); 145.2 (Cf-9); 145.0 (Cf-15); 144.3 (Cf-8); 142.9 (Cf-6); 142.4 (Cf-14); 142.0 (Cf-4); 141.8 (Cf-12,13); 141.7 (Cf-10); 140.0 (Cf-3); 137.0 (4C); 136.0 (4C); 117.8 (2CH); 70.4 (4C); 67.7 (4CH2); 54.6 (2CH2); 38.4 (2CH2); 28.7 (2CH2); 28.5 (2CH2); 27.1 (2CH2); 26.8 (2CH2). MALDI/TOF: m/z calcd for [C146H34N4O4+H]+ 1909, measured 1909.| Ecomplex = Eint − Edef | (1) |
Due to the large size and conformational flexibility of 2a–c, detailed computational study was performed only on compound 2a containing two CH2 groups. The conformational search was performed with molecular mechanic force field implemented in AMMP program, using systematic search method with Vega ZZ (version 3.0) interface.41 The geometries of three most abundant conformers of compound 2a were fully optimized in gas phase with B3LYP functional and 6-31G(d) basis set with included Grimme's D2 dispersion correction. Energies of the conformers were calculated on B3LYP/6-31G(d) optimized geometries using same functional and dispersion correction, but with larger 6-311G(d,p) basis set in toluene solution. Solvent effect was simulated with SMD implicit solvation method.42 Non-Covalent Interactions (NCI) visualization index is calculated using NCI plot program.43 HOMO and LUMO orbital energies are calculated in gas phase, on B3LYP/6-31G(d) optimized geometries, using PBEPBE functional and relatively large basis set 6-311G(d,p). This method was recently proven to be quite reliable in calculating HOMO–LUMO energies and other molecular properties in fulleropyrrolidine44 and PCBM-like fullerene derivatives.45
Properties of the first excited state were investigated using TD-DFT technique, with B3LYP functional and 6-311G(d,p) basis set. The Fukui function and spin densities were calculated with B3LYP/6-311G(d,p) method by adding electron to B3LYP/6-31G(d) optimized geometries. All the orbitals were drown in gOpenMol46 program from HF/6-311G(d,p) wavefunction.
All the quantum chemical calculations were done in Gaussian 09 program package.47
The UV-vis spectra of 2a–c appear to be a superposition of those corresponding to fulleropyrrolidine and PMDI components. The emission spectra of 2a–c (400–850 nm, Fig. S1†) were collected after excitation at 400 nm and are in good agreement with the absorption features. The *0 → 0 emission (λmax, 709 nm) and 0 → *0 absorption bands (λmax, 704 nm) differ very slightly to provide additional evidence for the absence of a charge transfer interaction between two chromophores.
:
1 solution mixture (Fig. 3). In order to understand the electron transfer pathways, CV data were collected for structurally similar reference compounds N,N-dihexylpyromellitic diimide (DHPMDI) and N-methylfulleropyrrolidine (NMFP).
![]() | ||
Fig. 3 Cyclic voltammogram (CV) of 2a at 50 mV s−1 rate in o-dichlorobenzene : DMF = 2 : 1 mixture, containing 0.1 M nBu4NPF6. | ||
The half-wave potential values for all compounds are given in Table 1. In the potential range between 0 and −2.8 V five quasi-reversible reductions were observed for 2a and 2b, whereas 2c showed four reduction waves. For illustrative clarity, CV curves for 2b–c and reference compounds are omitted from Fig. 3 and can be found in Fig. S2.† As can be seen from Table 1, the half-wave potential values, associated with the first fullerene reduction, are cathodically shifted by 100–120 mV, compared to the parent C60. Indeed, a decrease in the electron activity of the functionalized fullerene is expected due to a partial loss in π-conjugation.49 In the second reduction process, one electron is accepted by the PMDI moiety: the half wave potential of −1.16 V for 2a (−1.13 V for 2b, −1.15 V for 2c) compared with DHPMDI (1.10 V) indicates the increase of electronic density on this electron-poor organic addend. Third E1/2 value can be assigned to fullerene cages due to its good correlation with second reduction peak of the model NMFP (−1.28 V, Table 1). As the reduction goes to more negative potentials, the assignment of the redox processes, based on a similarity to the corresponding model compounds, becomes less straightforward. We assign reduction peaks IV and V, to the second PMDI and third C60 reduction, respectively. For compound 2c, the last two reduction peaks could not be resolved by changing the scan rate of the experiment. A gradual broadening of the redox waves might be ascribed to the presence of longer and more flexible alkyl chains (six CH2 groups) causing a decrease in the rate constant characterizing electron transfer processes at the electrode surface.
In line with reductions of other related symmetric C60-dimers and triads,30,50 each of the fullerene reduction waves is in 2a–c commensurate with a two-electron uptake. The peak current of these reductions is thus higher than for those associated with the PMDI reductions, indicating that the latter correspond to a one-electron uptake.
Cyclic voltammograms of all three triads (Fig. S2†) show that C60 and PMDI moieties are accommodating electrons independently of each other. It should also be noted that no significant influence of the alkyl spacer on the electrochemical properties was found, except for the dumbbell 2c where a greater conformational flexibility of alkyl chains, perhaps, causes a poor resolution of the reduction peaks. As shown in Table 2, dumbbell-like 2a–c display a substantial electrochemical response with the ability of accommodating up to eight electrons to emphasize its possible use as an PCBM-like electron acceptor in polymer solar cells.
| Compd | Parameters | PBEPBE/6-311G(d,p)a |
|---|---|---|
| a Cited from ref. 44.b Extended conformer.c Single-folded conformer.d Double-folded conformer. | ||
| C60 | HOMO/eV | −5.867 |
| LUMO/eV | −4.181 | |
| Egap/eV | 1.686 | |
| PCBM | HOMO/eV | −5.522 |
| LUMO/eV | −4.006 | |
| Egap/eV | 1.516 | |
| 2a | HOMO/eV | −5.575b, −5.447c, −5.417d |
| LUMO/eV | −4.234b, −4.238c, −4.257d | |
| Egap/eV | 1.341b, 1.209c, 1.16d | |
A π–π aromatic interaction between fullerene and PMDI (as a model complex) were investigated computationally with B97-D/TZVP method. In the optimized geometry of fullerene–PMDI complex (Fig. 4a) a PMDI molecule has adopted a concave geometry with 8.0° deviation from planarity. This bending of the PMDI molecule allows a greater contact surface between molecules which could give rise to stronger π–π interactions.35 The calculated complexation energy (Ecomplex) is 9.76 kcal mol−1. This complexation energy is relatively small and only half the size of the complexation energy of fullerene with corannulene (17.03 kcal mol−1) and sumanene (18.47 kcal mol−1) aromatics, both well known as concave hosts for fullerene. Despite, the obtained values indicate a possibility for noncovalent intermolecular interactions in compounds 2a–c, especially in polar non-aromatic solvents.
The nature of the interaction was further investigated using non-covalent interactions (NCI) visualization index.46 The NCI index, calculated for the fullerene–PMDI complex (Fig. 4b), reveals a large area of weak and attractive π–π contacts as a primary source of the interaction between these two molecules.
Conformational search for stable conformers of compound 2a produced three clusters of conformers (Fig. 5). In the first cluster (Fig. 5a) were the structures with both fullerene alkyl arms fully extended away from the PMDI part with no intramolecular π–π contacts. In the conformers of the second cluster (Fig. 5c), one fullerene arm is folded toward the PMDI part and in the third cluster both fullerene arms are folded towards PMDI part of the molecules (Fig. 5b). One representative structure of each cluster was additionally optimized and structures of extended, single-folded and double-folded conformers were obtained (Fig. 5a–c). The potential energies of all three conformers of compound 2a in toluene solution are evaluated at B3LYP/6-311G(d,p) level of theory. The most stabile conformation is single-folded (Fig. 5c), while double-folded (Fig. 5b) and extended (Fig. 5a) conformations are somewhat less stable (1.1 kcal mol−1 and 4.1 kcal mol−1, respectively).
It is worth noting that entropic factors are not included in our analysis. The entropy could, perhaps, significantly contribute to the stability of conformers, and especially in the solvent phase.
The second part of our computational investigation was focused on understanding the electronic properties of compound 2a. HOMO–LUMO energy gaps were calculated for the extended, single-folded (Fig. 5b) and double-folded (Fig. 5c) conformers. Obtained results were compared to the appropriate values for the reference C60 and PCBM. Summarized data are shown in Table 2. It can be seen that HOMO–LUMO energy gap is 0.34–0.52 eV lower for all conformers of compound 2a when compared to C60. Concurrently, both HOMO (0.29–0.45 eV) and LUMO energies (0.08–0.05 eV) are more negative for three conformers of 2a. On the other hand HOMO/LUMO energies and Egap of all 2a conformers more closely reassemble those of the PCBM acceptor (Table 2). All conformers of 2a display higher-lying LUMO and reduced bandgap which is very important in terms of improving the open-circuit voltage Voc and light harvesting capacity of the acceptor component in BHJ solar cells.53 Our further investigation of electron accepting features of compound 2a was oriented towards calculating the properties of excited state. TD-DFT calculations at B3LYP/6-311G(d,p) level of theory was employed in order to investigate the properties of first excited state. TD-DFT calculations predicts that first energy transitions will occur at energies of 1.881, 1.860 and 1.873 eV for extended, single folded and double folded conformer, respectively. The electron density difference maps between first excited and ground state are shown in Fig. 6. For the extended and single folded conformation there is no charge transfer between fullerene and PMDI parts of the molecule, indicating attenuated potential for the OPV n-phase material; excitation is localized on the fullerene cage (Fig. 6a and b). Due to favorable overlap of orbitals, a little of the electron transfer from fullerene to PMDI is observed in the double folded conformation (Fig. 6c).
Similarly, the Fukui function and Mulliken spin densities calculations (see ESI Fig. S3 and S4†) have shown that fullerene cage will be the most probable target of nucleophile attack. For radical monoanionic form of compound 2a, from 75% to 90% of additional electron spin density is located at the two fullerene cages.
To investigate the effect of the medium polarity on the aggregation of 2a–c, we used the following solvents: toluene, o-dichlorobenzene (ODCB) and chloroform as well as their 2
:
1 mixtures with 1,4-dioxane and iso-propanol. All samples were prepared in the following manner prior to SEM/TEM imaging. Compounds 2a–c were dissolved (1 mg) in 1 mL of solvent (∼0.5 μM) and a drop of the solution was deposited onto a glass-plate (SEM). Under toluene atmosphere, the solvent was allowed to evaporate during 48 h at a room temperature (SEM). For TEM measurement, 0.25 mM solution of 2a–c was deposited on a copper grid and subjected to drying under ambient conditions.
Different modes of assembly of 2a–c were observed by SEM and TEM measurements (Fig. 7 and 8). The apparent discrepancy resulted from different conditions used in the preparation of samples. The aggregation of molecules ought to be a function of the solid substrates on which it took place, the concentration of molecules and drying conditions (time, temperature, etc.). Compound 2a assembled into nanoparticles (Fig. 7a), while 2b and 2c (only 2c shown in Fig. 7b) showed a tendency to form unilamellar vesicles. The sizes of the vesicles were generally uniform with 200–300 nm in diameter. By a closer inspection of the single vesicle (HR TEM) we estimated the thickness of its membrane to be 4.4 nm. The width corresponds well to the length of the extended conformer of 2c (Fig. 7c). As presented in Fig. 8, compound 2a dissolved in toluene/dioxane (2
:
1) assembles into fiber containing sheets (or bundles) while in the same solvent mixture 2b aggregates into smaller hank-like objects (Fig. 8c). When processed from toluene/iPr-OH (2/1) 2a exhibits a flower-like structure (Fig. 8b), whereas 2b assembles into sheet fibers (Fig. 8d).
![]() | ||
| Fig. 8 SEM images of 2a prepared from (a) toluene/dioxane, (b) toluene/iPr-OH; 2b prepared from (c) toluene/dioxane, (d) toluene/iPr-OH and 2c prepared from (e) CHCl3/iPr-OH, (f) CHCl3/dioxane. | ||
When dissolved in CHCl3/iPr-OH (2
:
1) mixture compound 2c gave micron-size clusters (Fig. 8e), whereas in CHCl3/dioxane (2/1) mixture “necklace-like” chains (Fig. 8f). In other solvent combinations compounds 2a–c showed various self-organized architectures with the corresponding images presented in ESI (Fig. S5†).
To summarize, compounds 2a–c have a tendency to organize into nano- and micro-sized structures as a function of the length of their alkyl chains and experimental conditions. Importantly, the length of the extended conformer 2c bearing a hexyl chain is consistent with the thickness of the vesicle wall. This implies that the monolayer structure is in vesicles held by π–π (C60/C60, PMDI/PMDI) and van der Waals (CH2/CH2) interactions. With a large number of C60 exposed on the vesicle surface, “active cites” are created for fullerene–fullerene attraction between different vesicles, which could play a role in the observed hierarchical assembly by SEM.
We presume that compounds 2a and 2b, having shorter alkyl chains, give rise to weaker noncovalent contacts with a smaller solvophobic bias to lead to the formation of different structures, indicating a way for controlling the ordering of this type of supramolecular soft material.
Footnote |
| † Electronic supplementary information (ESI) available: Scheme S1; scanned NMR spectra of 2a–c; Tables S1 and S2; Fig. S1–S6. See DOI: 10.1039/c5ra16309a |
| This journal is © The Royal Society of Chemistry 2015 |