Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Boron difluorides with formazanate ligands: redox-switchable fluorescent dyes with large stokes shifts

M.-C. Chang a, A. Chantzis b, D. Jacquemin bc and E. Otten *a
aStratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands. E-mail: edwin.otten@rug.nl
bLaboratoire CEISAM – UMR CNRS 6230, Université de Nantes, 2 Rue de la Houssinière, BP 92208, 44322 Nantes Cedex 3, France
cInstitut Universitaire de France, 1, rue Descartes, 75005 Paris Cedex 5, France

Received 30th March 2016 , Accepted 10th May 2016

First published on 10th May 2016


Abstract

The synthesis of a series of (formazanate)boron difluorides and their 1-electron reduction products is described. The neutral compounds are fluorescent with large Stokes shifts. DFT calculations suggest that a large structural reorganization accompanies photoexictation and accounts for the large Stokes shift. Reduction of the neutral boron difluorides occurs at the ligand and generates the corresponding radical anions. These complexes are non-fluorescent, allowing switching of the emission by changing the ligand oxidation state.


Introduction

Fluorescent dyes have attracted increasing interest in various fields of modern research, including biological imaging,1 molecular probes,2 electroluminescent devices3 and photosensitizers.4 An ideal fluorescent dye presents high absorption coefficients and emission quantum yield, large Stokes shift, tuneable absorption/emission profiles, as well as high chemical and photochemical stability. Fluorescent dyes with large Stokes shift are more likely to show emission in the long wavelength region (>600 nm), which is desirable for applications in laser printing, information storage, displays and solar power conversion.5 Most importantly, fluorescent dyes emitting within the biological window (650–900 nm) are useful in bioimaging applications due to the small light scattering, low background emission and deep penetration into cells and tissues.6 Fluorescent dyes with small Stokes shifts usually suffer from self-quenching7 which limits the application in high concentration conditions and as solid state materials.

A large class of fluorescent dyes is based on the BF2 moiety bearing N,N′- or N,O-chelating ligands.8 Of these compounds, those with dipyrrin ligands, commonly known as BODIPYs (A in Chart 1),9 are the most popular systems due to their stability, high quantum yield and adaptable absorption/emission profiles. The optical properties of the BODIPYs are tuneable by changing the R1–R7 substituents of the ligand backbone.9,10 In addition, the F-substituents of the boron center can also be replaced by aryl, alkyl, and alkoxide groups.11


image file: c6dt01226d-c1.tif
Chart 1

Even though the largest documented Stokes shift for mono-BODIPYs is 185 nm,12 for most of these systems the Stokes shift is usually less than 40 nm. Recently, it was shown that incorporation of a triazole moiety on the R3 position of the BODIPY framework results in a substantial increase in Stokes shift (up to 160 nm).13 Also, chelates based on N,O donor atoms lead to dyes with relatively large Stokes shifts.8 In an alternative approach, the Stokes shift of fluorescent dyes can be increased by incorporation into an energy transfer fragment in which the energy absorbed by BODIPY unit is transferred to a second fluorophore.1a,4a A drawback, however, is that this strategy usually requires large synthetic efforts. In order to develop new fluorescent dyes, a variety of other N,N′-chelating ligands have been proposed as shown in Chart 1. Boron difluoride complexes bearing β-diketiminate ligands (B) exhibit strong absorption.14 In addition, compounds B show larger Stokes shifts (ca. 80 nm) than BODIPYs. A de-symmetrized β-diketiminate analog, the anilido-pyridine (C) framework, was reported by the groups of Piers and Heyne in 2011 and presents large Stokes shifts of 90–120 nm.15 Further modifications include 1,2-bis(pyrrolylmethylene) hydrazones (BOPHY) (D),16 and indigo-N,N′-diarylamines (E).17 Compounds D are highly fluorescent (Φ > 0.99) in solution with Stokes shifts around 40 nm.16 In addition, they show emission in thin films and as solid powders. Compounds E are a class of redox-active and near-infrared dyes that show Stokes shifts of 30–70 nm.17

Our group has been interested in the chemistry of formazanate ligands as nitrogen-rich, redox-active analogues of β-diketiminates,18,19 and we have previously communicated the synthesis of formazanate boron difluoride complexes.20 Concurrently, Gilroy and co-workers reported a series of similar compounds.21 In their work, the effect of ligand substituents on the optical properties was evaluated and the application of these compounds in electro-chemiluminescence22 and cell imaging23 was reported. Recently, Studer and co-workers used verdazyl radicals as “pro-fluorescent” radical probes.24

Here we describe the synthesis and characterization of (formazanate)boron complexes (LBF2, 1, Scheme 1) as well as the corresponding 1-electron reduction products, the radical anions [LBF2 (1˙), and discuss the impact of electronic and steric effects on both the optical and redox properties. We show that the neutral compounds 1 are fluorescent with large Stokes shifts, whereas in the radical anions 1˙ the fluorescence is quenched.

Results and discussion

Synthesis and characterisation of formazanate boron complexes

Two synthesis methods for formazanate boron difluoride complexes have been reported by us20 and the Gilroy group,21 shown as path A and path B in Scheme 1, respectively. In the present contribution, both methods were used to yield the desired products in moderate to good yields (60–90%). Using these methods, we prepared 9 (formazanate)boron difluoride compounds 1a–i. In addition, we synthesized two compounds in which the (formazanate)BF2 units are linked by phenylene moiety (1j and 1k). These compounds were recently reported by Gilroy and co-workers while this manuscript was in preparation.25 The formation of complexes 1 was confirmed by NMR spectroscopy, which shows diagnostic 1[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]1 triplets in the 11B NMR spectra for the 4-coordinate B centres (δ −0.7 to −2.3 ppm).19 In the 19F NMR spectra, 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 quartets between −145.6 and −159.4 ppm with JB–F = 20–30 Hz are observed that are consistent with the presence of a BF2 moiety.
image file: c6dt01226d-s1.tif
Scheme 1 Indirect (path A) and direct (path B) methods for the synthesis of (formazanate)boron difluorides (1).

Single crystals suitable for X-ray diffraction analysis were obtained by recrystallization from heptane or hexane (1a,b,d,h,i). The molecular structure of 1d as a representative example is shown in Fig. 1 whereas Table 1 lists selected geometrical parameters. The solid-state structures of all compounds show four-coordinate boron centres bound to the two terminal N atoms of a formazanate ligand to form a six-membered chelate ring, similar to those observed before.20,21 In most cases, the observed C–N and N–N bond lengths are in the range of 1.33–1.36 Å and 1.29–1.31 Å, respectively, consistent with a delocalized ligand backbone. For the derivative with electronically dissimilar N–Ar substituents (1d, R1 = Ph and R5 = C6F5) the data in Table 1 indicate a more localized bonding picture: the N–N bond length adjacent to the electron-withdrawing C6F5 group is significantly shorter (N1–N2: 1.299(2) Å) than that vicinal to the Ph substituent (N3–N4: 1.329(2) Å). All the boron centres in 1a, 1b, 1d, 1h and 1i show a distorted tetrahedral geometry, and the boron centre is significantly displaced from the formazanate backbone, with values ranging from 0.17–0.59 Å.


image file: c6dt01226d-f1.tif
Fig. 1 Molecular structure of 1d (left) and 1hBCF (right). Thermal ellipsoids are shown at 50% probability, and hydrogen atoms are removed for clarity.
Table 1 Selected bond lengths (Å) and bond angles (°) of LBF2 complexes (1) and their radical anions [1
  1a 1b 1d 1h 1i 1hBCF [Na(15-c-5)]+ [1c[thin space (1/6-em)]d [Cp2Co]+ [1f [Cp*2Co]2 [1j2−
a Displacement of the B atom from the least-squares plane defined by N1, N2, N3 and N4. b Angle between the planes defined by the N–Ar aromatic rings and the ligand backbone. c C10–N2 instead of C7–N2. d Values reported for one of the two independent molecules. e Data taken from ref. 20.
N1–N2 1.308(1) 1.3177(12) 1.2993(18) 1.294(1) 1.2931(15) 1.307(4) 1.376(3) 1.374(2) 1.3599(19)
N3–N4 1.308(1) 1.3289(17) 1.295(1) 1.300(4) 1.370(3) 1.356(2) 1.3620(18)
C7–N2 1.346(1) 1.3394(12) 1.361(2) 1.344(1) 1.3442(13)c 1.340(5) 1.337(4) 1.329(3) 1.335(2)
C7–N3 1.343(1) 1.329(2) 1.340(2) 1.339(5) 1.337(4) 1.333(2) 1.338(2)
N1–B1 1.559(2) 1.5523(15) 1.577(2) 1.573(2) 1.5665(16) 1.562(5) 1.520(4) 1.533(3) 1.541(2)
N4–B1 1.552(2) 1.548(2) 1.580(2) 1.553(5) 1.510(4) 1.512(3) 1.536(2)
N1–B1–N4 102.40(9) 100.15(11) 101.52(12) 105.87(9) 103.76(14) 104.4(3) 107.4(2) 108.49(17) 108.46(13)
B out-of-planea 0.50 0.59 0.51 0.17 0.29 0.49 0.38 0.08 0.05
Dihedral anglesb 47.98 47.97 47.54 24.16 88.11 41.53 27.54 12.64 4.01
42.00 50.09 22.35 44.39 65.05 83.44 13.23


The N-aryl substituents are rotated out of the plane of the ligand backbone with dihedral angles ranging from 23–50° for the compounds with unhindered Ar groups. The increased steric interactions in the N-Mes derivative 1i result, as expected, in a perpendicular orientation (dihedral angle of 88°).

Compounds with cyano substituents (1h,i) offer the possibility to influence the optical and redox properties by coordination of Lewis acids. To probe this, the B(C6F5)3-adduct 1hBCF was prepared by stirring 1h with B(C6F5)3 in toluene solution. Indicative of the formation of 1hBCF is the appearance of 3 additional resonances in the 19F NMR spectrum at −134, −155, and −163 ppm due to the C6F5 rings. An X-ray structure determination (molecular structure in Fig. 1, pertinent bond distances and angles in Table 1) confirms the coordination of the Lewis acidic B(C6F5)3 fragment to the cyano group.

UV-Vis absorption and emission spectroscopy

The optical properties of the formazanate boron complexes 1 were studied by UV-Vis absorption and emission spectroscopy in THF solution (Fig. 2 and Table 2). All compounds show medium to strong absorption bands with extinction coefficients between 9700 and 30[thin space (1/6-em)]200 L mol−1 cm−1 in the visible range of the spectrum (between 410 and 520 nm). The spectral trends observed here are very similar to those described for bis(formazanate)zinc complexes.19 Introducing an electron-donating tBu-group (1b) instead of a p-tolyl moiety (1a) results in a blueshift of the absorption maximum (517 and 473 nm for 1a and 1b, respectively).
image file: c6dt01226d-f2.tif
Fig. 2 UV-Vis absorption spectra (top) and normalised emission spectra (bottom) of 1c, 1f, and 1g. Data were collected in 10−5 M dry THF solution. The excitation wavelength of emission spectra is at 473 nm.
Table 2 Optical properties of formazan boron complexes 1 in THF solution
  λ max (nm) ε (M−1 cm−1) λ em[thin space (1/6-em)]c (nm) QY (%) SS (nm) SS (cm−1)
a Taken from ref. 21b. b Data collected in toluene solution. c Excitation wavelength = 473 nm.
1a 517 13[thin space (1/6-em)]736 641 0.2 124 3742
1b 473 13[thin space (1/6-em)]220 643 <0.1 170 5590
1c 464 21[thin space (1/6-em)]306 617 <0.1 153 5344
1d 482 12[thin space (1/6-em)]523 640 <0.1 158 5122
1e 460 12[thin space (1/6-em)]139 610 0.1 150 5346
1f 431 13[thin space (1/6-em)]702 612 <0.1 181 6862
1g 414 11[thin space (1/6-em)]852 650 <0.1 236 8770
1ha 489 25[thin space (1/6-em)]400 585 5 96 3356
1h , 502 30[thin space (1/6-em)]400 586 15 84 2855
1hBCF[thin space (1/6-em)]b 502 20[thin space (1/6-em)]582 632 0.5 130 4098
1i 428 9662 643 0.2 215 7812
1j 523 30[thin space (1/6-em)]227 664 0.6 141 4060
1k 508 16[thin space (1/6-em)]768 631 1.3 123 3837


In the case of the addition of an electron-withdrawing C6F5-group (1f, 1g) at the R3 position instead of a p-tolyl moiety (1c, 1e), the absorption maximum undergoes a blueshift as well (431 nm for 1fvs. 464 nm for 1c; 414 nm for 1gvs. 460 m for 1e). Upon substitution of phenyl for mestiyl at the R1 or R5 positions, the λmax is again blueshifted (489 nm for 1hvs. 428 nm for 1i). This likely results from the nearly perpendicular orientation of the mesityl group, which doesn't allow conjugation between the formazanate core and the side mestiyl ring, thereby limiting the size of the π-conjugated system.21c The introduction of the Lewis acid B(C6F5)3 in 1hBCF surprisingly does not influence λmax. The UV-Vis absorption spectra of the di-formazanate system (1j and 1k) are very similar to 1a except for the expected ca. two-fold increase in molar extinction coefficients.

Similar to other reported boron difluoride complexes, compounds 1 were shown to be emissive, with emission wavelengths λem in the 580 nm to 670 nm range in THF solution (Fig. 2). The reported (formazanate)boron difluoride complexes usually show large Stokes shifts (100–150 nm).21,23 For most of the compounds reported here, the Stokes shifts are even larger (>150 nm). As an example, 1g shows λmax at 414 nm and λem at 650 nm, a Stokes shift of 236 nm (8770 cm−1), which to the best of our knowledge is the highest value reported to date for this class of compounds. The large Stokes shift of compounds 1c–g might be due to the asymmetry in the substitution pattern of the ligands. The B(C6F5)3 adduct 1hBCF shows a redshift in its emission spectrum (λmax = 632 nm) relative to 1h (586 nm), which increases the Stokes shift from 84 nm (4098 cm−1) to 130 nm (4098 cm−1) in toluene solution. These results suggest that the emission profile of (3-cyanoformazanate)boron difluoride complexes is tuneable via binding of Lewis acids. The phenylene-linked systems 1j and 1k show emission spectra that are redshifted and blueshifted, respectively, compared to those of the parent compound 1a, whereas the obtained quantum yields are higher.

Cyclic voltammetry

The redox chemistry of compounds 1 was studied by cyclic voltammetry (CV) in tetrahydrofuran (THF) or 1,2-dichloroethane (DCE) solution under nitrogen atmosphere. All complexes show two (quasi)reversible redox processes corresponding to the formation of the radical anions [1]˙− and dianions [1]2−. The results of cyclic voltammetry studies are summarized in Table 3. From these data it is apparent that the redox potentials of (formazanate)BF2 complexes can be altered over a wide range (up to 530 mV) by changing the substituents on the ligand framework. The measured redox-potentials correlate with the electron-donating ability of the R3-substituent via inductive effects: an electron-donating t-butyl group (1b) induces a shift to more negative potential, while electron-withdrawing cyano (1h) or C6F5 substituents (1f) result in an anodic shift, in comparison to R3 = p-tolyl (1a). Changing the nature of the aromatic N-substituent(s) from Ph to an electron-withdrawing C6F5 group (1avs.1d, 1cvs.1e and 1fvs.1g) shifts the redox potentials in the positive direction by 130–180 mV. As expected, substitution of a phenyl group for an electron-donating mesityl group (1avs.1c, 1dvs.1e and 1hvs.1i) will shift the measured redox potentials to more negative values by around 240 mV. Surprisingly, coordination of the Lewis acid B(C6F5)3 in 1hBCF induces only a marginal variation of the first redox potential from −0.66 V (1h) to −0.67 V (1hBCF). This might be related to the high concentration of electrolyte: the B(C6F5)3 group in 1hBCF does not interact strongly in the high ionic strength solvent system used for collecting the CV data. The LBF20/−1 redox couple in 1j and 1k is observed at potentials similar to those in 1a, but corresponds to two sequential one-electron transfers as evidenced by the somewhat larger peak-to-peak separation in 1j and 1k (0.268–0.287 V, see Fig. 3) than in 1a (0.222 V). At more negative potential, 1j and 1k show two very close one-electron processes (between −2.0 and −2.4 V vs. Fc/Fc+). In the case of 1j, these are poorly resolved but for 1k there are clearly two separate events, indicating that the two (formazanate)BF2 units in these phenylene-linked systems do not behave fully independently.26 Gilroy and co-workers described similar electrochemical data for these compounds in dichloromethane solution. While the 2-electron reduction products appeared unstable in dichloromethane solution,25 our data in THF suggest that in this solvent the products are stable and might be isolable (vide infra).
image file: c6dt01226d-f3.tif
Fig. 3 Cyclic voltammograms of 1j and 1k recorded at 0.1 V s−1 in 1.5 mM THF solution containing 0.1 M tetrabutylammonium hexafluorophosphate.
Table 3 Electrochemical Data (V vs. Fc/Fc+) of LBF2 complexesa
  LBF20/−1 LBF2−1/−2
a Cyclic voltammetry experiments collected in THF solution (1.5 mM 1 and 0.1 M [Bu4N][PF6] as supporting electrolyte) at a scan rate if 0.1 V s−1. The data were referenced internally against the Fc/Fc+ couple. b Collected in DCE solution.
1a −0.98 −2.06
1b −1.08 −2.21
1c −1.19 −2.34
1d −0.85 −1.99
1e −1.01 −2.26
1f −1.02 −2.25
1g −0.84 −2.17
1h −0.66 −1.83
1h −0.65 −1.76
1hBCF[thin space (1/6-em)]b −0.67 −1.75
1i −0.90 −2.41
1j −0.95 −2.0 to −2.3
1k −0.94 −2.0 to −2.4


Synthesis and characterisation of radical anions [1

According to the CV data, the radical anions [1 are relatively stable and could be accessible chemically. Indeed, treatment of compounds 1 with Cp2Co precipitated the dark green salts [Cp2Co]+[1. For the phenylene-linked bis(formazanate) systems 1j and 1k, treatment with 1 equiv. of reducing agent resulted in precipitation of 0.5 equiv. of the corresponding diradical dianions [1j,k2− and 0.5 equiv. of unreacted starting material instead of the radical anions [1j,k. This is presumably driven by the poor solubility of the dianionic products. In order to obtain crystalline products, a variety of reducing agents was used; a representative series ([1a,20 [1c, [1f and [1j2−) could be obtained in high yield with either Cp2Co+ ([1a and [1f), Na(15-crown-5)+ ([1c) or Cp*2Co+ ([1j2−) as countercations. For all compounds, significant changes were observed in the metrical parameters of the reduced compounds in comparison to the neutral precursors. The most obvious change is the elongation of the N–N bonds (from 1.30–1.31 Å in 1 to 1.34–1.37 Å in [1) due to population of a π*-orbital that has N–N antibonding character.19a,20 The B–N bond lengths decrease upon reduction (1a: 1.559(2)/1.552(2) Å vs. [1a: 1.532(3)/1.536(3) Å), as anticipated. In addition to the changes of the N–N and B–N bond lengths, the formation of [1 is accompanied by a substantial planarization of the system: the B atom is virtually in the plane of the ligand (displacement <0.09 Å for both [1a and [1f) and the unhindered N-Ph substituents become coplanar with the formazanate backbone. The slightly larger boron displacement in [1c (0.376 Å) is due to an asymmetric interaction between the Na+ cation and the F atoms (2.383(4) and 2.419(4) Å) (Fig. 4).
image file: c6dt01226d-f4.tif
Fig. 4 Molecular structures of [Na(15-crown-5)]+[1c (left) and [Cp*2Co]+2[1j2− (right) showing 50% probability ellipsoids; cations in the latter omitted for clarity.

The metrical parameters of each of the (formazanate)boron fragments in the para-phenylene linked compound [1j2− are similar to that in [1a. These data suggest that [1j2− is best represented as two nearly-independent [LBF2 units, and a significant quinoidal contribution to the structure of the phenylene linker can be excluded.27

EPR spectroscopy for the radical anions [1 shows broad, featureless resonances at g ∼ 2, indicative of ligand-based radicals. While related organic verdazyl radicals show hyperfine interactions with the N nuclei,24,28 these are absent in the boron systems studied here. DFT calculations (see the ESI) suggest that the hyperfine coupling constants in [1a are indeed smaller for the boron compounds reported here than those in organic analogues. Additionally, hyperfine interactions with the boron/fluorine nuclei are present in the compounds [1a, which accounts for the broad signal with unresolved hyperfine interactions. The phenylene-linked systems [1j2− and [1k2− are diradicals similar to linked bis(verdazyl) radicals that have been reported,26,29 but in contrast to these purely organic systems our boron-based diradicals do not show zero-field splitting in their EPR spectra. This implies that the spin centres behave independently, effectively leading to S = 1/2 behaviour without intramolecular electronic coupling (J ∼ 0).30 EPR spectra were collected in MeTHF/DMSO glass in the range between 5 and 40 K to evaluate the spin ground state of these compounds (Fig. S1 in the ESI). Even though the absence of zero-field splitting indicates little or no electronic coupling, for the para-phenylene linker the signal intensity decreases upon cooling to 5 K. This suggests the two radical centres to be anti-ferromagnetically coupled (J < 0) leading to a singlet biradical ground-state. For the meta-phenylene linked compound [1k2−, the signal broadens upon cooling but its intensity does not change significantly and a conclusive assignment of the electronic ground-state cannot be made. DFT calculations for both systems (see the ESI) agree with the VT EPR data that [1j2− has a singlet biradical ground state. On the other hand, [1k2− is calculated to have a triplet ground state, which is in line with meta-phenylene linked bis(verdazyl) radicals reported in the literature.26

Optical properties of the radical anions [1

UV/Vis absorption spectroscopy of compounds [1 shows two absorption bands in the visible range of the spectrum (see Table S3 in the ESI). One of these is shifted to lower energy (up to 716 nm for [1a, Fig. 5) in comparison to the neutral precursors and is indicative of the presence of a formazanate ligand in the radical dianionic form (L˙2−).18–20 We were interested to evaluate the influence of the redox-state of the compounds on their emission properties, and thus emission spectra of [1a were collected. Regardless of the excitation wavelength, the compound was only very weakly emissive and the observed spectrum indicates that λem is identical to that of the neutral precursor 1a. Based on this, we conclude that a small amount of decomposition to 1a is responsible for the observed spectrum, and the radical anions [1 are non-emissive.
image file: c6dt01226d-f5.tif
Fig. 5 UV-Vis absorption spectrum of compounds 1a (dotted red line) and [1a (green line) in THF solution.

Theoretical calculations

We have performed calculations in which the ground- and excited-states structures were respectively determined with DFT and Time-Dependent DFT (TD-DFT), whereas the transition energies were obtained with both TD-DFT and SOS-CIS(D) model, corrected for solvent effects (see the ESI for details). This approach was shown previously to be adequate for fluoroborate complexes.31 In order to demonstrate the ability of this selected theoretical level in reproducing the structure, properties and electronic absorption spectra of formazanate boron complexes, systems 1a and [Cp2Co]+[1a have been chosen as representative examples. Inspection of the theoretically determined structure of 1a, shows that the B atom is displaced with respect to the formazanate backbone, as indicated by the dihedral angle (B1,N1,N2,N3) of 16.29°. Moreover, the DFT-computed C–N, N–N and B–N bond lengths of the neutral 1a system are 1.340 Å, 1.286 Å and 1.562 Å, respectively, showing electron delocalization along the chelating formazanate ligand. These values match well their experimental counterparts, though theory undershoots the N–N bond length (Table 1). We have determined the optimal structure of the lowest singlet excited-state as well. It shows a nearly perfectly planar core with a (B1,N1,N2,N3) dihedral of 0.03°. In the excited-state, the computed C–N, N–N and B–N bond lengths attain 1.345 Å, 1.320 Å and 1.567 Å, respectively, indicating a strong elongation of the N–N bonds. These significant geometrical changes are consistent with the sizeable Stokes shifts that have been measured experimentally (see Table 1). Passing to the theoretically determined ground-state structure of the [1a radical anion, large structural changes are also observed, in agreement with the experimental trends. Theory correctly predicts a significant planarization of the boron atom with respect to the plane of the formazanate ligand, resulting in a dihedral angle (B1,N1,N2,N3) value of 4.93°, much smaller than in 1a. Moreover, the experimentally observed elongation of the N–N bond and the simultaneous shortening of the B–N bond upon passing from neutral to radical anionic forms is also qualitatively reproduced by the theoretical calculations. Indeed, DFT foresees that the former goes from 1.286 Å to 1.343 Å and the latter from 1.562 Å to 1.543 Å. To investigate the importance of the counterion, we have also optimized the 1[thin space (1/6-em)]:[thin space (1/6-em)]1 [Cp2Co]+[1a complex, and obtained similar trends but quite different values, i.e., the dihedral is 9.34°, the N–N bond lengths are 1.342 Å and 1.347 Å, whereas the B–N distances are 1.539 Å and 1.541 Å. It has to be noted that while in the solid-state it is in principle possible to observe changes in the planarity of the two phenyl substituents with respect to the central ring upon reduction of 1a, in the theoretical liquid-phase calculations the essentially free rotation of the phenyl rings makes analogous comparisons less relevant. Clearly, DFT allows a reasonable determination of the ground-state equilibrium geometries of formazanate boron complexes.

In order to access the predictive capability of our model in reproducing the electronic absorption spectra of the neutral formazanate boron complexes, we have determined the vertical transition energies of all 11 1x [x = a,b…] compounds in Table 1 and the obtained results are listed in Table 4. We note that the chosen theoretical level, SOS-CIS(D), can nicely reproduce both the absolute experimental values and the trends upon substitution whereas TD-DFT overestimates the transition energies, as expected.31 For the full set of compounds, the mean signed error is −4.7 nm and the mean absolute error is 15.1 nm when SOS-CIS(D) is applied. Clearly, these values are rather small. At the same time, the linear correlation coefficient relating theoretical and experimental values is large (0.97).

Table 4 Experimental and theoretical wavelengths of maximum intensity for the absorption bands of systems [1x]. We show both the TD-DFT values and their SOS-CIS(D) counterparts
  λ TD-DFTmax (nm) λ SOS-CIS(D)max (nm) λ Expmax (nm)
1a 463 543 517
1b 418 485 473
1c 417 462 464
1d 445 501 482
1e 420 451 460
1f 400 434 431
1g 407 400 414
1h 414 486 502
1i 373 412 428
1j 469 545 523
1k 463 535 508


Let us now turn to the radical anions [1a, the experimentally recorded UV-Vis spectra of both 1a and its radical anionic form in THF are shown in Fig. 5. One clearly notes the emergence of two bands in the visible range for the latter structure. Indeed, using TD-DFT, we computed two intense bands at 595 nm and 399 nm for [1a which nicely surround the 463 nm values computed for 1a at the same level of theory. This fits well with the experimental trends. Indeed, the experiment reveals that the lowest-energy band of the neutral precursor 1a is significantly redshifted by ca. 198 nm upon passing to the radical anion, whereas theory underestimates the redshift (132 nm). For the second band, the experimental blueshift compared to the main band of 1a is ca. −59 nm, and theory provides −64 nm. Test calculations performed with the [Cp2Co]+[1a complex demonstrated that the counter-ion has a rather marginal impact on these variations.

In Fig. 6 the occupied-virtual molecular orbital (MO) pairs that characterise the most intense theoretically determined electronic transitions for both 1a and 1a˙ are shown. For 1a the lowest state can be mainly ascribed to a HOMO–LUMO transition, as expected. The pictures of the electron density redistribution that accompanies photoexcitation show that reduction of the neutral 1a compound does not change the nature of the transition (of HOMO to SOMO nature in 1a˙) involved in the lowest-energy band of the absorption spectrum but merely shifts its position, in agreement with the experimental suggestion. Indeed, for both 1a and 1a˙ the lowest-energy transition is characterized by withdrawal of electron density from the p-tolyl group and its redistribution on the rest of the system. In contrast, the long wavelength band of 1a˙ is characterized by electron density withdrawal from the formazanate-containing ring and two phenyls to the p-tolyl group.


image file: c6dt01226d-f6.tif
Fig. 6 Occupied-virtual molecular orbital (MO) pairs that characterize the most intense theoretically determined electronic transitions for compounds 1a and 1a˙. The numbering of the bands is based on the experimentally determined absorption spectra for both systems. Contour threshold: 0.03 a.u.

Conclusions

We have prepared a series of (formazanate)boron difluoride complexes and evaluated their structural, electrochemical and luminescent properties. The neutral compounds (1) show an intense absorption band in the visible range, which TD-DFT calculations confirm to be due to electronic transitions (π → π*) within the ligand. Evaluation of their luminescence properties shows large Stokes shifts across the series, albeit with low quantum yields. In case of the cyano-substituted formazans, the emission wavelength may be further increased by coordination of a Lewis acid (as shown for 1hBCF). For a representative series of compounds, the 1-electron reduction products (1˙) were characterised, which show two absorption maxima in the visible that occur at both higher and lower energy than in the neutral precursor. Theoretical calculations indicate that, similar to the experimentally verified structures of 1-electron reduction products, photochemical excitation to the lowest singlet excited state results in large structural reorganization/planarization following the partial population of the N–N π*-orbitals. Due to the non-emissive nature of the radical anions 1˙, the formazanate boron difluoride compounds presented here allow for redox-switching of their luminescent properties, and may find application in sensing and imaging applications.

Acknowledgements

E. O. is grateful to the Netherlands Organisation for Scientific Research (NWO) for a Veni grant. D. J. acknowledges the European Research Council (ERC) for financial support in the framework of a starting grant (Marches-278845). A. C. thanks the ERC (Marches-278845) for his postdoctoral grant. This research used resources of the GENCI-CINES/IDRIS, of the CCIPL, and of a local Troy cluster. We thank Prof. Wesley Browne for access to spectroscopic facilities and help with quantum yield determinations and Prof. Bas de Bruin for VT-EPR data.

Notes and references

  1. (a) A. Martin, R. D. Moriarty, C. Long, R. J. Forster and T. E. Keyes, Asian J. Org. Chem., 2013, 2, 763–778 CrossRef CAS; (b) K. Sreenath, Z. Yuan, J. R. Allen, M. W. Davidson and L. Zhu, Chem. – Eur. J., 2015, 21, 867–874 CrossRef PubMed; (c) A. Yamada, Y. Hiruta, J. Wang, E. Ayano and H. Kanazawa, Biomacromolecules, 2015, 16, 2356–2362 CrossRef CAS PubMed; (d) P. Agarwal, B. J. Beahm, P. Shieh and C. R. Bertozzi, Angew. Chem., Int. Ed., 2015, 54, 11504–11510 CrossRef CAS PubMed.
  2. (a) T. W. Hudnall and F. P. Gabbai, Chem. Commun., 2008, 4596–4597 RSC; (b) M. V. Kuperman, S. V. Chernii, M. Y. Losytskyy, D. V. Kryvorotenko, N. O. Derevyanko, Y. L. Slominskii, V. B. Kovalska and S. M. Yarmoluk, Anal. Biochem., 2015, 484, 9–17 CrossRef CAS PubMed.
  3. (a) A. M. Prokhorov, T. Hofbeck, R. Czerwieniec, A. F. Suleymanova, D. N. Kozhevnikov and H. Yersin, J. Am. Chem. Soc., 2014, 136, 9637–9642 CrossRef CAS PubMed; (b) X. Cao, J. Miao, M. Zhu, C. Zhong, C. Yang, H. Wu, J. Qin and Y. Cao, Chem. Mater., 2015, 27, 96–104 CrossRef CAS; (c) L. Yao, S. Zhang, R. Wang, W. Li, F. Shen, B. Yang and Y. Ma, Angew. Chem., Int. Ed., 2014, 53, 2119–2123 CrossRef CAS PubMed.
  4. (a) S. D. Topel, G. T. Cin and E. U. Akkaya, Chem. Commun., 2014, 50, 8896–8899 RSC; (b) T. Yogo, Y. Urano, Y. Ishitsuka, F. Maniwa and T. Nagano, J. Am. Chem. Soc., 2005, 127, 12162–12163 CrossRef CAS PubMed.
  5. (a) S. Erten-Ela, M. D. Yilmaz, B. Icli, Y. Dede, S. Icli and E. U. Akkaya, Org. Lett., 2008, 10, 3299–3302 CrossRef CAS PubMed; (b) A. Mishra, M. K. R. Fischer and P. Bäuerle, Angew. Chem., Int. Ed., 2009, 48, 2474–2499 CrossRef CAS PubMed.
  6. V. Ntziachristos, J. Ripoll, L. V. Wang and R. Weissleder, Nat. Biotechnol., 2005, 23, 313–320 CrossRef CAS PubMed.
  7. (a) S. J. Farley, D. L. Rochester, A. L. Thompson, J. A. K. Howard and J. A. G. Williams, Inorg. Chem., 2005, 44, 9690–9703 CrossRef CAS PubMed; (b) W. Qin, M. Baruah, M. Van der Auweraer, F. C. De Schryver and N. Boens, J. Phys. Chem. A, 2005, 109, 7371–7384 CrossRef CAS PubMed.
  8. D. Frath, J. Massue, G. Ulrich and R. Ziessel, Angew. Chem., Int. Ed., 2014, 53, 2290–2310 CrossRef CAS PubMed.
  9. (a) A. Loudet and K. Burgess, Chem. Rev., 2007, 107, 4891–4932 CrossRef CAS PubMed; (b) G. Ulrich, R. Ziessel and A. Harriman, Angew. Chem., Int. Ed., 2008, 47, 1184–1201 CrossRef CAS PubMed; (c) N. Boens, V. Leen and W. Dehaen, Chem. Soc. Rev., 2012, 41, 1130–1172 RSC.
  10. (a) J. Rosenthal and S. J. Lippard, J. Am. Chem. Soc., 2010, 132, 5536–5537 CrossRef CAS PubMed; (b) J.-J. Shie, Y.-C. Liu, Y.-M. Lee, C. Lim, J.-M. Fang and C.-H. Wong, J. Am. Chem. Soc., 2014, 136, 9953–9961 CrossRef CAS PubMed; (c) C. Cheng, N. Gao, C. Yu, Z. Wang, J. Wang, E. Hao, Y. Wei, X. Mu, Y. Tian, C. Ran and L. Jiao, Org. Lett., 2015, 17, 278–281 CrossRef CAS PubMed.
  11. (a) T. Lundrigan, S. M. Crawford, T. S. Cameron and A. Thompson, Chem. Commun., 2012, 48, 1003–1005 RSC; (b) G. Duran-Sampedro, I. Esnal, A. R. Agarrabeitia, J. Bañuelos Prieto, L. Cerdán, I. García-Moreno, A. Costela, I. Lopez-Arbeloa and M. J. Ortiz, Chem. – Eur. J., 2014, 20, 2646–2653 CrossRef CAS PubMed; (c) T. Lundrigan, T. S. Cameron and A. Thompson, Chem. Commun., 2014, 50, 7028–7031 RSC; (d) A. B. More, S. Mula, S. Thakare, N. Sekar, A. K. Ray and S. Chattopadhyay, J. Org. Chem., 2014, 79, 10981–10987 CrossRef CAS PubMed.
  12. A. Martin, C. Long, R. J. Forster and T. E. Keyes, Chem. Commun., 2012, 48, 5617–5619 RSC.
  13. J. C. Er, M. K. Tang, C. G. Chia, H. Liew, M. Vendrell and Y.-T. Chang, Chem. Sci., 2013, 4, 2168–2176 RSC.
  14. (a) F. P. Macedo, C. Gwengo, S. V. Lindeman, M. D. Smith and J. R. Gardinier, Eur. J. Inorg. Chem., 2008, 3200–3211 CrossRef CAS; (b) R. Yoshii, A. Hirose, K. Tanaka and Y. Chujo, J. Am. Chem. Soc., 2014, 136, 18131–18139 CrossRef CAS PubMed; (c) R. Yoshii, A. Hirose, K. Tanaka and Y. Chujo, Chem. – Eur. J., 2014, 20, 8320–8324 CrossRef CAS PubMed; (d) S. M. Barbon, V. N. Staroverov, P. D. Boyle and J. B. Gilroy, Dalton Trans., 2014, 43, 240–250 RSC.
  15. J. F. Araneda, W. E. Piers, B. Heyne, M. Parvez and R. McDonald, Angew. Chem., Int. Ed., 2011, 50, 12214–12217 CrossRef CAS PubMed.
  16. (a) C. Yu, L. Jiao, P. Zhang, Z. Feng, C. Cheng, Y. Wei, X. Mu and E. Hao, Org. Lett., 2014, 16, 3048–3051 CrossRef CAS PubMed; (b) I.-S. Tamgho, A. Hasheminasab, J. T. Engle, V. N. Nemykin and C. J. Ziegler, J. Am. Chem. Soc., 2014, 136, 5623–5626 CrossRef CAS PubMed; (c) Q. Huaulmé, A. Mirloup, P. Retailleau and R. Ziessel, Org. Lett., 2015, 17, 2246–2249 CrossRef PubMed.
  17. (a) G. Nawn, R. McDonald and R. G. Hicks, Inorg. Chem., 2013, 52, 10912–10919 CrossRef CAS PubMed; (b) G. Nawn, S. R. Oakley, M. B. Majewski, R. McDonald, B. O. Patrick and R. G. Hicks, Chem. Sci., 2013, 4, 612–621 RSC.
  18. J. B. Gilroy, M. J. Ferguson, R. McDonald, B. O. Patrick and R. G. Hicks, Chem. Commun., 2007, 126–128 RSC.
  19. (a) M.-C. Chang, T. Dann, D. P. Day, M. Lutz, G. G. Wildgoose and E. Otten, Angew. Chem., Int. Ed., 2014, 53, 4118–4122 CrossRef CAS PubMed; (b) M.-C. Chang, P. Roewen, R. Travieso-Puente, M. Lutz and E. Otten, Inorg. Chem., 2015, 54, 379–388 CrossRef CAS PubMed.
  20. M. C. Chang and E. Otten, Chem. Commun., 2014, 50, 7431–7433 RSC.
  21. (a) S. M. Barbon, J. T. Price, P. A. Reinkeluers and J. B. Gilroy, Inorg. Chem., 2014, 53, 10585–10593 CrossRef CAS PubMed; (b) S. M. Barbon, P. A. Reinkeluers, J. T. Price, V. N. Staroverov and J. B. Gilroy, Chem. – Eur. J., 2014, 20, 11340–11344 CrossRef CAS PubMed; (c) S. M. Barbon, V. N. Staroverov and J. B. Gilroy, J. Org. Chem., 2015, 80, 5226–5235 CrossRef CAS PubMed.
  22. M. Hesari, S. M. Barbon, V. N. Staroverov, Z. Ding and J. B. Gilroy, Chem. Commun., 2015, 51, 3766–3769 RSC.
  23. R. R. Maar, S. M. Barbon, N. Sharma, H. Groom, L. G. Luyt and J. B. Gilroy, Chem. – Eur. J., 2015, 21, 15589–15599 CrossRef CAS PubMed.
  24. D. Matuschek, S. Eusterwiemann, L. Stegemann, C. Doerenkamp, B. Wibbeling, C. G. Daniliuc, N. L. Doltsinis, C. A. Strassert, H. Eckert and A. Studer, Chem. Sci., 2015, 6, 4712–4716 RSC.
  25. S. M. Barbon, J. T. Price, U. Yogarajah and J. B. Gilroy, RSC Adv., 2015, 5, 56316–56324 RSC.
  26. J. B. Gilroy, S. D. J. McKinnon, P. Kennepohl, M. S. Zsombor, M. J. Ferguson, L. K. Thompson and R. G. Hicks, J. Org. Chem., 2007, 72, 8062–8069 CrossRef CAS PubMed.
  27. L. K. Montgomery, J. C. Huffman, E. A. Jurczak and M. P. Grendze, J. Am. Chem. Soc., 1986, 108, 6004–6011 CrossRef CAS PubMed.
  28. R. G. Hicks, in Stable Radicals, John Wiley & Sons, Ltd, 2010, pp. 245–279 Search PubMed.
  29. (a) K. Mukai, N. Azuma, H. Shikata and K. Ishizu, Bull. Chem. Soc. Jpn., 1970, 43, 3958–3960 CrossRef CAS; (b) R. M. Fico, M. F. Hay, S. Reese, S. Hammond, E. Lambert and M. A. Fox, J. Org. Chem., 1999, 64, 9386–9392 CrossRef CAS.
  30. M. Abe, Chem. Rev., 2013, 113, 7011–7088 CrossRef CAS PubMed.
  31. B. Le Guennic and D. Jacquemin, Acc. Chem. Res., 2015, 48, 530–537 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: Experimental section, X-ray data, computational methods. CCDC 1471287–1471289, 1471291–1471293 and 1471295. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6dt01226d

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.