Modulating Mechanochromic Luminescence Quenching of Alkylated

Jun 24, 2016 - Zhang , G.; Evans , R. E.; Campbell , K. A.; Fraser , C. L. Role of Boron in ... Zojer, Egbert; Barlow, Stephen; Bredas, Jean-Luc; Perr...
0 downloads 0 Views 5MB Size
Article pubs.acs.org/JPCC

Modulating Mechanochromic Luminescence Quenching of Alkylated Iodo Difluoroboron Dibenzoylmethane Materials William A. Morris,† Michal Sabat,†,‡ Tristan Butler,† Christopher A. DeRosa,† and Cassandra L. Fraser*,† †

Department of Chemistry, University of Virginia, McCormick Road, Charlottesville, Virginia 22904, United States Department of Materials Science and Engineering, University of Virginia, 395 McCormick Road, Charlottesville, Virginia 22904, United States



S Supporting Information *

ABSTRACT: Difluoroboron β-diketonate compounds exhibit solid-state luminescence phenomena. Among these are reversible mechanochromic luminescence (ML), aggregation induced emission (AIE), and mechanochromic luminescence quenching (MLQ). These properties can be tuned by alterations to the molecular structure. Dyes with varying halide substituents exhibit tunable ML, MLQ, and solid-state emission with high quantum yields. A series of difluoroboron dibenzoylmethane (BF2dbm) dyes with iodide and alkoxyl substituents (BF2dbm(I)OR) were synthesized where R = CH3 (C1), C5H11 (C5), C6H13 (C6), C12H25 (C12), and C18H37 (C18)). The 4-iodo parent compound BF2dbm(I) (H) was made for comparison. By keeping the heavy atom static, the dependence of ML properties on alkyl chain length was probed. The hydrogen derivative is only weakly emissive as a solid and exhibited minimal mechanoresponsive behavior. In contrast, alkoxy dyes exhibited tunable ML and MLQ properties depending on the length of the alkyl chain. Longer chain dyes corresponded to smaller singlet triplet energy gaps, greater triplet emission enhancement (77 K), and longer recovery times after smearing under ambient conditions. Shorter chain dyes have a much greater affinity for the ordered emissive state, as confirmed by atomic force microscopy (AFM). Powder X-ray diffraction (XRD) was performed on pristine dye powders as well as drop-cast films to gauge crystallinity in various forms. Single crystal XRD analysis of the H and C5 dyes revealed significant differences in crystal packing and π−π stacked dimers for dyes bearing alkyl chains and dyes without alkyl chains. Unique I−I and F−π close interactions were discovered in the dye crystals.



INTRODUCTION Organic stimuli-responsive materials have many potential applications as sensors, for memory storage, security inks, OLED development, criminal justice, and even art and design.1−8 In particular, organoboron materials are a research area of much interest given their tunable optical properties.9−15 Boron coordination induces striking luminescent properties in β-diketonates.16 Difluoroboron β-diketonates (BF2bdks) frequently display two-photon absorption cross sections,17,18 high quantum yields,18−20 large extinction coefficients,18−20 tunable absorption in the near-UV range,18,19 intramolecular charge transfer (ICT) character,20,21 a range of emission colors in the solid-state,19,22 oxygen-sensitive room temperature phosphorescence in rigid media such as polymers,23−25 organic vapor sensitivity,26,27 and reversible mechanochromic luminescence (ML).22,28−32 ML materials are so named because they undergo a change in emission color in response to mechanical perturbation.1−5 Types of ML materials include inorganic and organic molecular solids,33,34 liquid crystals,35,36 and polymers,3,37,38 to name a few. In recent years, many groups have made progress in synthesizing, understanding, and applying ML materials. The first ML report from our group, difluoroboron avobenzone (BF2AVB),28 exhibited high contrast green to yellow ML with spontaneous recovery of the green emissive © XXXX American Chemical Society

state at room temperature or more rapid recovery with heating. AFM images revealed that amorphous spin-cast films become more ordered upon thermal annealing. Crystals of BF2AVB subjected to single crystal X-ray diffraction (XRD) analysis revealed distinctly emissive polymorphs whose emissions were dependent upon crystal packing. Zhang et al. have postulated that smearing BF2bdk films forms ground-state and excimeric aggregate species, namely Haggregates, in which the dye molecules adopt a face-to-face arrangement with red-shifted fluorescence.39,40 Building upon seminal findings with BF2(avobenzone),28 Sket et al. recently reported a BF2bdk system (difluoroboron 1-phenyl-3-(3,5dimethoxyphenyl)-propane-1,3-dione) that also exhibits polymorphism and ML. The compound formed two distinct emissive polymorphs differing in the orientation of the methoxy groups. One mode of packing exhibited distinct ML behavior as well as crystallization induced emission enhancement (CIEE). Furthermore, the ML was reversible either by applying CH2Cl2 dropwise, heating, or allowing the sample to sit for a period of time at room temperature. They also found it was possible to Received: March 30, 2016 Revised: June 3, 2016

A

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C switch between the two polymorphs by heating and evaporation from various solvents.41 Furthermore, Chujo et al. have demonstrated many examples of organoboron compounds with highly tunable emission properties.10,42,43 Among them is the recent discovery of boron ketoiminates exhibiting aggregation induced emission (AIE) as well as ML. They found that the reversible ML properties could be tuned by altering end functional groups. Substitution with halogens afforded much more dramatic ML than dyes without such substituents. This was attributed to intermolecular interactions, such as halogen−halogen bonding, not present in non-halogensubstituted dyes.44 Though ML materials have potential for practical applications, logical design of these materials remains challenging.1,2 Prior work in our group has demonstrated that reversible ML properties of BF2bdk dyes bearing alkyl chain substituents are highly dependent on the length of the chain, and thus, their properties can be tuned by varying chain length.29 The iodine heavy atom substituent in BF2dbm(I)OC12H25 can also facilitate mechanically induced luminescence quenching (MLQ) in annealed films by enhancing crossover to the triplet state, which is quenched by oxygen and other nonradiative decay pathways at room temperature under air. Total emission and delayed emission spectra recorded at 77 K submerged in liquid N2 revealed that phosphorescence intensity increased substantially for the annealed films after smearing, presumably due to the creation of aggregate species with singlet excited state energies closer in energy to that of the aggregate triplet state.32 A more recent study probed how varying the halide atom affected this phenomenon. 30 Changing a single substituent, from F to Cl, Br, or I, can alter film morphology, solid-state fluorescence wavelength, phosphorescence intensity, and the ability of the materials to recover the ordered emissive state after smearing. Utilizing both fluorescence lifetime measurements under ambient conditions and phosphorescence lifetime measurements at low temperature, the proposed MLQ mechanism was substantiated. Enhancement of the triplet excited state population may be a phenomenon common to all BF2bdk dyes exhibiting ML. Heavy atoms such as iodide simply make the effect much more obvious. All of the dyes in the aforementioned halide study bore C12H25 alkyl chain substituents, and all of their singlet excited state energies were found to be alterable by mechanical perturbation. However, their triplet states were relatively insensitive to this stimulus. Therefore, we were curious to see if we could synthesize materials with triplet states that could be perturbed by mechanical force. Given that solid-state fluorescence properties can be tuned by alkyl chain length,29 we wanted to test whether varying the chain length might afford mechanoresponsiveness of the triplet state as well as probe what other effects chain length might have on MLQ and triplet emission enhancement. Thus, we performed a study wherein the halide substituent remains static and the alkyl chain length is modulated. Since the iodine substituent yielded the most prominent MLQ and phosphorescence enhancement in a previous study,30 it seemed the most promising substituent for probing the effects of mechanical perturbation on the triplet state. Six dyes of the form BF2dbm(I)R, where R = H (H), OCH3 (C1), OC5H11 (C5), OC6H13 (C6), OC12H25 (C12), and OC18H37 (C18) were synthesized (Figure 1). Both short and long chain derivatives were targeted. Additionally, C5 and C6 were chosen to probe for the presence of even/odd effects as it has previously been reported that these effects play an

Figure 1. Chemical structure of BF2dbm(I)R dyes (R = H (H), OCH3 (C1), OC5H11 (C5), OC6H13 (C6), OC12H25 (C12), OC18H37 (C18).

important role in the self-assembly of alkyl chain molecules.45 The dyes were studied in CH2Cl2 solution and as films on both weighing paper and glass substrates. Density functional theory (DFT) calculations were performed on the H, C1, and C5 dyes in order to optimize ground-state geometries, generate HOMO and LUMO molecular orbital (MO) diagrams, and simulate absorption and emission spectra. Films on glass and pristine dye powders were subjected to powder X-ray diffraction (XRD) analysis to reveal crystalline or amorphous character. Films on glass were also studied using atomic force microscopy (AFM) to investigate morphologies. Solid-state quantum yield measurements were performed. Single crystals of the H and C5 dyes were grown and studied by single crystal XRD. Differential scanning calorimetery (DSC) was also utilized to detect thermal transitions in the pristine dye powders.



EXPERIMENTAL SECTION Materials. 4-Methoxyacetophenone, 4-pentyloxyacetophenone, 4-hexyloxyacetophenone, 4-dodecyloxyacetophenone, and 4-octadecyloxyacetophenone were synthesized via a Williamson ether synthesis as previously reported.29 Solvents CH2Cl2 and THF were dried over 3 Å molecular sieves according to a previously reported method.46 All other chemicals were purchased from Sigma-Aldrich and used without further purification. Synthetic procedures for the dyes are provided in the Supporting Information. Methods. 1H NMR spectra were recorded on Varian UnityInova 300/51 (300 MHz) or Varian VMRS/600 (600 MHz) instruments in CDCl3. 1H NMR spectra were referenced to the signal for the chloroform residual proton at 7.26 ppm, and coupling constants are given in Hertz. Mass spectra were recorded using either an Applied Biosystems 4800 spectrometer with a MALDI TOF/TOF analyzer or with a Micromass Q-TOF Ultima spectrometer using electrospray ionization (ESI) MS techniques. Melting points were recorded on a MelTemp II by Laboratory Devices. UV−vis spectra were recorded on a Hewlett-Packard 8452A diode-array spectrophotometer. Steady-state fluorescence emission spectra were recorded on a Horiba Fluorolog-3 model FL3-22 spectrofluorometer (doublegrating excitation and double-grating emission monochromator). A 2 ms delay was used in recording delayed emission spectra. Time-correlated single-photon counting (TCSPC) fluorescence lifetime measurements were performed with a NanoLED-370 (λex = 369 nm; duration 2σ(I).



RESULTS AND DISCUSSION Optical Properties of Dyes in Solution. Synthesis of the BF2dbm(I)R dyes via Claisen condensation of methyl 4iodobenzoate with the appropriate 4-alkoxyacetophenone followed by boronation, purification by passage through a silica column with CH2Cl2, and then recrystallization yielded yellow, emissive powders. The optical properties were studied in dilute solution (Table 1 and Figure 2). With the exception of Table 1. Absorption and Emission Properties of Boron Dyes in CH2Cl2.a dye

λabsb [nm]

ε [M−1 cm−1]

λemc [nm]

ΦF [%]

τ [ns]

H C1 C5 C6 C12 C18

390 411 409 409 409 408

45,000 56,000 61,000 62,000 63,000 58,000

412 439 444 445 444 443

9 55 70 64 67e 29

0.2 2.0 1.2 1.2 1.5e 1.2

λex = 369 nm; room temperature, air. bAbsorbance maximum. Emission maximum; fluorescence. eValues taken from ref 32.

a c

the H dye (i.e., without an alkoxy tail), all dyes exhibited extinction coefficients >50 000 M−1 cm−1. The H dye was also, by far, the most blue-shifted of the dyes with respect to both absorption and emission. The presence of an electron-donating alkoxy group has the effect of red-shifting the absorption by ∼20 nm and the emission by ∼30 nm. This is presumably due to the π donating lone pairs on the oxygen atom having a similar effect to increasing conjugation (i.e., raising the energy of the HOMO). The alkoxyl substituent also effects a dramatic increase in both fluorescence quantum yield and lifetime; the H dye has a quantum yield of only 9% and a lifetime of 0.2 ns while all other dyes have quantum yields >25% and lifetimes >1 ns. As with iodide-free BF2dbm complexes,29 there is an increase in quantum yield when the tail length is increased from C1 to C5. For iodide dyes, there is also a decrease in quantum yield when the chain length is increased from C12 to C18 perhaps due to more ways to dissipate energy non-radiatively for the longer chain. In order to gain further insight into the optical properties in solution, density functional theory (DFT) calculations were performed on the H, C1, and C5 compounds. Computational limitations prevented analysis of the longer chain dyes using the same method. The computed absorption and emission maxima C

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

absorption and emission maxima of the dyes bearing chains differ very little from each other. Mechanochromic Luminescence on Paper. For ready comparison with previous reports, weighing paper was used as a substrate for visualizing optical phenomena and taking measurements at 77 K submerged in liquid N2.30,32 First, it was necessary to determine the optimum annealing temperature for each dye as this can vary on a case-by-case basis.30 Different temperatures were tested for each dye until a stable maximal blue-shift was achieved (Figure S3). As with previous systems, 110 °C was sufficient for annealing the H, C1, and C12 dyes.29−32 However, the C5, C6, and C18 dyes required slightly higher temperatures to achieve the same blue-shifted maxima and narrow full widths at half maxima (fwhm). These temperatures were 150 °C for C5 and C6 and 120 °C for C18. Just as in solution, the H dye had the most blue-shifted emission as a film on weighing paper (λem = 451 nm). However, this dye is extremely dim at room temperature under air even when annealed, suggesting a high degree of quenching. In addition, the H dye did not show the typical red-shift in emission upon smearing the annealed material (Figure 4, Table

Figure 2. UV−vis absorption spectra (top) and steady-state fluorescence spectra (bottom) of BF2dbm(I)R dyes (R = H, C1, C5, C6, C12, C18) in CH2Cl2 solution (5 × 10−6 M; λex = 369 nm; room temperature, air).

for all three dyes are in good agreement with the results obtained experimentally (Table S2, Figure S1, Table S4, and Figure S2). The HOMO and LUMO diagrams show a transition mostly π to π* in character with only a slight shift in amplitude from the arene rings to the BF2 diketone moiety when going from HOMO to LUMO (Figure 3).

Figure 4. Emission spectra of BF2dbm(I)OR dyes as films on weighing paper in both thermally annealed (TA) and smeared (SM) forms (λex = 369 nm; room temperature, air).

Figure 3. Highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) diagrams of the H, C1, and C5 dyes (left to right).

2). The C1 dye also did not show a red-shift in peak emission; however, it did show a broadening of the fwhm, an increase in intensity of a red-shifted shoulder at ∼510 nm, and a decrease in overall emission intensity suggesting the formation of lower energy emitting aggregates. The C5, C6, and C12 dyes show only slight red-shifts in emission after smearing (i.e., only ∼10 nm) along with, once again, a broadening of the peaks and a decrease in emission intensity. Carrying the C1, C5, C6, and C12 dye films through seven cycles of annealing followed by smearing, it was found that the emission intensity was recoverable by reannealing (Figure S4). All of these dyes also show perturbations in their pre-exponential weighted lifetimes (τpw0) after smearing. The C18 dye showed ML behavior typical of most other BF2dbm derived materials, with a more dramatic red-shift upon smearing the annealed material (i.e., ∼30 nm), and an increase in both τpw0 and fwhm.28,29 The C18 dye was unique in the set because, instead of the usual decrease in emission intensity, this dye actually showed a slight increase in emission intensity when

A difference in amplitude distribution was observed when comparing the generated molecular orbital diagram images for the three compounds. For the H dye there is, qualitatively, more amplitude concentrated on the iodine heavy atom in the HOMO when compared to the C1 and C5 dyes. This means that the H dye should experience a more pronounced heavy atom effect in which enhanced spin−orbit coupling increases crossover from the S1 excited state to the T1 excited state, thus decreasing fluorescence lifetime and quantum yield. This prediction is confirmed by the experimental results wherein the C1 dye has a quantum yield of 55% and a fluorescence lifetime of 2.0 ns, and these same values for the C5 dye are 70% and 1.2 ns, respectively. The H dye has corresponding values of only 9% and 0.2 ns. Looking at the MOs of the C5 dye, it can be seen that the length of the alkyl chain beyond C1 has no effect on the distribution of amplitude in the HOMO or LUMO. This bears out experimentally, in that solution D

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C Table 2. Luminescence Properties of Dye Films on Weighing Papera thermally annealed

smeared

dye

λemb [nm]

τpw0c [ns]

fwhmd [nm]

λem [nm]

τpw0c [ns]

fwhmd [nm]

H C1 C5 C6 C12 C18

451 481 493 504 502 480

0.37 0.04 1.05 2.42 1.29 0.16

74 28 67 78 87 83

457 480 497 511 513 518

0.83 0.06 0.69 0.45 1.01 0.98

117 68 123 90 95 122

λex = 369 nm; room temperature, air. bEmission maximum; fluorescence. cPre-exponential weighted fluorescence lifetime.47 dFull width at half maximum.

a

Table 3. Delayed Emission Properties of Dye Films on Weighing Papera thermally annealed

smeared

dye

λemb [nm]

τpw0c [ms]

fwhmd [nm]

λem [nm]

τpw0c [ms]

fwhmd [nm]

H C1 C5 C6 C12 C18

607 581 598 580 573 567

0.06 16.6 10.3 11.4 14.4 3.6

106 92 89 81 78 72

611 621 618 590 580 578

0.05 9.8 8.6 11.5 35.2 8.2

95 90 95 95 81 85

λex = 369 nm; 77 K, liquid N2. bEmission maximum; phosphorescence. cPre-exponential weighted phosphorescence lifetime.47 dFull width at half maximum.

a

smeared, suggesting the formation of unique aggregate species (Figure S5). In order to observe the effects of mechanical perturbation on triplet emission (i.e., phosphorescence), measurements were performed on the films while submerged in liquid N2 at 77 K. This was necessary because triplet emission of small molecules is often quenched under air either by nonradiative relaxation or collisional quenching with O2. The C1, C5, C6, and C12 dyes all showed typical MLQ behavior, manifesting as a decrease in emission intensity at room temperature under air along with an increase in phosphorescence intensity at 77 K in liquid N2 (Figures 5 and 6).30,32 Once again, the H dye was unique among the set in that it showed very little mechanical responsiveness; only a slight increase in phosphorescence lifetime and fwhm were observed after smearing the annealed dye (Table 3). Unlike the other dyes, the phosphorescence intensity actually decreased upon smearing. Both this and the Figure 6. Delayed emission spectra of boron dyes on weighing paper in both thermally annealed (TA) and smeared (SM) states (λex = 369 nm; 77 K, liquid N2). The intensities were recorded as photon counts per second (CPS) and shown as CPS/106.

decrease in fluorescence emission intensity at room temperature in air are most likely due to the simple removal of dye material caused by smearing. The MLQ phenomenon is most likely due to a lowering in energy of the S1 excited state, leading to increased intersystem crossing to the T1 excited state.30,32 However, the extent to which triplet emission is enhanced varies. In order to more easily gauge the extent of phosphorescence enhancement, we have defined a phosphorescence enhancement parameter (ISM/ ITA) as the peak intensity of phosphorescence for the smeared sample divided by the peak intensity of phosphorescence for the annealed sample at 77 K in liquid N2. This is displayed in Table 4 along with the difference in the fluorescence and phosphorescence peak maxima calculated from the total

Figure 5. Photographs of the MLQ effect using the C1, C5, and C12 dyes as examples. Samples were photographed under ambient conditions (left) and then immediately after removal from liquid N2 (right). E

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

pronounced in these dyes as thinner spin-cast films on glass than it was when the dyes were smeared as thicker films on weighing paper substrates. This could either be a thickness effect, a substrate effect, or a combination of both. For C5− C18, the emission wavelengths of the as-spun (AS) films are intermediate to the annealed and smeared states. This suggests that the AS films may be a mixture of both the ordered and amorphous emissive species. Once annealed, these dyes exhibit the typical blue-shifted emissions and narrow fwhms. However, the C1 dye shows very little change when the AS film is annealed, indicating that this dye has a strong propensity to form the ordered emissive state similar to iodide-free BF2dbmOMe.29 The C1 dye also shows a significant red-shift in emission when smeared, unlike its behavior as a film on weighing paper. Once again, this may be attributed to a thickness or substrate effect. The H dye showed very little change when the AS film was annealed, only a slight decrease in fwhm. When smeared, the peak emission actually blue-shifted slightly, but this is probably just a side effect of the broadening fwhm. Overall, the emission of the H dye does not seem to change substantially in response to thermal or mechanical processing. Once again, the presence of the donating alkoxy group seems to be necessary to effect significant mechano-responsiveness in this class of dyes. Solid-state quantum yields were recorded for the dyes in both annealed and smeared states as spin-cast films on glass (Table 6). As expected, on the basis of their MLQ behavior, the

Table 4. Triplet-Singlet Energy Gaps and Phosphorescence Enhancement Parameters dye

T1 − S1a [nm]

ISM/ITAb

H C1 C5 C6 C12 C18

151 135 ∼83c 62 77 ∼78c

N/A 1.84 2.62 3.30 6.74 4.79

a T1 − S1 = estimated gap between S1 and T1 excited states; 77 K, liquid N2. bISM/ITA = phosphorescence enhancement parameter, (intensity of phosphorescence when smeared)/(intensity of phosphorescence when annealed); 77 K, liquid N2. cFluorescence maximum estimated from blue-shifted shoulder in total emission spectrum.

emission spectra at 77 K in liquid N2 (Figure S6). It can be clearly seen that a wider gap between fluorescence and phosphorescence peak emissions corresponds to lower phosphorescence enhancement parameters (Table 4). In previous studies we have put forth, and demonstrated compelling evidence for, the idea that the formation of aggregates with lower S1 excited states in response to mechanical force could enhance phosphorescence by increasing intersystem crossing between the S1 and T1 excited states in the presence of a heavy atom.30,32 However, in all of these earlier studies, mechanical force only perturbed the S1 excited state, not the T1 excited state. As can be seen in Table 3 and Figure 6, the shorter-chain C1 and C5 dyes experience a significant redshifting of phosphorescence in response to smearing compared to the other dyes bearing alkyl chains. The reason for this is unclear at this time. Furthermore, the C1 and C5 dyes have the lowest ISM/ITA values. If the T1 excited state energy is lowered as well as the S1 energy upon application of mechanical force, the two states will not be as close to one another as they would be if only the S1 energy was lowered. This results in a larger singlet−triplet energy gap, less intersystem-crossing enhancement, and less enhancement of phosphorescence. Therefore, the extent of phosphorescence enhancement after smearing is tunable by the singlet−triplet energy gap after smearing which is tunable by alkyl chain length. Mechanochromic Luminescence on Glass. Thin films were fabricated by spin-casting dye solutions onto microscope cover glass slides to study substrate effects, spontaneous recovery at room temperature, and solid-state quantum yields. Atomic force microscopy was also used to gauge sample morphology in both TA and SM states. The C5, C6, C12, and C18 dyes all exhibit ML behavior when the annealed films are smeared (Table 5). In fact, the red-shifting is much more

Table 6. Solid-State Luminescence Quantum Yields for SpinCast Dye Films on Glassa thermally annealed Φ [%]

dye

1.94 4.44 14.25 22.80 8.10 1.04

H C1 C5 C6 C12 C18 a

smeared Φ [%]

(±0.28) (±0.18) (±0.26) (±0.33) (±0.41) (±0.05)

1.54 1.81 7.09 4.12 2.37 2.15

(±0.19) (±0.17) (±0.20) (±0.16) (±0.11) (±0.07)

λex = 369 nm; room temperature, air.

C1−C12 dyes showed a decrease in fluorescence quantum yield upon smearing of the annealed films. The C6 and C12 dyes showed the must substantial decreases in fluorescence quantum yield in response to mechanical force (82% and 70% decreases, respectively). The H dye showed only a very slight decrease in quantum yield upon smearing, which is consistent with its lesser degree of mechanical responsiveness. A trend of increasing quantum yield of the annealed dyes is observed moving from H to C6, with C6 having the highest quantum

Table 5. Emission Properties of Spin-Cast Dye Films on Microscope Cover Glassa as-spun

thermally annealed

smeared

dye

λemb [nm]

τpw0c [ns]

fwhmd [nm]

λemb [nm]

τpw0c [ns]

fwhmd [nm]

λemb [nm]

τpw0c [ns]

fwhmd [nm]

H C1 C5 C6 C12 C18

561 481 549 527 513 484

2.21 0.21 1.54 0.41 0.75 0.27

127 34 132 120 107 69

565 480 488 490 479 478

1.66 0.04 1.34 3.25 1.44 0.11

112 30 76 79 68 69

541 550 557 536 513 522

1.24 1.35 1.84 1.40 0.15 0.79

154 132 128 122 104 110

λex = 369 nm; room temperature, air. bEmission maximum; fluorescence. cPre-exponential weighted fluorescence lifetime.47 dFull width at half maximum.

a

F

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C yield of the set. At C12 and C18, however, the quantum yield diminishes substantially. It seems that, much like in solution, increasing tail length raises quantum efficiency until a threshold is reached, at which point the quantum yield begins to attenuate. Finally, The C18 dye showed a slight increase in fluorescence quantum yield upon smearing the annealed film consistent with the previously observed increase in intensity of the weighing paper films when smeared. This could possibly explain why the C18 dye shows a mitigated increase in phosphorescence intensity at low temperature by comparison to C12. The process causing this increase in fluorescence quantum yield must diminish the effects of MLQ and lowtemperature phosphorescence enhancement. To measure spontaneous recovery of thin films at room temperature, the films were fabricated, annealed at the appropriate temperature for 10 min, smeared, monitored with steady-state fluorescence spectroscopy for 1 week, and then reannealed (Figures 7 and 8). Alkyl chain length had substantial

Figure 8. Spin-cast films of the C1 and C18 dyes going from as-spun (AS) to thermally annealed (TA) to smeared (SM) states (room temperature, air).

annealed state and the recovery ability of the dye after smearing. The C1 dye also displays very few peak emissions between the smeared and recovered states while the C5 dye, for example, recovers much more slowly with many intermediate peak emissions. In studies concerning the reversible ML behavior of BF2AVB molecular crystals by Reddy et al., it was discovered that the recovery to the ordered emissive state after mechanical perturbation of more rigid forms of the crystals was more facile compared to more elastic forms.52 In this study, the narrow fwhm and more structured peak observed for the C1 dye suggest a more ordered, rigid annealed state. Recovery to such a state should happen more quickly than that to a less ordered, more elastic annealed state. Excitation spectra were recorded for the C1 and C5 dyes in their TA and SM states as well as several points during the spontaneous recovery of the materials under ambient conditions to see if emissions were arising from distinct ground-state species (Figure S8). After smearing the annealed C1 dye, the excitation peak red-shifts to a new peak and then rapidly recovers the original peak, just like in the emission spectra. On the other hand, the excitation spectra of the annealed C5 dye exhibits two peaks, a more intense blue-shifted peak and a less intense red-shifted peak. When the film is smeared, the intensity ratio switches in favor of the red-shifted peak, and then the intensity of the blue-shifted peak is slowly recovered. Therefore, each peak emission along the recovery path does not represent a distinct ground-state species; rather, the partially recovered spectra represent unique mixtures of ordered and metastable amorphous emissive species. The emission profile of both the C6 and C18 dyes cease recovery at a certain point after which reannealing is required to recover the ordered emissive state. The materials may become trapped in a local energy minimum. This is probably also true of the C12 dye. It seems likely that the longer alkyl chains may create a more elastic environment and promote the formation of a much less rigid ordered emissive state than the shorter chains. This could be a reason why recovery to the ordered emissive states is much slower in the longer chained dyes. In previous studies, annealing the AS films of BF2dbk dyes has caused a change from an amorphous morphology to a much more ordered morphology.22,28−30,32,53 This phenomenon also supports why the TA dye films have blue-shifted emissions and narrow fwhms when compared to those of the AS films and why these dyes exhibit ML, since the AS films seem to correlate with the SM state.28 However, as was mentioned before, the alkyl chain length has a significant effect on the tendency to form the crystalline TA form, with shorter-chained dyes having a much greater affinity for this state. These variable tendencies toward the ordered emissive state are further demonstrated utilizing AFM imaging (Figure 9). The H, C1, C5, and C18 dyes were chosen as representative examples. The H dye exhibits rod-like crystallites in the AS state which, qualitatively,

Figure 7. (A) Emission spectra of boron dyes as spin-cast films on glass (λex = 369 nm) (room temperature, air). The dyes were thermally annealed (TA) and smeared (SM), and then the emission spectra were monitored over time. After 1 week, the films were reannealed (RA). Note: min = minutes, d = days. (B) Recovery of the C1−C18 dyes represented as percent recovery of the annealed emission maxima. The graph on the right shows the first hour of recovery for all dyes.

effects on spontaneous recovery as previously reported for other BF2dbmOR samples.29 The length of the chain had a direct influence on the ability of the dyes to recover the ordered emissive state, with longer chains hindering recovery. In fact, the C12 dye did not recover fully with reannealing after 1 week. Bulky substituents, such as long alkyl chains, seem to hinder the recovery of the ordered emissive state. The fwhm is relatively narrow for the C1 dye in the TA state, and this dye, by far, shows the most rapid recovery in the set. Therefore, there seems to be a correlation between the degree of order in the G

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

devoid of peaks. The C12 dye showed the least change after smearing with only slight decreases in the intensities of a peaks at ∼17°, ∼27°, ∼23°, and ∼41°. Single Crystal XRD. Crystals suitable for single crystal XRD were grown of the H and C5 dyes by slow evaporation from hexanes/acetone. Crystal growth of the other dyes was attempted using similar methods, but was not successful. The H and C5 crystals exhibited distinct emission properties much like their respective solutions and films on both weighing paper and glass. The C5 crystals were visibly emissive under UV excitation and appeared green to the eye (506 nm). On the other hand, emission from crystals of the H dye was not visible, but fluorescence spectroscopy revealed a broad emission profile with a relatively sharp peak at 479 nm (Figure 10).

Figure 9. AFM images of boron dyes as spin-cast films on glass in both as-spun (AS) and thermally annealed (TA) states. Images depict a 10 × 10 μm2 scan. Figure 10. Emission spectra (λex = 369 nm; room temperature, air) of the C5 and H dye crystals used for single crystal XRD analysis under UV light. A photograph of the C5 crystals under UV light is inset. H crystals (not shown) are dim.

do not appear to change significantly when the film is annealed. This lack of change correlates to the limited stimuli responsiveness of emission of the H dye, given that molecular packing controls solid-state emissions. The C1 dye film is also composed of rod-like crystallites that show very little change in size or morphology when annealed. Since this dye does exhibit ML, this seems to be indicative of an affinity for the ordered emissive state and could further help to explain why the emission recovers so rapidly after smearing. On the other hand, the C5 and C18 dyes, which both show significant changes in emission between the AS and TA states, experienced substantial morphological changes when the AS films were annealed. The C5 dye exhibited a change from small, rod-like crystallites to larger block-like crystallites when annealed, and the morphology of the C18 dye was, by and large, amorphous in the AS state. When annealed, a thick patchwork of lamellar crystallites formed, corresponding to the change in emission for these dyes when annealed. X-ray Diffraction (XRD) of Pristine Powders and DropCast Films. X-ray diffraction (XRD) analyses were performed on the dyes as pristine powders and drop-cast films on glass in order to ascertain crystallinity of the samples in various forms (Figure S10). Drop-casting was used to fabricate the films because this method produces thicker films than spin-casting, which are more suitable for diffraction. The diffractograms of the pristine powders exhibited many strong peaks, indicating a high degree of crystallinity. In general, the diffractograms of the dye films showed few differences between the as-cast (AC) and thermally annealed (TA) states. The AC form may represent a mixture of both ordered and amorphous emissive species that is still at least partly crystalline. All dyes showed a decrease in crystallinity after smearing of the TA film. This manifested as either a decrease in intensity of certain peaks or a diffractogram

In addition to differences in emission properties, the H and C5 dye crystals show key differences in crystal packing. Both compounds are highly planar and arrange themselves in offset, J-aggregate type dimer packing (Figures 11 and 12A,B). However, as can be seen in Figures 11 and 12B, the two dyes differ significantly in their dimer π−π stacking. The H dye packs with the BF2 moieties facing in the same direction such that the phenyl ring bearing the iodine substituent is overlapping with the unsubstituted phenyl ring. Conversely, the C5 dye packs with the BF2 moieties facing in opposite directions with the phenyl ring bearing the alkyl chain partially overlapping with both the dioxaborine core of the dye and the chain-bearing phenyl ring of the opposing molecule. This type of stacking with the BF2 moieties facing in opposite directions is typical for BF2bdk dyes and, in particular, BF2bdk dyes known to exhibit ML properties.28,39,52,54,55 With the BF2 moieties facing in opposite directions, the dye molecules would be free to form the H-aggregates proposed by Zhang et al. to be responsible for the red-shifted emission observed upon smearing.39 However, when the BF2 moieties are facing in the same direction, forming the face-to-face H-aggregate would bring the fluorine atoms into direct contact with each other in a stereoelectronically unfavorable interaction. This may explain why the C5 dye exhibits strong mechano-responsiveness while the H dye does not. This structure is also quite different from what has been reported for the BF2dbm dye absent I substitution. Mirochnik et al. have reported J-aggregate type packing of BF2dbm with overlap of the phenyl rings and the molecules arranged in an antiparallel fashion with the BF2 moieties of adjacent molecules positioned opposite each H

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figures 11B and 12B emphasize the short contacts present in the crystal structures. In the H dye, short contacts between fluorine and aromatic hydrogen atoms are present as well as a short I−I contact of 3.85 Å (the van der Waals radius of I−I is 3.96 Å). In comparison, for the C5 dye, the alkyl chains seem to disrupt short I−I contacts (i.e., contacts within the van der Waals radius of 3.96 Å), as none are present. There is, however, a unique short F−C contact occurring between the fluorine atoms of the BF2 moiety and the carbon atoms of the phenyl ring bearing the alkyl chain of an adjacent molecule. The shortest F−C contact is 2.65 Å (the van der Waals radius of F− C is 3.20 Å). A close F−π contact such as this has been previously reported by Ono et al. in the crystal packing of a difluoroboron perfluorotetracene derivative.58 The C5 dye also exhibits intermolecular I−H contacts involving the H atoms of the C5 tail, F−H and B−H contacts with aromatic H atoms, and close C−H and interactions between the C atoms of the iodine-substituted phenyl ring and the H atoms on the alkyl chain of the opposing molecule. There are also B−C interactions between the B atom on one molecule and the phenyl ring bearing the C5 tail on the other molecule. The dimer packing of the H and C5 dye crystals may help to explain their differing emissions. Though they are both packed in offset J-aggregate type dimers, the C5 dye exhibits greater overlap of conjugated π-bonds. When this overlap is maximized in an H-aggregate conformation, emission is much more dramatically red-shifted as demonstrated by Zhang and coworkers,39 so it stands to reason that greater overlap yields more red-shifted emissions. Mirochnik et al. have also demonstrated in crystals of BF2bdks that more efficient πoverlap leads to bathochromic shifts in emission.54,56 Differential Scanning Calorimetry of Dye Powders. Differential scanning calorimetry (DSC) analyses were performed on the dyes in pristine powder form in order to gain more insight into their thermal properties. The results of these studies are summarized in Table 7, and the thermograms

Figure 11. Crystal structure and packing of the H dye from single crystal XRD analysis. (A) Stacked dimer with offset, J-aggregate type π-stacking. The stacking motif overlaps the I-substituted phenyl rings with the non-I-substituted phenyl rings. (B) View highlighting short atom−atom interactions. Close I−I and F−H interactions are shown.

Table 7. Differential Scanning Calorimetry (DSC) Data for Pristine Dyesa dye H C1 C5 C6 C12 C18

Tmb (ΔHc) 235.48 251.32 198.91 191.51 151.87 150.17

(268.6) (326.5) (410.9) (345.7) (289.9) (261.8)

Tcd (ΔHc) 197.90 238.51 182.36 177.30 141.91 144.04

(260.6) (306.2) (403.1) (335.5) (266.8) (285.0)

All data was taken from the 2nd cycle. bMelting point given in °C as the peak of the major endothermic transition. cEnthalpy of the transition given in kJ/mol. dCrystallization temperature given in °C as the peak of the major exothermic transition. a

Figure 12. Crystal structure and packing of the C5 dye from single crystal XRD analysis. (A) Stacked dimer with offset, J-aggregate type π-stacking. The stacking motif places the alkoxyl chain-substituted phenyl COCH2 over the BF2bdk core α-CCO of an adjacent molecule. (B) View highlighting short atom−atom interactions. Close F−C, I−H, I−C, F−H, B−H, and B−C interactions are shown.

are provided in the SI (Figure S9). As expected, the melting temperatures (Tm’s) varied as a function of alkyl chain length, with the H and C1 dyes having the highest melting points (∼235−250 °C), the C5 and C6 dyes having intermediate melting points (∼190−200 °C), and the C12 and C18 dyes having the lowest melting points (∼150 °C). Crystallization temperatures (Tc’s) follow a similar trend and fall in the range ∼140−240 °C. The C12 and C18 dyes both show broad, lower temperature Tm and Tc transitions in addition to the major peaks. We are unable to determine the exact origin of these transitions other than that they must be inherent to longer

another.56 BF2dbm systems with tert-butyl groups and ML display this same antiparallel arrangement of the BF 2 moieties.28,57 I

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX

The Journal of Physical Chemistry C



chained dyes and have been observed in other similar studies.29,30 We can say with a high degree of certainty that they are not liquid crystal transitions due to their large enthalpy values.59 Summary. In conclusion, the effects of alkyl chain length on ML, MLQ, and phosphorescence enhancement were probed by synthesizing and screening a series of iodine-substituted BF2dbm dyes. When compared to the H dye, dyes bearing alkoxy chains displayed much more impressive optical properties in both solution and the solid-state. The presence of an alkoxy chain afforded the high quantum yields in solution as well as significant ML and MLQ behavior as films on both weighing paper and glass. Furthermore, it was found that the length of the chain, itself, had a significant impact on the MLQ and phosphorescence enhancement properties of the dyes. Dyes with shorter chains exhibited smaller phosphorescence enhancement parameters (ISM/ITA) due to a larger gap between the S1 and T1 excited states after smearing. This is due to the fact that mechanical perturbation not only lowers the energy of S1 for dyes bearing shorter chains, but the T1 energy as well. As spin-cast films on glass, dyes bearing longer alkyl chains recovered more slowly and incompletely after smearing. In general, emission spectra, excitation spectra, and AFM images of spin-cast films suggest that shorter-chain dyes are more crystalline than longer chain dyes. Powder XRD of drop-cast films revealed changes from crystalline to amorphous states when annealed films were smeared. Single crystal XRD of the H and C5 dyes revealed differences in crystal packing. The alignment of the BF2 moieties opposite to each another appears to be key to engineering ML functionality. Also, crystals of the H dye revealed unique short I−I contacts while those of the C5 dye revealed close F−π contacts. Finally, the presence of even/ odd effects was probed, but no significant impacts on emission properties were identified as the biggest differences were observed for long versus short chains. With greater insight into the tunable properties of dyes, systematic design of mechanoresponsive materials may be achieved.



REFERENCES

(1) Chi, Z.; Zhang, X.; Xu, B.; Zhou, X.; Ma, C.; Zhang, Y.; Liu, S.; Xu, J. Recent Advances in Organic Mechanofluorochromic Materials. Chem. Soc. Rev. 2012, 41, 3878−3896. (2) Sagara, Y.; Kato, T. Mechanically Induced Luminescence Changes in Molecular Assemblies. Nat. Chem. 2009, 1, 605−610. (3) Kinami, M.; Crenshaw, B. R.; Weder, C. Polyesters with Built-in Threshold Temperature and Deformation Sensors. Chem. Mater. 2006, 18, 946−955. (4) Black, A. L.; Lenhardt, J. M.; Craig, S. L. From Molecular Mechanochemistry to Stress-Responsive Materials. J. Mater. Chem. 2011, 21, 1655−1663. (5) Davis, D. A.d; Hamilton, A.; Yang, J.; Cremar, L. D.; Van Gough, D.; Potisek, S. L.; Ong, M. T.; Braun, P. V.; Martínez, T. J.; White, S. R.; Moore, J. S.; et al. Force-Induced Activation of Covalent Bonds in Mechanoresponsive Polymeric Materials. Nature 2009, 459, 68−72. (6) Yoon, S.-J.; Chung, J. W.; Gierschner, J.; Kim, K. S.; Choi, M.-G.; Kim, D.; Park, S. Multistimuli Two-Color Luminescence Switching via Different Slip-Stacking of Highly Fluorescent Molecular Sheets. Y. J. Am. Chem. Soc. 2010, 132, 13675−13683. (7) Karasawa, S.; Hagihara, R.; Abe, Y.; Harada, N.; Todo, J.-I.; Koga, N. Crystal Structures, Thermal Properties, and Emission Behaviors of N,N-R-Phenyl-7-amino-2,4-trifluoromethylquinoline Derivatives: Supercooled Liquid-to-Crystal Transformation Induced by Mechanical Stimuli. Cryst. Growth Des. 2014, 14, 2468−2478. (8) Chung, K.; Kwon, M. S.; Leung, B. M.; Wong-Foy, A. G.; Kim, M. S.; Kim, J.; Takayama, S.; Gierschner, J.; Matzger, A. J.; Kim, J. Shear-Triggered Crystallization and Light Emission of a Thermally Stable Organic Supercooled Liquid. ACS Cent. Sci. 2015, 1, 94−102. (9) Mukherjee, S.; Thilagar, P. Stimuli and Shape Responsive BoronContaining Luminescent Organic Materials. J. Mater. Chem. C 2016, 4, 2647−2662. (10) Matsumi, N.; Naka, K.; Chujo, Y. Extension of π-Conjugation Length via the Vacant p-Orbital of the Boron Atom. Synthesis of Novel Electron Deficient π-Conjugated Systems by Hydroboration Polymerization and Their Light Blue Emission. J. Am. Chem. Soc. 1998, 120, 5112−5113. (11) Yamaguchi, S.; Akiyama, S.; Tamao, K. Colorimetric Fluoride Ion Sensing by Boron-Containing π-Electron Systems. J. Am. Chem. Soc. 2001, 123, 11372−11375. (12) Yakubovskyi, V. P.; Shandura, M. P.; Kovtun, Y. P. Boradipyrromethenecyanines. Eur. J. Org. Chem. 2009, 19, 3237−3243. (13) Li, D.; Zhang, H.; Wang, Y. Four-Coordinate Organoboron Compounds for Organic Light-Emitting Diodes (OLEDs). Chem. Soc. Rev. 2013, 42, 8416−8433. (14) Hesari, M.; Barbon, S. M.; Staroverov, V. N.; Ding, Z.; Gilroy, J. B. Efficient Electrochemiluminescence of Readily Accessible Boron Difluoride Formazanate Dye. Chem. Commun. 2015, 51, 3766−3769. (15) Zhang, C.; Zhao, J.; Wu, S.; Wang, Z.; Wu, W.; Ma, J.; Guo, S.; Huang, L. Intramolecular RET Enhanced Visible Light-Absorbing Bodipy Organic Triplet Photosensitizers and Application in Photooxidation and Triplet-Triplet Annihilation Upconversion. J. Am. Chem. Soc. 2013, 135, 10566−10578. (16) Zhang, G.; Evans, R. E.; Campbell, K. A.; Fraser, C. L. Role of Boron in the Polymer Chemistry and Photophysical Properties of Difluoroboron−Dibenzoylmethane Polylactide. Macromolecules 2009, 42, 8627−8633. (17) Halik, M.; Wenseleers, W.; Grasso, C.; Stellacci, F.; Zojer, E.; Barlow, S.; Bredas, J.-L.; Perry, J. W.; Marder, S. R. Bis(dioxaborine) Compounds with Large Two-Photon Cross Sections, and their Use in the Photodeposition of Silver. Chem. Commun. 2003, 1490−1491. (18) Cogné-Laage, E.; Allemand, J.-F.; Ruel, O.; Baudin, J.-B.; Croquette, V.; Blanchard-Desce, M.; Jullien, L. Diaroyl(methanato) boron Difluoride Compounds as Medium-Sensitive Two-Photon Fluorescent Probes. Chem. - Eur. J. 2004, 10, 1445−1455. (19) Ono, K.; Yoshikawa, K.; Tsuji, Y.; Yamaguchi, H.; Uozumi, R.; Tomura, M.; Taga, K.; Saito, K. Synthesis and Photoluminescence Properties of BF2 Complexes with 1,3-Diketone Ligands. Tetrahedron 2007, 63, 9354−9358.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.6b03239. Additional synthetic, computational and spectral details; DSC and XRD data (PDF) Crystallographic data for H derivative (CIF) Crystallographic data for C5 derivative (CIF)



Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Phone: (434) 924-7998. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank the National Science Foundation (NSF CHE1213915) for support for this research and Professor Carl O. Trindle for guidance and helpful discussions concerning computational studies and methods. J

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C (20) Chow, Y. L.; Johansson, C. I.; Zhang, Y.-H.; Gautron, R.; Yang, L.; Rassat, A.; Yang, S.-Z. Spectroscopic and Electrochemical Properties of 1,3-Diketonatoboron Derivatives. J. Phys. Org. Chem. 1996, 9, 7−16. (21) Xu, S.; Evans, R. E.; Liu, T.; Zhang, G.; Demas, J. N.; Trindle, C. O.; Fraser, C. L. Aromatic Difluoroboron β-Diketonate Complexes: Effects of π-Conjugation and Media on Optical Properties. Inorg. Chem. 2013, 52, 3597−3610. (22) Liu, T.; Chien, A. D.; Lu, J.; Zhang, G.; Fraser, C. L. Arene Effects on Difluoroboron β-Diketonate Mechanochromic Luminescence. J. Mater. Chem. 2011, 21, 8401−8408. (23) Zhang, G.; Chen, J.; Payne, S. J.; Kooi, S. E.; Demas, J. N.; Fraser, C. L. Multi-Emissive Difluoroboron Dibenzoylmethane Polylactide Exhibiting Intense Fluorescence and Oxygen-Sensitive Room-Temperature Phosphorescence. J. Am. Chem. Soc. 2007, 129, 8942−8943. (24) Samonina-Kosicka, J.; DeRosa, C. A.; Morris, W. A.; Fan, Z.; Fraser, C. L. Dual-Emissive Difluoroboron Naphthyl-Phenyl βDiketonate Polylactide Materials: Effects of Heavy Atom Placement and Polymer Molecular Weight. Macromolecules 2014, 47, 3736−3746. (25) DeRosa, C. A.; Samonina-Kosicka, J.; Fan, Z.; Hendargo, H. C.; Weitzel, D. H.; Palmer, G. M.; Fraser, C. L. Oxygen Sensing Difluoroboron Dinaphthoylmethane Polylactide. Macromolecules 2015, 48, 2967−2977. (26) Zhang, X.; Liu, X.; Lu, R.; Zhang, H.; Gong, P. Fast Detection of Organic Vapors Based on Fluorescent Nanofibrils Fabricated from Triphenylamine Functionalized β-Diketone-Boron Difluoride. J. Mater. Chem. 2012, 22, 1167−1172. (27) Liu, X.; Zhang, X.; Lu, R.; Xue, P.; Xu, D.; Zhou, H. LowDimensional Nanostructures Fabricated from Bis(dioxaborine)carbazole Derivatives as Fluorescent Chemosensors for Detecting Organic Amine Vapors. J. Mater. Chem. 2011, 21, 8756−8765. (28) Zhang, G.; Lu, J.; Sabat, M.; Fraser, C. L. Polymorphism and Reversible Mechanochromic Luminescence for Solid-State Difluoroboron Avobenzone. J. Am. Chem. Soc. 2010, 132, 2160−2162. (29) Nguyen, N. D.; Zhang, G.; Lu, J.; Sherman, A. E.; Fraser, C. L. Alkyl Chain Length Effects on Solid-State Difluoroboron β-Diketonate Mechanochromic Luminescence. J. Mater. Chem. 2011, 21, 8409− 8415. (30) Morris, W. A.; Liu, T.; Fraser, C. L. Mechanochromic Luminescence of Halide-Substituted Difluoroboron β-Diketonate Dyes. J. Mater. Chem. C 2015, 3, 352−363. (31) Zhang, G.; Singer, J. P.; Kooi, S. E.; Evans, R. E.; Thomas, E. L.; Fraser, C. L. Reversible Solid-State Mechanochromic Fluorescence from a Boron Lipid Dye. J. Mater. Chem. 2011, 21, 8295−8299. (32) Zhang, G.; Lu, J.; Fraser, C. L. Mechanochromic Luminescence Quenching: Force-Enhanced Singlet-to-Triplet Intersystem Crossing for Iodide-Substituted Difluoroboron-Dibenzoylmethane-Dodecane in the Solid-State. Inorg. Chem. 2010, 49, 10747−10749. (33) Ito, H.; Saito, T.; Oshima, N.; Kitamura, N.; Ishizaka, S.; Hinatsu, Y.; Wakeshima, M.; Kato, M.; Tsuge, K.; Sawamura, M. Reversible Mechanochromic Luminescence of [(C6F5Au)2(μ-1,4Diisocyanobenzen)]. J. Am. Chem. Soc. 2008, 130, 10044−10045. (34) Ni, J.; Zhang, X.; Qiu, N.; Wu, Y.-H.; Zhang, L.-Y.; Zhang, J.; Chen, Z.-N. Mechanochromic Luminescence Switch of Platinum(II) Complexes with 5-Trimethylsilylethynyl-2,2′-bipyridine. Inorg. Chem. 2011, 50, 9090−9096. (35) Sagara, Y.; Kato, T. Stimuli-Responsive Luminescent Liquid Crystals: Change of Photoluminescent Colors Triggered by a ShearInduced Phase Transition. Angew. Chem., Int. Ed. 2008, 47, 5175− 5178. (36) Sagara, Y.; Yamane, S.; Mutai, T.; Araki, K.; Kato, T. A StimuliResponsive, Photoluminescent, Anthracene-Based Liquid Crystal: Emission Color Detemined by Thermal and Mechanical Processes. Adv. Funct. Mater. 2009, 19, 1869−1875. (37) Löwe, C.; Weder, C. Oligo (p-phenylene vinylene) Excimers as Molecular Probes: Deformation-Induced Color Changes in Photoluminescent Polymer Blends. Adv. Mater. 2002, 14, 1625−1629.

(38) Crenshaw, B. R.; Weder, C. Deformation-Induced Color Changes in Melt-Processed Photoluminescent Polymer Blends. Chem. Mater. 2003, 15, 4717−4724. (39) Sun, X.; Zhang, X.; Li, X.; Liu, S.; Zhang, G. A Mechanistic Investigation of Mechanochromic Luminescent Organoboron Materials. J. Mater. Chem. 2012, 22, 17332−17339. (40) Liang, W. Y. Excitons. Phys. Educ. 1970, 5, 226−228. (41) Galer, P.; Korošec, R. C.; Vidmar, M.; Šket, B. Crystal Structures and Emission Properties of the BF2 Complex 1-Phenyl-3-(3,5dimethoxyphenyl)-propane-1,3-dione: Multiple Chromisms, Aggregation- or Crystallization-Induced Emission, and the Self-Assembly Effect. J. Am. Chem. Soc. 2014, 136, 7383−7394. (42) Nagai, A.; Kokado, K.; Nagata, Y.; Arita, M.; Chujo, Y. J. Highly Intense Fluorescent Diarylboron Diketonate. J. Org. Chem. 2008, 73, 8605−8607. (43) Yoshii, R.; Nagai, A.; Tanaka, K.; Chujo, Y. Highly Emissive Boron Ketoiminate Derivatives as a New Class of Aggregation-Induced Emission Fluorophores. Chem. - Eur. J. 2013, 19, 4506−4512. (44) Yoshii, R.; Suenaga, K.; Tanaka, K.; Chujo, Y. Mechanofluorochromic Materials Based on Aggregation-Induced EmissionActive Boron Ketoiminates: Regulation of the Direction of the Emission Color Changes. Chem. - Eur. J. 2015, 19, 7231−7237. (45) Hibino, M.; Sumi, A.; Tsuchiya, H.; Hatta, I. Microscopic Origin of the Odd-Even Effect in Monolayer of Fatty Acids Fromed on a Graphite Surface by Scanning Tunneling Microscopy. J. Phys. Chem. B 1998, 102, 4544−4547. (46) Williams, D. B. G.; Lawton, M. Drying of Organic Solvents: Quantitative Evaluation of the Efficiency of Several Desiccants. J. Org. Chem. 2010, 75, 8351−8354. (47) Carraway, E. R.; Demas, J. N.; DeGraff, B. A.; Bacon, J. R. Photophysics and Photochemistry of Oxygen Sensors Based on Luminescent Transition-Metal Complexes. Anal. Chem. 1991, 63, 337−342. (48) Zhu, H.; Wang, X.; Li, Y.; Wang, Z.; Yang, F.; Yang, X. Microwave Synthesis of Fluorescent Carbon Nanoparticles with Electroluminescence Properties. Chem. Commun. 2009, 5118−5120. (49) 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, Revision A.1; Gaussian, Inc.: Wallingford, CT, 2009. (50) Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999−3094. (51) Sheldrick, G. M. SHELXTL; Bruker AXS, Inc.: Madison, WI, 1997. (52) Krishna, G. R.; Kiran, M. S. R. N.; Fraser, C. L.; Ramamurty, U.; Reddy, C. M. The Relationship of Solid-State Plasticity to Mechanochromic Luminescence in Difluoroboron Avobenzone Polymorphs. Adv. Funct. Mater. 2013, 23, 1422−1430. (53) Butler, T.; Morris, W. A.; Samonina-Kosicka, J.; Fraser, C. L. Mechanochromic Luminescence and Aggregation Induced Emission for a Metal-Free β-Diketone. Chem. Commun. 2015, 51, 3359−3362. (54) Mirochnik, A. G.; Bukvetskii, B. V.; Fedorenko, E. V.; Karasev, V. E. Crystal Structures and Excimer Fluorescence of Anisoylbenzoylmethanatoboron Dianisoylmethanatoboron Difluorides. Russ. Chem. Bull. 2004, 53, 291−296. (55) Sakai, A.; Ohta, E.; Yoshimoto, Y.; Tanaka, M.; Matsui, Y.; Mizuno, K.; Ikeda, H. New Fluorescence Domain “Excited Multimer” Formed upon Photoexcitation of Continuously Stacked Diaroylmethanatoboron Difluoride Molecules with Fused π-Orbitals in Crystals. Chem. - Eur. J. 2015, 50, 18128−18137. (56) Mirochnik, A. G.; Bukvetskii, B. V.; Gukhman, E. V.; Zhikhareva, P. A.; Karasev, V. E. Crystal Structure and Excimer Fluorescence of Dibenzoylmethanatoboron Difluoride. Russ. Chem. Bull. 2001, 50, 1612−1615. (57) Krishna, G. R.; Devarapalli, R.; Prusty, R.; Liu, T.; Fraser, C. L.; Ramamurty, U.; Reddy, C. M. Structure-Mechanical Property Correlations in Mechanochromic Luminescent Crystals of Boron Difluoride Dibenzoylmethane Derivatives. IUCrJ 2015, 2, 611−619. K

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C (58) Ono, K.; Hashizume, J.; Yamaguchi, H.; Tomura, M.; Nishida, J.-i.; Yamashita, Y. Synthesis, Crystal Structure, and Electron-Accepting Property of BF2 Complex of Dihydroxydione with a Perfluorotetracene Skeleton. Org. Lett. 2009, 11, 4326−4329. (59) Singh, S. Phase Transitions in Liquid Crystals. Phys. Rep. 2000, 324, 107−269.

L

DOI: 10.1021/acs.jpcc.6b03239 J. Phys. Chem. C XXXX, XXX, XXX−XXX