Chemistry Properties of Distyryl-BODIPY Derivatives

6 hours ago - Singlet oxygen is the key elements for photodynamic therapy. In this paper, six novel distyryl-BODIPY compounds were synthesized and ...
0 downloads 0 Views 449KB Size
Subscriber access provided by Miami University Libraries

A: Spectroscopy, Molecular Structure, and Quantum Chemistry

Photophysical/Chemistry Properties of Distyryl-BODIPY Derivatives: An Experimental and Density Functional Theoretical Study Hongwei Kang, YuBing Si, Yuxiu Liu, Xiaofan Zhang, Weiwei Zhang, Yi Zhao, Baocheng Yang, Yaxuan Liu, and Zhongyi Liu J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b02656 • Publication Date (Web): 05 Jun 2018 Downloaded from http://pubs.acs.org on June 5, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Photophysical/chemistry Properties of Distyryl-BODIPY Derivatives: An Experimental and Density Functional Theoretical Study Hongwei Kanga, Yubing Sia*, Yuxiu Liub, Xiaofan Zhanga, Weiwei Zhangc*, Yi Zhaob, Baocheng Yanga, Yaxuan Liua and Zhongyi Liud* a.

Henan Provincial Key Laboratory of Nanocomposites and Applications, Institute of Nanostructured

Functional Materials, Huanghe Science and Technology College, Zhengzhou 450006, People’s Republic of China. b.

State Key Laboratory of Physical Chemistry of Solid Surfaces, Collaborative Innovation Center of

Chemistry for Energy Materials, and College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, People’s Republic of China. c.

Department of Mechanical and Nuclear Engineering, Pennsylvania State University, University Park,

Pennsylvania 16802, United States. d.

College of Chemistry and Molecular Engineering, Zhengzhou University, 100 Kexue Avenue, 450001,

People’s Republic of China. *Corresponding author, E-mail address: [email protected] (Y. Si), [email protected] (W. Zhang), [email protected] (Z. Liu).

Abstract: Singlet oxygen is the key elements for photodynamic therapy. In this paper, six novel distyryl-BODIPY compounds were synthesized and investigated in detail to fully evaluate their photophysical/chemistry characteristics. Specially, the singlet oxygen 1O2 quantum yield of compounds 2 and 4 each bearing two bromine atoms in their skeleton revealed the position effect of heavy atom for 1O2 production. The 1O2 quantum yield of 4, which was brominated at 2/6 position of BODIPY skeleton, was much higher than that of compound 2, brominated at styryl group with a long distance towards BODIPY core. Importantly, theoretical calculations were carried out to elaborate the essential reason for the difference of 2 and 4 by investigating intersystem crossing rate.

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1 Introduction Dipyrromethene boron difuoride (BODIPY) dyes have attracted widespread attention in recent decades.1-3 The distinguished features, including easy availability by organic synthesis, high fluorescence quantum yield as well as low toxicity and splendid stabilities, allow them to be promising candidates for application in contemporary technologies, such as sensors, biology imaging and solar cells.4-9 Among these applications, one notable example is the photodynamic therapy, in which the BODIPY-based materials, having the abilities to translate triplet oxygen molecule into singlet oxygen, are widely employed.10-14 However, due to spin-forbidden transition, these conventional BODIPYs suffering from the low triplet state are invalid. So, heavy atoms are often introduced to enhance the transition probability,15 and it is proved that these atoms as well as their position are very crucial to trigger the singlet to triplet states (S→T) transition. In clinical photodynamic therapy process, long-wavelength light sources are preferable for the greater tissue penetration of long-wavelength light than that of short-wavelength light. Thus, for utilizing the long wavelength, these BODIPY sensitizers having broad absorption are in great request. In this sense, much effort has been paid to broaden the absorption region and regulate the frontier orbital levels of BODIPYs, by developing the sophisticated styryl-BODIPYs. Based on the current achievements, a plenty of methods are available now to synthesize and modify BODIPY materials, including extension the π conjugation along 2/6 direction,16-18 decoration in the meso-position,19-22 fusing aromatic rings onto pyrrole units,23,24 or 2

ACS Paragon Plus Environment

Page 2 of 24

Page 3 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

functionalization in α/β positions.25,26 Specially, embedding styryl groups in α/β positions of BODIPYs represents a particular interest, and it turned out to be a very effective way to tune the optical or electrical properties of these advanced BODIPYs, via simply reacting the active methyl group with benzaldehyde.27-29 However, to the best of our knowledge, the comparative study of heavy atoms especially for the halogenated distyryl-BODIPYs is not enough. Therefore, investigation of the heavy atom as well as their position effect is significant towards the preparation of the next generation drugs for photodynamic therapy. Herein, six novel molecules bearing 3, 5-distyryl-BODIPY backbone were synthesized, their optical and electrical properties were investigated in detail to evaluate their physical characteristics. The singlet oxygen (1O2) photosensitizing experiments were carried out to calculate the quantum yield of 1O2. Importantly, theoretical calculations were carried out to elaborate the essential reason for the different 1O2 production abilities of 2 and 4 by investigating the spin-orbit coupling (SOC) effect.

2. Experiment section 2.1 Instrumentation and method The microwave experiments were carried out using Biotage microwave reactor with normal power model. 1H NMR (400 MHz) and

13

C NMR (100 MHz) spectra were

performed in CDCl3 with TMS as internal reference on a Bruker ADVANCE 400 NMR Spectrometer. Mass spectra (Q-TOF-EIS) were determined on a Bruker 3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 24

BIFLEX III Mass Spectrometer. Elemental analysis was carried on PerkinElmer 2400. Thermal properties were detected using Thermogravimetric Analysis (Q600 SDT). UV-Vis

spectra

were

measured

with

Hitachi

(Model

U-4010)

UV-Vis

spectrophotometer. Differential pulse voltammetrys (DPV) and cyclic voltammetry (CV) were tested on a Zahner IM6e electrochemical workstation at a scan rate of 50 mV s-1. All the measurements were carried out using an electrochemical cell with three electrodes, with a glassy carbon discs as the working electrode, a Pt wire as the counter electrode, and an Ag/AgCl electrode as the reference electrode, meanwhile ferrocene/ferrocenium (Fc/Fc+) was used as an external reference for the calibration of potential and 0.1 M tetrabutylammonium hexafluorophosphate (Bu4NPF6) dissolved in CH2Cl2 (HPLC grade) was employed as the supporting electrolyte. The condensation reaction procedures for compounds 1-3 were as following: 0.2 mmol of BODIPY substrate and 0.35 equivalent of aromatic aldehyde were added to a microwave tube, 4 ml of dry DMF was used to dissolve the substrates, then 100 µl of piperidine and 100 µl of HAc were added to the solution. The tube was sealed with an aluminum cap and heated for 10 - 45 min at 130 ºC under microwave radiation. After that, the solvent was evaporated with rotary evaporator using oil pump as vacuum source. Then the target compounds were purified by column chromatography (100-200 M, silica gel) using DCM/PE as eluent. 2.2 Theoretical calculation The singlet ground states were optimized at the density functional theory (DFT) with B3LYP/6-31G(d) level implemented with polarizable continuum model (PCM) 4

ACS Paragon Plus Environment

Page 5 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

to consider the solvent effect, and the static dielectric constant of dichloromethane (ϵ = 8.93) was used. The vibrational frequency analysis was performed to verify the optimized geometries are local minima. All the optimizations were manipulated by the Gaussian 09 suite.30 The intersystem crossing rate from a singlet to triplet state is estimated from Marcus formula:31 𝑘𝐼𝑆𝐶 =

1 π −( 𝜆 + ∆𝐺0 )2 VSOC 2 √ exp[ ] ℏ 𝜆𝑘𝐵 𝑇 4𝜆𝑘𝐵 𝑇

where, kB is the Boltzmann constant and T is the room temperature of 298.15K. λ is reorganization energy and ΔG0 represents the driving force. The SOC constants (VSOC) were calculated with ADF 4.0 program package32 at the DFT/B3LYP/TZ2P level, in which TZ2P is a core double-ξ, valence triplet-ξ, doubly polarized basis known as slater-type orbits. The two-component zero-order regular approximation (ZORA) is employed for the solution of the four-component ̂ |𝜓𝑗𝑇 ⟩, where 𝜓𝑖𝑆 Dirac−Kohn−Sham equation. The SOC is given by ESO ≈ ⟨𝜓𝑖𝑆 |𝐻𝑠𝑜 ̂ is and 𝜓𝑗𝑇 are the wave functions of single and triplet states, respectively, and 𝐻𝑠𝑜 SOC operator within ZORA.

3. Results and discussion 3.1 Synthesis procedure optimization

Scheme 1. The synthetic route of styryl-BODIPY via Knoevenagel reaction. 5

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Initially,

the

reaction

of

8-(4-bromophenyl)-tetramethyl

difluoroboradipyrromethene and 4-dimethylamino benzaldehyde was chosen as a model, as shown in Scheme 1, which took nearly 100 hours to consume the substrates completely in toluene.33 It was observed that the solvents had the greatest influence on conversion and yields than other conditions. As shown in Table S1, the reaction in DMF expressed the fastest rate and the BODIPY substrate was consumed completely within 2h, while the reaction was proceeded much slowly in toluene or ethanol, with the conversion no more than 20.4% after heating for 24 h. The faster reaction rate in DMF is probably due to the polarity of the solvent since the transition state complexes of Knoevenagel reaction are easily solvated in polar solvent, thereby decreasing the transition state energy, and accelerating the condensation reaction.34 Additionally, it is clear that the microwave could promote the reaction both in reaction rate and yield than conventional heating in ethanol or DMF. Besides the solvents, the second most important factor is the catalyst, the rate is reduced in this order: piperidine > piperidine·HAc >> piperidine·HCl. Despite that piperidine could accelerate the reaction, however, in most cases, piperidine·HAc system showed a better selectivity for the target product, and gave high yields. We also investigated the effects of temperature, reaction time and the amount of benzaldehyde to get a better reaction condition in a yield of nearly 80% for distyryl-BODIPY.

6

ACS Paragon Plus Environment

Page 6 of 24

Page 7 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Scheme 2. The chemical structures of compounds 1-6. Compounds 1-3 were synthesized under Knoevenagel condition, 4 was obtained by bromination of 3, while 5, 6 were prepared via Pd catalyzed coupling reaction.

As shown in Scheme 2, this optimized method was applied to synthesize styryl-BODIPYs 1 and 4, though the yields of 1, 2, 3 are relatively high, while the yield of compound 4 is rather low, probably due to the extremely low solubility of 2, 6 dibromo BODIPY, which needs nearly 10 ml of DMF to dissolve 0.1 mmol of the substrate. Thus, the compound 4 was prepared nearly quantitatively via a changed synthetic condition, by reacting compound 3 with NBS in DCM. The route was illustrated in Scheme 1. Base on the as-prepared compounds 2 and 4, compounds 5 and 6, were synthesized under Suzuki coupling reaction (see supporting information).

3.2 Experimental spectra The absorption spectra of the compounds 1-6 were investigated in dichloromethane and shown in Figure 1. For these distyryl compounds, the strongest absorption over 7

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

630 nm with the molar absorption coefficient up to 1.0 × 105 L mol-1cm-1 are attributed to typical S0-S1 transition; while the broad shoulder absorption peaks at shorter wavelength are assigned to S0-S2 transition.35 The λmax of compounds 1, 2, 3, 4 are 630, 637, 650, and 672 nm respectively, showing that the introduction of electrondonating groups at 3, 5 positions or bromination at 2, 6 position could extend the absorption region to long wavelength direction. The λonset values are 648, 657, 675, 710 nm respectively for 1, 2, 3 and 4, corresponding to narrow highest occupied molecular orbital (HOMO) – lowest unoccupied molecular orbital (LUMO) gaps about 1.89 - 1.75 eV for these styryl-BODIPYs. Besides, the fluorescent properties were also studied in DCM with the concentration of 1 × 10-5 M by excitation at the lowest-energy wavelength. It is clear that, compounds 1 has the highest emission intensity, while compound 4, possessing two bromine atoms at 2/6 position, shows the lowest emission intensity. It indicates that the styryl groups determine the intensity of fluorescence emission greatly, when electron-donating units was incorporated at styryl groups, the fluorescence intensity was depressed. As shown in Table S2, fluorescence quantum yields were measured using rhodamine 6G (φf = 0.95) as standard in ethanol with the absorbance less than 0.05.36 It can be seen that these data are approximately in accordance with the fluorescence intensity.

8

ACS Paragon Plus Environment

Page 8 of 24

Page 9 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1. UV-vis absorption spectra (left) and fluorescence emission (right) of 1 - 6 in DCM (1 × 10-5 M). The fluorescence spectra were obtained by excitation at lowest energy absorption wavelength at room temperature.

Differential pulse voltammetry (DPV) and cyclic voltammetry (CV) were adopted to explore the electrical properties of these novel materials, the data were shown in Figure S1 and S2. It is distinct that the oxidative peaks are changed more notably, with the HOMO levels ranging from -5.50 eV to -5.80 eV, on the contrary, the reduction peaks are shifted in a narrow range with LUMO energy levels at about -4.0 eV, implying that the HOMO levels other than LUMO levels of styryl-BODIPYs are dramatically influenced by changed styryl substituents. The CV waves for oxidation process of these compounds are all reversible, however the reduction waves are much irreversible, suggesting that these materials donate electrons easier than accept electrons, which is agreed with the application as donor materials in solar cell research.37

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. The singlet oxygen (1O2) photosensitizing measurement of compounds 2 and 4. Methylene blue was used as standard (λex= 650 nm, the irradiation power is 10 mW). The concentration of DPBF was 4×10-5 mol/L, and the concentration of dyes is 1×10 -7 mol/L in DCM.

3.3 Intersystem crossing rate Generally, the heavy atom, such as Br here, often facilitates intersystem crossing (ISC) process via vibronic spin-orbit coupling, thus increase the probability of triplet photosensitizer to trigger 1O2 by energy transfer process. Also, some aromatic boronic acid esters with enhanced phosphorescence, generated by triplet process are probably good

1

O2 sensitizers.38 So, we synthesized the brominated and boronic acid

derivatives of distyryl BODIPY to investigate their 1O2 photosensitizing abilities by singlet oxygen generation experiments. In these tests, 1, 3-diphenylisobenzofuran (DPBF) was used as scavenger to probe 1O2. As 1O2 reacts with DPBF immediately and thus the characteristic absorption peak at 414 nm of DPBF declines gradually, hence, the slope of the linear decrease in absorbance can be used to estimate 1O2 quantum yields vicariously. As shown in Figure 2, the linear slopes are -0.00192, -0.01126 and -0.01467 for 2, 4 and methylene blue respectively. Using methylene blue as standard (0.57 in DCM),39, 40 the 1O2 quantum yields of 2 and 4 were deduced to be 0.075 and 0.44 respectively. It indicates that the difference position of heavy atom induced different singlet oxygen generation yields.41, 42 While for both 5, and 6, 10

ACS Paragon Plus Environment

Page 10 of 24

Page 11 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

the significant 1O2 quantum yields were not observed with the low 1O2 yields less than 0.07, suggested that the boronic acid esters units here are no help for the formation of triple states of 5, and 6. According to Marcus equation, the phosphorescence radiative rate constant is proportional to the square of the spin-orbit coupling (SOC) constant, which is linear correlated with the yield of singlet oxygen molecule. Therefore, we firstly compared the SOC constant for compounds 2 and 4. To obtain the SOC, the 20 lowest singlet−singlet and singlet−triplet transitions are taken into account in our calculations. The SOC values are 0.02 cm−1 and 0.16 cm−1 of S1 → T1 for 2 and 4 respectively, while 0.04 cm−1 and 0.10 cm−1 of S1 → T2 higher-level transitions. Based on Marcus theory, the intersysterm crossing rate of 4 will be 64 and 6.25 times larger than 2 for the first (S1 → T1) and higher level states (S1 → T2) respectively in the approximation of the same Frank-Condon weighted density of states (FCWD). However, the widely-used single-reference DFT may lose the accuracy on open shell system, especially for the triplet excited states,43-45 thus, the complete active space multiconfiguration self-consistent field (CASSCF) computation is used to calculate the SOC for the purpose of comparison. The split-valence plus polarization (def2-SVP) basis set (similar to TZ2P) is employed for the atoms. The CASSCF wave functions for the triplet and singlet states of 2 and 4 involve an active space of 10 electrons in 10 orbitals, denoted as (10, 10). The single-point energies of compounds (2 and 4) determined by CASSCF are calibrated by the completed active space second 11

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

order perturbation theory (CASPT2) method, which incorporates the dynamic correlation effects. In the CASPT2 calculation, a level shift of 0.25 a.u. is adopted to avoid the effect of intruder state.46 All the SOC integrals were computed at the 2 and 4 using the atomic mean-field approximation (AMFI) as implemented in the Molpro 2009 software package.

47-51

The SOC values are 0.01 cm-1 and 0.22 cm-1 of S1 → T1

for 2 and 4, while 0.07 cm-1 and 0.31 cm-1 of S1 → T2 transitions. This SOC constants show that DFT/TDDFT predicted consistent results with the high computational cost CASPT2 method, DFT is thus adopted in present work. Considering the intersystem crossing rate is also strongly dependent on other factors such as driving force and reorganization energy,52 here, the lowest singlet excited states of S1 was optimized by time-dependent DFT (TDDFT) at the 6-31+G (d, p) level, and lowest triplet state T1 was optimized by unrestricted DFT (UDFT) with the same basis set to balance the accuracy and efficiency. It is worth to note that only S1 → T1 process was considered since the excited energy of T2 state are higher than S1 for both 2 and 4 at the DFT/B3LYP/TZ2P level and the 6-31+G (d, p) level. It is shown that a 0.82 Debye change in dipole moment accompanies the transition from S1 (7.38 Debye) to T1 (8.20 Debye) for 2 and 1.06 Debye change for 4 (the dipole moments are 4.57 Debye and 3.51 Debye for S1 and T1), therefore, the corresponding solvent reorganization energy should be very small.53, 54 In the following, we only consider the calculations vibrational reorganization energy. Based on the obtained adiabatic states (S1 and T1) the driving forces are -0.03 eV and -0.02 eV for 2 and 4 and the vibrational reorganization energies are 0.08 eV and 0.11 eV for 2 and 4 12

ACS Paragon Plus Environment

Page 12 of 24

Page 13 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

respectively (the details for reorganization energy calculations can be seen in elsewhere,55, 56 are not shown here), the final ISC rate are 3.7 × 105 s-1 and 3.7 × 107 s-1 for compounds 2 and 4, respectively. The ISC rates can be used to semi-quantitative measurement of singlet oxygen quantum yield since the higher excited states are not considered, our calculation is also consistent with the previous conclusion of iodization of BODIPY dye at 2/6 position results in large heavy atom effect. 57 Conclusion In summary, six novel molecules bearing 3, 5-distyryl-BODIPY backbones were synthesized, and the 1O2 photosensitizing experiments of dibromo compounds 2, 4 and two boronic acid esters 5, 6 were investigated. The 1O2 quantum yields of 2 and 4 were 0.075 and 0.44 respectively, and the 1O2 quantum yields of 5, 6 were no more than 0.07, showing that halogenation at the benzene rings of distyryl groups of 3, 5 distyryl-BODIPYs, or boronate esterification of BODIPY had no positive effect on 1

O2 production, while halogenation at the 2/6 position resulted in obviously enhanced

1

O2 yields. Furthermore, theoretical calculations were used to explain the deep reason

for the 1O2 yields difference between the dibromo compounds 2 and 4. The calculated intersystem crossing rates were 3.7 × 105 s-1 and 3.7 × 107 s-1 for 2 and 4, respectively, which was consistent with their 1O2 yields abilities. It was hoped that these results would enrich the research of BODIPYs, and would provide a useful envision for preparation of powerful BODIPYs drugs for photodynamic therapy process.

Acknowledgements 13

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Authors thank Prof. Wenli Zou for his helpful discussions. This work is supported by Science and Technology Development Program of Henan province (162102210329, 172102310164), college science and technology innovation team of Henan province (16IRTSTHN001) and the National Natural Science Foundation of China (21703077, 51602120).

Supporting Information Available: Synthesis details, NMR spectra of BODIPY compounds, photophysical data, electrical properties, thermal properties, IR spectra, HRMS spectra. References: 1. Loudet, A.; Burgess, K. BODIPY Dyes and Their Derivatives:  Syntheses and Spectroscopic Properties. Chem. Rev. 2007, 107, 4891–4932. 2. Li, X.; Gao, X.; Shi, W.; Ma, H. Design Strategies for Water-Soluble Small Molecular Chromogenic and Fluorogenic Probes. Chem. Rev. 2014, 114, 590-659. 3. Boens, N.; Leen, V.; Dehaen, W. Fluorescent Indicators Based on BODIPY. Chem. Soc. Rev. 2012, 41, 1130-1172. 4. Du, R.; Cui, S.; Sun, Z.; Liu, M.; Zhang, Y.; Wu, Q.; Wu, C.; Guo, F.; Zhao, L. Highly Fluorescent Hyperbranched BODIPY-based Conjugated Polymer Dots for Cellular Imaging. Chem. Commun. 2017, 53, 8612-8615.

14

ACS Paragon Plus Environment

Page 14 of 24

Page 15 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

5. Chen, H.; He, X.; Su, M.; Zhai, W.; Zhang, H.; Li, C. A General Strategy toward Highly Fluorogenic Bioprobes Emitting Across the Visible Spectrum. J. Am. Chem. Soc. 2017, 139, 10157-10163. 6. Mao, W.; Xia, L.; Xie, H. Detection of Carbapenemase-Producing Organisms with a Carbapenem-Based Fluorogenic Probe. Angew. Chem. Int. Ed. 2017, 56, 4468-4472. 7. Su, D.; Teoh, C. L.; Gao, N.; Xu, Q. H.; Chang,Y. T. A Simple BODIPY-Based Viscosity Probe for Imaging of Cellular Viscosity in Live Cells. Sensors 2016, 16, 1397. 8. Luo, G. G.; Lu, H.; Zhang, X. L.; Dai, J. C.; Wu, J. H.; Wu, J. J. The Relationship Between the Boron Dipyrromethene (BODIPY) Structure and the Effectiveness

of

Homogeneous

and

Heterogeneous

Solar

Hydrogen-Generating Systems as well as DSSCs. Phys. Chem. Chem. Phys. 2015, 17, 9716-9729. 9. Ziessel, R.; Ulrich, G.; Haefele, A.; Harriman, A. An Artificial Light-Harvesting Array Constructed From Multiple BODIPY Dyes. J. Am. Chem. Soc. 2013, 135, 11330-11344. 10. Zhao, J.; Xu, K.; Yang,W.; Wang, Z.; Zhong, F. The Triplet Excited State of BODIPY: Formation, Modulation and Application. Chem. Soc. Rev. 2015, 44, 8904-8939.

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

11. Kamkaew, A.; Lim, SH.; Lee, HB.; Kiew, LV.; Chung, LY.; Burgess, K. Burgess, BODIPY Dyes in Photodynamic Therapy. Chem. Soc. Rev. 2013, 42, 77-88. 12. Filatov, MA.; Karuthedath, S.; Polestshuk, PM.; Savoie, H.; Flanagan, K. J.; Sy, C.; Sitte, E.; Telitchko, M.; Laquai, F.; Boyle, RW.; Senge, MO. Generation of Triplet Excited States via Photoinduced Electron Transfer in Meso-Anthra-BODIPY: Fluorogenic Response toward Singlet Oxygen in Solution and in Vitro. J. Am. Chem. Soc. 2017, 139, 6282-6285. 13. Zhang, Q.; Cai, Y.; Li, QY.; Hao, LN.; Ma, Z.; Wang, XJ.; Yin, J. Targeted Delivery of Mannose-Conjugated BODIPY Photosensitizer by Nanomicelles for Photodynamic Breast Cancer Therapy. Chem. Eur. J. 2017, 57, 14307–14315. 14. He, H.; Ji, S.; He, Y.; Zhu, A.; Zou, Y.; Deng, Y.; Ke, H.; Yang, H.; Zhao, Y.; Guo, Z.; Chen, H. Photoconversion-Tunable Fluorophore Vesicles for Wavelength-Dependent Photoinduced Cancer Therapy. Adv. Mater. 2017, 19, 1606690. 15. Marin, D. M.; Payerpaj, S.; Collier, G. S.; Ortiz, A. L.; Singh, G.; Jones, M.; Walter, M. G. Efficient Intersystem Crossing Using Singly Halogenated Carbomethoxyphenyl Porphyrins Measured Using Delayed Fluorescence, Chemical Quenching, and Singlet Oxygen Emission. Phys. Chem. Chem. Phys. 2015, 17, 29090-29096.

16

ACS Paragon Plus Environment

Page 16 of 24

Page 17 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

16. Barros, L.; Cardoso, T.; Bihlmeier, A.; Wagner, D.; Kölmel, D.; Hörner, A.; Bräse, S.; Brito-Cruz, C.; Padilha, L. Two-Photon Absorption in a Series of 2, 6-Disubstituted BODIPY Dyes. Phys. Chem. Chem. Phys. 2017, 19, 21683-21690. 17. Zhang, W.; Sheng, W.; Yu, C.; Wei, Y.; Wang, H.; Hao, E.; Jiao, L. One-pot Synthesis and Properties of Well-Defined Dutadiynylene-BODIPY Oligomers. Chem. Commun. 2017, 53, 5318-5321. 18. Gai, L.; Mack, J.; Lu, H.; Yamada, H.; Kuzuhara, D.; Lai, G.; Li, Z.; Shen, Z. New 2,6-Distyryl-Substituted BODIPY Isomers: Synthesis, Photophysical Properties, and Theoretical Calculations. Chem. Eur. J. 2014, 20, 1091-1102. 19. Manjare, S. T.; Kim, J.; Lee, Y.; Churchill, D. G. Facile Meso-BODIPY Annulation and Selective Sensing of Hypochlorite in Water. Org. Lett. 2014, 16, 520-523. 20. Dhokale, B.; Jadhav, T.; Mobin, S. M.; Misra, R. Meso Alkynylated Tetraphenylethylene (TPE) and 2, 3, 3-Triphenylacrylonitrile (TPAN) Substituted BODIPYs. J. Org. Chem. 2015, 80, 8018-8025. 21. Lager, E.; Liu, J.; Aguilar-Aguilar, A.; Tang, B. Z.; Peña-Cabrera, E. Novel Meso-Polyarylamine-BODIPY Hybrids: Synthesis and Study of Their Optical Properties. J. Org. Chem. 2009, 74, 2053-2058. 22. Shivran, N.; Mula, S.; Ghanty, T. K.; Chattopadhyay, S. Steric Strain Release-Directed Regioselective Functionalization of Meso-Methyl Bodipy Dyes. Org. Lett. 2011, 13, 5870-5873. 17

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

23. Hayashi, Y.; Obata, N.; Tamaru, M.; Yamaguchi, S.; Matsuo, Y.; Saeki, A.; Seki, S.; Kureishi, Y.; Saito, S.; Yamaguchi, S.; Shinokubo, H. Facile Synthesis of Biphenyl-Fused BODIPY and its Property. Org. Lett.2012, 14, 866-869. 24. Jiao, C.; Huang, K. W.; Wu, J. Perylene-Fused BODIPY Dye with Near-IR Absorption/Emission and High Photostability. Org. Lett. 2011, 13, 632-635. 25. Leen, V.; Gonzalvo, V. Z.; Deborggraeve, W. M.; Boens, N.; Dehaen, W. Direct Functionalization of BODIPY Dyes by Oxidative Nucleophilic Hydrogen Substitution at the 3- or 3, 5-Positions. Chem. Commun. 2010, 46, 4908-4910. 26. Yu, C.; Jiao, L.; Yin, H.; Zhou, J.; Pang, W.; Wu, Y.; Wang, Z.; Yang, G.; Hao, E. α-/β-Formylated Boron–Dipyrrin (BODIPY) Dyes: Regioselective Syntheses and Photophysical Properties. Eur. J. Org. Chem. 2011, 28, 5460-5468. 27. He, H.; Lo, P. C.; Yeung, S. L.; Fong, W. P.; Ng, D. K. Preparation of Unsymmetrical Distyryl BODIPY Derivatives and Effects of the Styryl Substituents On Their in Vitro Photodynamic Properties. Chem. Commun. 2011, 47, 4748-4750. 28. Frenette, M.; Hatamimoslehabadi, M.; Bellinger-Buckley, S.; Laoui, S.; La, J.; Bag, S.; Mallidi, S.; Hasan, T.; Bouma, B.; Yelleswarapu, C.; Rochford, J. Shining Light On the Dark Side of Imaging: Excited State Absorption

18

ACS Paragon Plus Environment

Page 18 of 24

Page 19 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Enhancement of a Bis-Styryl BODIPY Photoacoustic Contrast Agent. J. Am. Chem. Soc. 2014, 136, 15853-15856. 29. Kostereli,

Z.;

Ozdemir,

T.;

Buyukcakir,

O.;

Akkaya,

E.

U.

Tetrastyryl-BODIPY-Based Dendritic Light Harvester and Estimation of Energy Transfer Efficiency. Org. Lett. 2012, 14, 3636-3639. 30. 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 D.01, Gaussian, Inc.: Wallingford, CT, 2010. 31. Marcus, R. A.; Sutin, N. Electron Transfers in Chemistry and Biology. Biochim. Biophys. Acta 1985, 811, 265−322. 32. Wang,

F.;

Ziegler,

T.

A

Simplified

Relativistic

Time-Dependent

Density-Functional Theory Formalism for the Calculations of Excitation Energies Including Spin-Orbit Coupling Effect. J. Chem. Phys. 2005, 123, 154102. 33. Kang, H.; Si, Y.; Liu, J.; Chen, L.; Li, Y.; Chen, H.; Groeper, J.; Yang, B. An Experimental

and

Theoretical

Study

of

Dimethylaminostyryl

BODIPY-Perylenetetracarboxylic Derivative Dyads: Synthesis, Properties and DFT Calculation. RSC Adv. 2016, 6, 23094-23101. 34. Burgoyne, A. R.; Meijboom, R. Knoevenagel Condensation Reactions Catalysed by Metal-Organic Frameworks. Catal. Lett. 2013, 143, 563-571. 35. Zhao, Y.; Lv, X.; Liu, Y.; Liu, J.; Zhang, Y.; Shi, H.; Guo, W. The Emission Enhancement of the NIR Distyryl BODIPY Dyes by the Indirect S0→S2 19

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 24

Excitation and their Application towards a Hg2+ Probe. J. Mater. Chem. 2012, 22, 11475-11478. 36. Kubin, R. F.; Fletcher, A. N. Fluorescence Quantum Yields of Some Rhodamine Dyes. J. Lumin. 1982, 27, 455-462. 37. Xiao, L.; Wang, H.; Gao, K.; Li, L.; Liu, C.; Peng, X.; Wong, W. Y.; Wong, W. K.; Zhu, X. A-D-A Type Small Molecules Based on Boron Dipyrromethene for Solution-Processed Organic Solar Cells. Chem. Asian. J. 2015, 10, 1513-1518. 38. Shoji, Y.; Ikabata, Y.; Wang, Q.; Nemoto, D.; Sakamoto, A.; Tanaka, N.; Seino, J.; Nakai, H.; Fukushima, T. Unveiling a New Aspect of Simple Arylboronic Esters: Long-Lived Room-Temperature Phosphorescence from Heavy-Atom-Free Molecules. J. Am. Chem. Soc. 2017, 139, 2728-2733. 39. Yang,

Y.;

Guo,

Q.;

Chen,

H.;

Zhou,

Z.;

Guo,

Z.;

Shen,

Z.

Thienopyrrole-Expanded BODIPY as a Potential NIR Photosensitizer for Photodynamic Therapy. Chem. Commun. 2013, 49, 3940-3942. 40. Li, M.; Tian, R.; Fan, J.; Du, J.; Long, S.; Peng, X. A Lysosome-Targeted BODIPY as Potential NIR Photosensitizer for Photodynamic Therapy. Dyes Pigments 2017, 147, 99-105. 41. Adarsh, N.; Shanmugasundaram, M.; Avirah, R.; Ramaiah, D. Aza-BODIPY Derivatives: Enhanced Quantum Yields of Triplet Excited States and the Generation of Singlet Oxygen and their Role as Facile Sustainable Photooxygenation Catalysts. Chem. Eur. J. 2012, 18, 12655-62. 20

ACS Paragon Plus Environment

Page 21 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

42. Lim, S.; Thivierge, C.; Nowak-Sliwinska, P.; Han, J.; van den Bergh, H.; Wagnieres, G.; Burgess, K.; Lee, H. B. In Vitro and in Vivo Photocytotoxicity of Boron Dipyrromethene Derivatives for Photodynamic Therapy. J. Med. Chem. 2010, 53, 2865-2874. 43. Götze, J. P.; Thiel, W. TD-DFT and DFT/MRCI Study of Electronic Excitations in Violaxanthin and Zeaxanthin. Chem. Phys. 2013, 415, 247-255. 44. Jovanović, V.; Lyskov, I.; Kleinschmidt, M.; Marian, C. M. On the Performance of DFT/MRCI-R and MR-MP2 in Spin–Orbit Coupling Calculations on Diatomics and Polyatomic Organic Molecules. Mol. Phys. 2017, 115, 109-137. 45. Lyskov, I.; Kleinschmidt, M.; Marian, C. M. Redesign of the DFT/MRCI Hamiltonian. J. Chem. Phys. 2016, 144, 034104. 46. Ghigo, G.; Roos, B. O.; Malmqvist, P.-Å. A Modified Definition of the Zeroth-Order Hamiltonian in Multiconfigurational Perturbation Theory (CASPT2). Chem. Phys. Lett. 2004, 396, 142-149. 47. Schimmelpfennig B. Atomic Spin-Orbit Mean-Field Integral Program. Stockholms Universitet, Sweden, 1996. 48. Samzow, R.; Hess, B. A. Spin-Orbit Effects in the Br Atom in the Framework of the No-Pair Theory. Chem. Phys. Lett. 1991, 184, 491-496. 49. Heß, B. A.; Marian, C. M.; Wahlgren, U.; Gropen, O. A Mean-Field Spin-Orbit Method Applicable to Correlated Wavefunctions. Chem. Phys. Lett. 1996, 251, 365-371. 21

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

50. Christiansen, O.; Gauss, J.; Schimmelpfennig, B. Spin-Orbit Coupling Constants from Coupled-Cluster Response Theory. Phys. Chem. Chem. Phys. 2000, 2, 965-971. 51. Werner, H., Knowles, P. J., Manby, F. R., Schütz, M., Celani, P., Knizia, G., Adler, T. B. MOLPRO, version 2009.1, A Package of Ab Initio Programs; Molpro, 2009. 52. Liu, Y. X.; Lin, M. S.; Zhao, Y. Intersystem Crossing Rates of Isolated Fullerenes: Theoretical Calculations. J. Phys. Chem. A 2017, 121, 1145−1152. 53. Si, Y.; Liang, W.; Zhao, Y. Theoretical Prediction of Triplet–Triplet Energy Transfer Rates in a Benzophenone–Fluorene–Naphthalene System. J. Phys. Chem. C 2012, 116, 12499-12507. 54. Place, I.; Farran, A.; Deshayes, K.; Piotrowiak, P. Triplet Energy Transfer through the Walls of Hemicarcerands: Temperature Dependence and the Role of Internal Reorganization Energy. J. Am. Chem. Soc. 1998, 120, 12626-12633 55. Si, Y.; Yang, B.; Qin, H.; Yuan, J.; Wang, S.; Chen, H.; Zhao, Y. Atomistic Modeling of Triplet–Triplet Energy-Transfer Rates from Drug (S)-Propranolol to (R)-Cinacalcet in Human α1-Acid Glycoprotein. J. Phys. Chem. C 2015, 119, 8014-8022. 56. Vura-Weis, J.; Newton, M. D.; Wasielewski, M. R.; Subotnik, J. E. Characterizing the Locality of Diabatic States for Electronic Excitation Transfer by Decomposing the Diabatic Coupling. J. Phys. Chem. C 2010, 114, 20449-20460. 22

ACS Paragon Plus Environment

Page 22 of 24

Page 23 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

57. Zhu, S.; Zhang, J.; Vegesna, G.; Luo, F. T.; Green, S. A.; Liu, H. Highly Water-Soluble Neutral BODIPY Dyes with Controllable Fluorescence Quantum Yields. Org. Lett. 2011, 13,438-441.

23

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC Graphic

24

ACS Paragon Plus Environment

Page 24 of 24