A Comprehensive Scope of Peripheral and Axial Substituent Effect on

Email: [email protected]; [email protected]. ¶ - these two authors contributed equally to the preparation of this manuscript and the ..... compound...
0 downloads 5 Views 3MB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

A: Spectroscopy, Photochemistry, and Excited States

A Comprehensive Scope of Peripheral and Axial Substituent Effect on the Spectroelectrochemistry of Boron Subphthalocyanines Kathleen L. Sampson, Xiaoqin Jiang, Esmeralda Bukuroshi, Aleksa Dovijarski, Hasan Raboui, Timothy P Bender, and Karl M. Kadish J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b02023 • Publication Date (Web): 19 Apr 2018 Downloaded from http://pubs.acs.org on April 19, 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 44 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

A Comprehensive Scope of Peripheral and Axial Substituent Effect on the Spectroelectrochemistry of Boron Subphthalocyanines. Kathleen L. Sampson,a,¶ Xiaoqin Jiang,b,¶ Esmeralda Bukuroshi,a Aleksa Dovijarski,a Hasan Raboui,a Timothy P. Bender,*,a,c,d and Karl M. Kadish*,b

a

Department of Chemical Engineering and Applied Chemistry, University of Toronto, 200

College Street, Toronto, Ontario, Canada M5S 3E4. b

Department of Chemistry, University of Houston, Houston, TX 77204-5003, USA

c

Department of Chemistry, University of Toronto, 80 St George Street, Toronto, Ontario, Canada

M5S 3H6. d

Department of Materials Science and Engineering, University of Toronto, 184 College Street,

Toronto, Ontario, Canada, M5S 3E4. * - to whom correspondences should be addressed. Email: [email protected]; [email protected]. ¶ - these two authors contributed equally to the preparation of this manuscript and the outlined experimentation.

Page 1 of 37 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

ABSTRACT An extensive study of the electrochemical and spectroelectrochemical properties of 14 boron subphthalocyanine (BsubPc) derivatives with various axial and peripheral substituents was performed in 1,2-dichloromethane (CH2Cl2) containing 0.1 M tetra-n-butyl-ammonium perchlorate (TBAP) as the supporting electrolyte. From the cyclic voltammetry results, all compounds exhibit one oxidation and at least two reduction processes within the solvent potential window of +1.6 to -1.8 V vs. SCE. It was found that the reversibility of the redox reactions depends on the axial and peripheral substituents and the dipole moment of the boronto-axial substituent. In general, UV-visible absorption spectra of the singly reduced BsubPc derivatives exhibit three equal-intensity peaks in the 450 to 650 nm region that are derived from the maximum BsubPc absorbance peak upon reduction. Axial substituents affect the intensity of the three peaks upon reduction, while peripheral substituents shift the position of the peaks to higher wavelengths. Upon oxidation, the UV-visible absorption profile flattens considerably with only a single broad (~300 nm) band apparent. Understanding the effect of substituents on the stability of the redox processes of BsubPcs will aid in further development of these materials for applications in organic electronic devices.

Page 2 of 37 ACS Paragon Plus Environment

Page 2 of 44

Page 3 of 44 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

INTRODUCTION Boron subphthalocyanines (BsubPcs) are semiconducting organic small molecules that are a variant from the widely studied family of phthalocyanine (Pc) macrocycles. BsubPcs are uniquely templated and formed by cyclotrimerization of phthalonitrile in the presence of boron trihalides (BCl3 and BBr3). Because of their unique photophysical properties, BsubPcs are applicable within organic electronic devices, such as organic photovoltaics (OPVs) and organic light emitting diodes (OLEDs).1-5 Within OPVs, BsubPcs have been proven to have multiple roles including facilitating exciton formation through strong photonic absorption6-7 and exciton dissociation through both hole-8-9 and electron-transport.10-11 The energy level alignment of the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) of organic semiconductor materials has an intrinsically important role in converting (sun)light into electrical current in OPV devices. The dissociation of excitons (electron-hole pairs) generated within the active layer of an OPV depends on the spacing between the HOMO and the LUMO of the electron-donating and electron-accepting materials.5, 12-13 Upon dissociation of the excitons, the active material either becomes a radical cation (electron removal) or a radical anion (electron addition). Therefore, an understanding of the radical cation and radical anion stability is desirable for a material. The ability to determine and alter the HOMO/LUMO energy levels is also desirable and essential to the engineering of efficient charge transfer and OPV device performance. Several handles are available to alter the energy levels of BsubPcs, two of which are the ability to change the axial ligands on the boron center atom or the substituents on the peripheral positions of the subPc macrocycle.14-17 Morse et al., have shown that the redox potentials of BsubPcs can be altered more significantly by changing the peripheral rather than axial substituents.14 Thus, the ability to

Page 3 of 37 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

simultaneously characterize the redox behavior and the stability of BsubPcs is of particular importance to confirm the alignment of the energy levels, to efficiently extract charge, and to obtain a radical anion or a radical cation over long periods of time. The electrochemical, spectroscopic, and spectroelectrochemical properties of Pcs have previously been measured to determine the effect of both the macrocycle and its substituents on the electrochemical behavior of the compounds.18-22 Both the number and potential of redox reactions of Pcs can be affected by the substituents on the Pc macrocycle.23 While electrochemical techniques, such as cyclic voltammetry (CV) and differential pulse voltammetry (DPV), have been used extensively to study the electrochemical properties of BsubPcs,14, 24 only a few comprehensive studies have compared a wide range of BsubPc derivatives under the same solution and experimental conditions (solvent, supporting electrolyte, type of electrode material, and type of electrochemical experiment). Moreover, most of these studies were on BsubPcferrocene dyads and not on the individual monomeric units.25-28 The first oxidation of most BsubPcs is typically irreversible and the product is less stable than that of the first reduction, which is often a reversible process.14, 24-27, 29 Scanning to more positive or more negative potentials beyond the first oxidation or first reduction often reveals several additional redox reactions.26-27, 30 These are not necessarily related to starting compounds, but may instead be due to redox reactions involving new products generated from the singly oxidized or singly reduced BsubPcs. Understanding the stability and reversibility of the redox species and identification of side reactions is, therefore, essential for evaluating the potential application of these materials in organic electronics. Spectroelectrochemistry is a useful technique to simultaneously determine the redox stability and to spectrally characterize both

Page 4 of 37 ACS Paragon Plus Environment

Page 4 of 44

Page 5 of 44 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

intermediates and final products of the redox reactions, which is important for organic electronic device performance. However, to date, it has been used only once to study BsubPcs.28 In this study, a series of 14 BsubPc derivatives with different axial and peripheral substituents

were

characterized

using

cyclic

voltammetry

and

UV-visible

spectroelectrochemistry (Figure 1). The BubPcs with differing axial substituents include the prototypical chloro-BsubPc (Cl-BsubPc, 13) and the less explored fluoro-BsubPc (F-BsubPc, 12).31-33 We also included in our study an extensive set of phenoxylated-BsubPcs: 3methylphenoxy-BsubPc

(3-MePhO-BsubPc,

8),

phenoxy-BsubPc

(PhO-BsubPc,

1),

pentafluorophenoxy-BsubPc (F5BsubPc, 7), and five additional fluorophenoxy-BsubPc compounds. These fluorophenoxy-BsubPcs include 3-fluorophenoxy-BsubPc (3F1PhO-BsubPc, 2),

3,5-difluorophenoxy-BsubPc

(35F2PhO-BsubPc,

3),

2,4,6-trifluorophenoxy-BsubPc

(246F3PhO-BsubPc, 4), 3,4,5-trifluorophenoxy-BsubPc (345F3PhO-BsubPc, 5) and 2,3,5,6tetrafluorophenoxy-BsubPc (2356-F4PhO-BsubPc, 6). In order to elucidate the influence of peripheral substituents, chloro-hexachloro-BsubPc (Cl-Cl6BsubPc, 14) and phenoxy-hexachloroBsubPc (PhO-Cl6BsubPc, 9) were chosen as part of this study due to their known functionality in OPVs.34-38 Finally, phenoxy-dodecachloro-BsubPc (PhO-Cl12BsubPc, 10) was also included based on previous observations that chloro-dodecachloro-BsubPc (Cl-Cl12BsubPc) harvested triplets from pentacene in OPVs.37 Cl-Cl12BsubPc has low solubility, but its phenoxylated version was easily characterized by spectroelectrochemistry.

Page 5 of 37 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 1. Structures of the investigated BsubPcs. EXPERIMENTAL SECTION Spectroelectrochemistry All chemicals and solvents were of the highest grade available and were used without further purification. Dichloromethane (CH2Cl2) and tetra-n-butyl-ammonium perchlorate (TBAP), used as the supporting electrolyte, were purchased from Sigma Aldrich Co. UV-visible spectra were recorded with a Hewlett-Packard Model 8453 diode array spectrophotometer. Cyclic voltammetry was carried out at 298 K using an EG&G Princeton Applied Research 173 potentiostat/galvanostat. A three-electrode system was used for cyclic voltammetry measurements and consisted of a glassy carbon working electrode, a platinum counter electrode, and a saturated calomel reference electrode (SCE). The SCE was separated from the bulk of the solution by a fritted glass bridge of low porosity which contained the solvent/supporting electrolyte mixture. Thin-layer UV-visible spectroelectrochemical experiments were performed with a home-built thin-layer cell, which had a light transparent platinum-net working Page 6 of 37 ACS Paragon Plus Environment

Page 6 of 44

Page 7 of 44 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

electrode.39-40 Potentials were applied and monitored with an EG&G PAR Model 173 potentiostat. High purity nitrogen from Trigas was used to deoxygenate the solution and a stream of nitrogen was kept over the solution during each electrochemical and spectroelectrochemical experiment. Synthesis of Boron Subphthalocyanines All solvents used in the synthetic procedures and the 50-200 µm standard basic alumina for Kauffman chromatography were purchased from Caledon Laboratories (Caledon, Ontario, Canada) and used as received. The procedures and setup of the Kauffman column and train sublimation apparatus have been previously described.41 3-Methylphenol and phenol for the synthesis of 3-MePhO-BsubPc (8) and PhO-Cl12BsubPc (10) were purchased from Sigma Aldrich (Oakville, Ontario, Canada). High performance liquid chromatography (HPLC) was performed by using a Waters 2695 separation module with a Waters 2998 photodiode array and a Waters 4.6 mm x 100 mm SunFire C18 3.5 µm column. The HPLC mobile phase was composed of 80/20 acetonitrile/dimethylformamide solvent mixture (HPLC solvents purchased from Caledon Laboratories, used as received) and eluted at a rate of 0.6 mL·min-1. Extinction coefficients were measured using our previously described method and are summarized in Table S1 of the Supporting Information.7, 31 The synthetic procedures for PhO-BsubPc (1), 3F1PhO-BsubPc (2), 35F2PhO-BsubPc (3), 246F3PhO-BsubPc (4), 345F3PhO-BsubPc (5), 2356F4PhO-BsubPc (6),42 F5BsubPc (7),41 PhBsubPc (11),43 F-BsubPc (12),31 Cl-BsubPc (13),7 and Cl-Cl6BsubPc (14)14 have been previously described. 3-Methylphenoxy-boron subphthalocyanine (3-MePhO-BsubPc, 8)

Page 7 of 37 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

Br-BsubPc was prepared based on the method previously reported44 by modifying a published procedure by Potz et al.32 A similar procedure to the one previously described by Paton et al. was followed.17 Br-BsubPc was used as a starting material instead of Cl-BsubPc to decrease the reaction time. It has been shown in our lab that Br-BsubPc has a higher rate of substitution with phenol than Cl-BsubPc.31 To a pre-dried round bottom flask with a condenser and under argon gas pressure, Br-BsubPc (5.0 g, 10.52 mmol) was dissolved in 75 mL of toluene. After five minutes of stirring, two equivalents of 3-methylphenol (2.28 g, 21.05 mmol) was added to the reaction flask and the mixture was refluxed at 111 °C overnight. The next day, the heating was stopped and the mixture was allowed to cool to room temperature. The crude product was purified by first dissolving the mixture in 160 mL of toluene and washing three times with 300 mL of 3M KOH in a separatory funnel, keeping the organic layer. Toluene was removed by rotary evaporation and further purified, first by Kauffman chromatography using standard basic alumina and dichloromethane as the solvent and then by train sublimation. Reaction Yield 3.02 g, 47.7%. Sublimation Yield 1.47 g, 48.5%. HPLC purity 98%, 1H NMR (400 MHz, CDCl3, Me4Si): δ1.94 (3H, s), 5.12 – 5.14 (1H, m), 5.25 (1H, s), 6.42 – 6.43 (1H, d), 6.63 – 6.65 (1H, t), 7.88 – 7.93 (6H, m), 8.83 – 8.88 (6H, m). MS (ESI) m/z: 502.2 [M + H]+. Phenoxy-hexachloro-boron subphthalocyanine (PhO-Cl6BsubPc, 9) A more time-efficient synthetic procedure for the BsubPc acceptor used in this study (PhO-Cl6BsubPc) was adapted from Ebenhoch et al., based on our synthetic methodology.31, 34, 45 Rather than refluxing phenol with Cl-Cl6BsubPc in toluene for 7 days, we have previously found that bromine in the axial position of BsubPc is more labile than chlorine and, thus, undergoes a faster reaction. Therefore, bromo-hexachloro-boron-subphthalocyanine (Br-Cl6BsubPc) was first synthesized. To an oven-dried 3-neck round bottom flask equipped with a condenser and gas

Page 8 of 37 ACS Paragon Plus Environment

Page 8 of 44

Page 9 of 44 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

inlet, dichlorophthalonitrile (9.654 g, 49 mmol) was added to bromobenzene (14.5 mL) and toluene (40.5 mL) under argon gas pressure. Boron tribromide (1.5 mL, 15.8 mmol) was added slowly through one neck of the reaction flask. The reaction was stirred overnight. In the morning, the reaction was turned off and left for 1 hour. The mixture was filtered by gravity filtration and immediately washed with ~75 mL of acetonitrile. The solid was dried with nitrogen blowing over top for a few hours. Once sufficiently dry, the product was placed in the vacuum oven overnight. Yield 7.74 g, 70.4 %. HPLC purity 89 %, MS (ESI) m/z: 681.8 [M + H]+. One equivalent of Br-Cl6BsubPc (0.8 g, 1.2 mmol) was combined with five equivalents of phenol (0.55 g, 5.9 mmol) in chlorobenzene (15 mL) in a 50 mL round-bottom flask fitted with a condenser and held under a constant pressure of argon gas. The reactants were heated at reflux for 4 hours, at which time the reaction was confirmed to be complete by HPLC analysis (absence of Br-Cl6BsubPc) and was subsequently cooled to room temperature. The solvent was removed by rotary evaporation, yielding the crude product. The crude product was purified using Kauffman column chromatography with silica gel and dichloromethane as the eluent. Removal of dichloromethane by rotary evaporation yielded the product, which was further purified by train sublimation. HPLC purity > 99.9%. 1H NMR (400 MHz, CDCl3, Me4Si) δ 5.34 – 5.36 (2H, m), δ 6.65 – 6.75 (1H, t), δ 6.76 – 6.79 (2H, m), δ 8.90 (6H, s). MS (ESI) m/z: 693.9 [M + H]+. Phenoxy-dodecachloro-boron subphthalocyanine (PhO-Cl12BsubPc, 10) A method adapted from Morse and Bender was used to synthesize PhO-Cl12BsubPc, the main difference being a new activating agent, tin(IV) chloride.46 Cl-Cl12BsubPc (1.026 g, 1.22 mmol, 1 eq) was added to a round bottom flask containing 1,2-dichlorobenzene (51 mL). To the stirring solution was added tin chloride (0.213 mL, 1.82 mmol, 1.5 eq). The resulting mixture was heated at 60 °C under argon for 25 minutes. After 25 minutes, phenol (0.572 g, 6.08 mmol, 5 Page 9 of 37 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

eq) was added and the reaction mixture was left under heating at 60 °C for 27 hours. The mixture was quenched with pyridine (10 mL). After cooling to room temperature, the solvent and other volatiles were removed by a rotary evaporator. The crude product was purified on a Kauffman column using standard silica gel as the adsorbent and dichloromethane as the eluent. The product eluted from the Kauffman column, while most of the excess phenol remained adsorbed. Dichloromethane was then removed by a rotary evaporator and the product was further purified by train sublimation with a yield of 0.064 g (24%). HPLC purity > 99%. 1H NMR (400 MHz, CDCl3, Me4Si): δ 5.35 – 5.37 (2H, d), 6.70 – 6.77 (1H, t), 6.81 – 6.78 (2H, t). MS (EI) m/z: 901.7 [M]+.

RESULTS AND DISCUSSION The main goal of this study was to survey the spectroelectrochemistry of boron subphthalocyanines (BsubPcs), a previously unexplored analysis technique applied broadly to these compounds. To start, we first performed cyclic voltammetry (CV) measurements, on the series of 14 BsubPc derivatives with various axial and peripheral substituents (Figure 1) in order to measure the redox potentials and evaluate reversibility. This was followed by spectroelectrochemical measurements at applied potentials which were at least 100 mV beyond the peak or half-wave potential of the investigated redox processes. In this study, it was assumed that the solubility does not significantly change between the neutral, reduced, and oxidized species of each compound during the time of the cyclic voltammetric or spectroelectrochemical measurements. Our discussion of the electrochemical and spectroelectrochemical data is divided into two main sections based on the effect of the axial and peripheral substituents on the redox behavior. Table 1 summarizes the peak or half-wave oxidation and reduction potentials for each Page 10 of 37 ACS Paragon Plus Environment

Page 10 of 44

Page 11 of 44 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

compound and Table 2 lists the maximum absorption wavelengths for major peaks of the neutral, anionic, and cationic forms of the BsubPcs.

Table 1. Peak and half-wave potentials (Ep and E1/2, V vs. SCE) of investigated BsubPcs in

Compound PhO-BsubPc (1) 3F1PhO-BsubPc (2) 35F2PhO-BsubPc (3) 246F3PhO-BsubPc (4) 345F3PhO-BsubPc (5) 2356F4PhO-BsubPc (6) F5BsubPc (7) 3-MePhO-BsubPc (8) PhO-Cl6BsubPc (9) PhO-Cl12BsubPc (10) Ph-BsubPc (11) F-BsubPc (12) Cl-BsubPc (13) Cl-Cl6BsubPc (14) CH2Cl2, 0.1 M TBAP. a

Oxidation Ep (ox 1) E1/2 (ox 1) 1.09 1.10 1.13 1.12 1.15 1.14 1.15 1.12 1.31 1.44 1.05 1.10 1.14 1.37 -

Ep (red 1) -0.99 -0.72

Quasi-reversible redox reaction

Page 11 of 37 ACS Paragon Plus Environment

Reduction E1/2 (red 1) Ep (red 2) -1.06 -1.57 -1.04 -1.57 -1.03 -1.57 -1.04 -1.58 -1.02 -1. 56 -1.01 -1.55 -1.01 -1.58 -1.06 -1.62 -0.77 -0.54 -1.11 -1.00 -1.48 -1.22

E1/2 (red 2) -1.31 -1.09 -1.60a -1.53a -

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 12 of 44

Table 2. UV-visible Spectral Data of Neutral Species, Anion Radicals, and Cation Radicals of Some of the Investigated BsubPcs in CH2Cl2 Containing 0.1 M TBAP. Compound

Neutral PhO-BsubPc (1) 307 513sh F5BsubPc (7) 307 515sh 3-MePhO-BsubPc (8) 304 518sh PhO-Cl6BsubPc (9) 318 520sh PhO-Cl12BsubPc (10) 328 539sh Ph-BsubPc (11) 308 521sh F-BsubPc (12) 306 513sh Cl-BsubPc (13) 302 502sh Cl-Cl6BsubPc (14) 314 525sh Sh = shoulder peak, br = broad peak

561 561 562 568 586 566 560 563 570

301 301 300 309 326 307 300 -

λmax (nm) λ Reduced (nm) 483 513 566 483 515 567 488 518 565 489 534 606 505 539 613 487 521 584 484 512 566 502 552 -

600-800br 600-800br 670 720 703 690 657 657 705

Oxidized 321 350 321 365 325 364 350 348 400 317 522 366 322 322 -

*Absorption peaks for fluorophenoxy-BsubPcs (2 – 6) from spectroelectrochemical measurements can be found in Table S2 of the Supporting Information Effect of Axial Substituents Electrochemistry In terms of axial substituents, we analyzed a series of axially phenoxy-BsubPcs, including PhO-BsubPc (1), a range of fluorophenoxy-BsubPcs (3F1PhO-BsubPc, 2; 35F2PhOBsubPc, 3; 246F3PhO-BsubPc, 4; 345F3PhO-BsubPc, 5; 2356F4PhO-BsubPc, 6; and F5BsubPc, 7),

and

a

m-methyl-phenoxy-BsubPc

(3-MePhO-BsubPc,

8).

Representative

cyclic

voltammograms are depicted in Figure 2 and comprehensive set of the voltammograms are depicted in Figure S1. We also examined phenyl-BsubPc (Ph-BsubPc, 11) and two halogenatedBsubPcs (F-BsubPc, 12, and Cl-BsubPc, 13). Each BsubPc derivative undergoes one oxidation and at least two reductions within the potential limit of the electrochemical solvent (+1.6 to -1.8 V vs. SCE), a result which is similar to other electrochemistry results reported in the literature (Figure 2).26-27, 29-30, 43

Page 12 of 37 ACS Paragon Plus Environment

Page 13 of 44 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 2. Cyclic voltammograms of axially-substituted BsubPcs in CH2Cl2 containing 0.1 M TBAP (PhO-BsubPc, 1; Ph-BsubPc, 11; F-BsubPc, 12; and Cl-BsubPc, 13).

A comparison of voltammograms for the eight investigated phenoxy-substituted BsubPcs (Figure S1) shows only a very slight variation in redox potentials. This can be explained from previous reports that explored modelling of the molecular orbitals for a series of BsubPcs and found that the first one-electron oxidation and first one-electron reduction occur at the π−ring Page 13 of 37 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

system of the conjugated subphthalocyanine macrocycle.47-50 Thus, axial substituents do not have much effect on the HOMO and LUMO values of the BsubPc. In our previous study of the fluorophenoxy-BsubPcs (1 – 7), we found a trend in this small variation which was due to the large electronic influence of fluorine in the meta position of the fluorophenoxy axial substituent.42 However, the cyclic votammograms in this study were done without an internal standard and the redox potentials, therefore, have an error of 10 – 25 mV. Thus, the same trends cannot be observed from the electrochemical data in which the redox potentials vary slightly from compound to compound. Our previous study used multiple other techniques, such as differential pulse voltammetry, ultraviolet photoelectron spectroscopy, semi-empirical modelling, and OPV device characteristics with the fluorophenoxy-BsubPcs as electron acceptors, in order to firmly establish the influence of a fluorine in the meta position of the phenoxy axial substituent and the resulting effect of this substituted placement on OPV device performance.42 Redox potentials of the other axially-substituted BsubPcs (Ph-BsubPc, 11; F-BsubPc, 12; and Cl-BsubPc, 13) are also fairly similar to the axial-phenoxy-substituted BsubPcs. However, the potentials of Ph-BsubPc (11) are slightly more negatively shifted than the other axiallysubstituted BsubPcs, outside of the ~25 mV error of these measurements. This could be due to the small dipole moment of the boron to axial substituent bond, which is described in more detail on the following pages. In this study, we also scanned to more negative potentials in order to observe the presence of a second reduction process. The second reduction involves a one electron transfer and was seen for all of the axially-substituted BsubPcs. This reduction is also coupled to a new irreversible oxidation peak between -0.35 and -0.48 V on the return scan back towards 0 V (an exception is Cl-BsubPc, 13). The new oxidation peak may involve a new chemical species Page 14 of 37 ACS Paragon Plus Environment

Page 14 of 44

Page 15 of 44 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

generated from the doubly reduced BsubPc, but further studies of compound identification of this product was not carried out and is of secondary interest to this study. While redox potentials of the axially-substituted BsubPcs are comparable to each other, the reversibility of these processes depends upon the type of axial substituent. For the phenoxyBsubPcs (1 – 8), the first reduction is reversible, while the second reduction and first oxidation processes are both irreversible. This result is consistent with what was previously reported for these phenoxy-BsubPc derivatives.14,

42

F-BsubPc (12) exhibits a reversible one electron

reduction and an irreversible one electron oxidation at similar potentials, which are similar to both PhO-BsubPc (1) and the fluorophenoxy-BsubPcs (2 – 7). However, the first reduction is less reversible for F-BsubPc (12), a compound with a fluorine axial ligand, than for the compounds with phenoxy axial ligands. In addition, the second reduction becomes quasireversible for F-BsubPc (12). When the BsubPc is axially coordinated with a phenyl group (PhBsubPc, 8), the first oxidation becomes reversible, a rare result for BsubPcs, and the second reduction becomes quasi-reversible. An exception to this trend of two reductions with the first being reversible is for Cl-BsubPc (13), where all of the redox reactions are irreversible and there are more than two reduction processes. The irreversible first reduction at -0.99 V is followed by two other irreversible peaks at -1.48 and -1.70 V as shown in Figure 2. The peak at -1.70 V is coupled to a reoxidation peak on the return scan at -1.48 V. It is difficult to determine the nature of these peaks and whether they are separate reduction and reoxidation processes of Cl-BsubPc (13) or if a new species is formed after reduction. This unclear cyclic voltammetry curve could be due to the known lower solubility of Cl-BsubPc (13) compared to other axially-substituted BsubPcs.7, 17, 31

Page 15 of 37 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

A previous theoretical study by Ferro et al. using semi-empirical, density functional theory (DFT), and Hartree-Fock (HF) molecular modeling methods, found that the boron-toaxial ligand bond length varies with the axial substituent type and obeys the following sequence: F-BsubPc < PhO-BsubPc < Ph-BsubPc < Cl-BsubPc.47 This order of boron-to-ligand bond length is the same order determined from X-ray diffraction measurements in the literature.17, 31, 43, 51 However, this order does not correlate well with the reversibility of redox potentials, with PhBsubPc having the second longest boron-to-axial ligand bond, but the most reversible redox processes and the most stable products of the electron transfer. The reversibility of the redox potentials seems to coincide more closely to the dipole moment of the axial ligand as measured with the same DFT calculations performed by Ferro et al., which results in the following order: Ph-BsubPc (3.1 D) < F-BsubPc (4.4 D) ~ PhO-BsubPc (4.5 D) < Cl-BsubPc (5.5 D).47 The more electron withdrawing the axial substituent and the longer the boron-to-axial ligand bond, the higher the dipole moment and the more difficult it is to reversibly oxidize or reduce the compound. For example, Ph-BsubPc (11) has the lowest dipole moment and the only studied compound with a reversible oxidation, a reversible first reduction, and a quasi-reversible second reduction. F-BsubPc (12) and PhO-BsubPc (1) have similar dipole moments and comparable electrochemistry with an irreversible oxidation, reversible first reduction, and an irreversible or quasi-reversible second reduction. Finally, Cl-BsubPc (13) has the bulkiest and most electronegative axial substituent, resulting in the largest dipole moment and all irreversible redox events. From the DFT study on the redox processes of Cl-BsubPc (13), Ferro et al. found that the boron-to-axial substituent dipole moment increases with oxidation and the boron-to-axial substituent bond length changes more with the oxidized species than the reduced species.48 This Page 16 of 37 ACS Paragon Plus Environment

Page 16 of 44

Page 17 of 44 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

may contribute to the reversibility of the redox processes.48 Further studies on the change in boron-to-axial substituent bond length and dipole moment with reduction and oxidation of other axially substituted BsubPcs is needed in order to further understand this mechanism and the trends in stability. Spectroelectrochemistry UV-visible spectra were recorded during controlled-potential reduction or oxidation of the axially-substituted BsubPc derivatives. As predicted from the similar electrochemistry of the fluorophenoxy-BsubPc derivatives (1 – 7) described above, Figure S2 depicts the comparable neutral UV-visible spectra, which are characterized by a major absorption band between 561 – 562 nm, shoulder peak in the range of 513 – 517 nm, and a second peak at 307 nm of lower intensity and extinction coefficient. During the one-electron addition to these fluorophenoxyBsubPcs (1 – 7), the major absorption band between 561 – 562 nm decreases in intensity and is replaced by three major absorption bands of fairly equal intensity at 483 – 488, 513 – 517, and 563 – 569 nm. The peak at 307 nm also decreases in intensity and shifts to 301 – 305 nm as a new broad band in the range of 600 – 800 nm forms. Similar changes are seen for 3-MePhOBsubPc (8), which has bands before reduction at 304, 518, and 562 nm as compared to 300, 488, 518, 565 nm, and 670 nm after the addition of one electron (Figure S3). The UV-visible spectra are also similar to that of the eight singly-oxidized axial phenoxysubstituted BsubPcs. Examples of the spectral changes are given in Figure S3, which depicts the first reduction and oxidation spectra of F5BsubPc (7) and 3-MePhO-BsubPc (8), compounds representative of the axial phenoxy-substituted BsubPcs. In general, the major peak at ~560 nm flattens considerably to the baseline upon the one-electron abstraction. For 3-MePhO-BsubPc (8), a peak at ~600 nm appears upon oxidation but flattens to the baseline with further scans. The Page 17 of 37 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

peak at ~305 nm decreases in intensity and broadens with two ill-defined peaks at ~320 – 325 nm and ~365 nm. The absorption profile of the oxidized species is considerably more altered compared to the changes in spectra for the reduced species. This may indicate that the oxidized compound is less stable than the reduced compound. However, further studies of the spectroelectrochemical reversibility are needed to confirm this. Changing the axial ligand on Ph-BsubPc (11), F-BsubPc (12), and Cl-BsubPc (13), has very little effect on the spectra of the reduced or oxidized compounds (Figure 3). For the neutral species, these BsubPcs exhibit two major absorption bands: a major peak between 560 – 566 nm with a shoulder between 502 – 521 nm and a second lower intensity band in the range of 302 – 308 nm; these are typical bands of BsubPcs.7, 14, 29, 31, 43 However, the presence of the phenyl and chlorine axial ligands also leads to a moderate red-shift in the absorption band, which are located at 566 and 564 nm, respectively, as compared to 561 nm for PhO-BsubPc (1) and 560 nm for FBsubPc (12). Similar UV-visible spectra are also observed for the anion radicals of Ph-BsubPc (11) and PhO-BsubPc (1), each compound being characterized by three major bands of similar intensity between 400 – 650 nm and a newly formed broad band between 600 – 700 nm upon reduction. The three absorption bands for the anionic form of Ph-BsubPc (11) are located at 487, 521, and 584 nm, slightly red-shifted from the bands of PhO-BsubPc (1) at 483, 513, and 566 nm. In the case of the fluoride and chloride derivatives, the bands in the 400 – 650 nm region are flattened to two non-obvious bands at 485 and 512 nm for F-BsubPc (12) and further levelled out to one broad band at around 500 nm for Cl-BsubPc (13) after reduction. In addition, the broad band in the 600 – 700 nm region is more pronounced with these compounds and there is a peak at around 657 nm (Figure 3). This result is fairly consistent with the electrochemical data in Page 18 of 37 ACS Paragon Plus Environment

Page 18 of 44

Page 19 of 44 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 2, in which the first reduction is less reversible for F-BsubPc (12) and irreversible for ClBsubPc (13). Again, the reason for this is attributed to the large dipole moment of the boron-tohalogen axial substituent and the resulting stretching of this bond during the reduction processes. As seen in Figure 3, the major 560 – 566 nm peak of the neutral compounds significantly flattens to the baseline upon oxidation. The peak at ~ 300 nm also decreases in intensity and broadens, showing a maximum between 317 – 366 nm for Ph-BsubPc (11), F-BsubPc (12), and Cl-BsubPc (13). As the ~560 nm peak of these three axially-substituted compounds decreases, a small new absorption is initially observed at ~600 nm, but with further scans, this new peak also reduces in intensity. Future work in analyzing the reversibility of the first reduction and oxidation processes by spectroelectrochemistry would be useful in establishing the stability of the anion and cation species. Further investigations of the spectra and reversibility during the second reduction would also contribute to our understanding of the overall stability of BsubPcs.

Page 19 of 37 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 3. UV-visible spectral changes during controlled-potential first reduction (left spectra) and oxidation (right spectra) for PhO-BsubPc (1), Ph-BsubPc (11), F-BsubPc (12), and ClBsubPc (13) in CH2Cl2 containing 0.1 M TBAP as the electrolyte.

Page 20 of 37 ACS Paragon Plus Environment

Page 20 of 44

Page 21 of 44 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

Effect of Peripheral Substituents Electrochemistry As reported previously, the redox behavior of BsubPcs can be altered much more significantly by changing the peripheral substituents than by changing the axial ligands.14 To confirm this and to observe the reversibility of the reduction and oxidation for the currently investigated compounds, the BsubPc macrocycle was altered by the stepwise addition of six peripheral chlorine atoms to both PhO-BsubPc (1) and Cl-BsubPc (13), forming PhO-Cl6BsubPc (9), PhO-Cl12BsubPc (10), and Cl-Cl6BsubPc (14), respectively. The substantial effect of the peripheral electron-withdrawing chlorine substituents is depicted in the cyclic voltammograms of Figure 4. The addition of six chlorine groups to the macrocycle (PhO-BsubPc, 1, to PhO-Cl6BsubPc, 9) shifts the oxidation peak potential, the first reduction half-wave potential, and the second reduction peak potential by 220, 290, and 200 mV, respectively. These peak or half-wave potentials are further shifted positively by 170, 230, and 220 mV, respectively, when six more chlorine atoms are added to the macrocycle (comparing PhO-Cl6BsubPc, 9, and PhO-Cl12BsubPc, 10). Moreover, the second reduction becomes reversible for PhO-Cl6BsubPc (9) and the first oxidation and the two reductions are reversible for PhOCl12BsubPc (10). Thus, the electron-withdrawing chlorine peripheral substituents have two effects: they shift the redox potentials in a positive direction, making the two reductions easier and single oxidation harder, and they also cause the redox reactions to become more reversible. With peripheral substituents, the HOMO and LUMO energy levels of the BsubPcs can be deepened to be used as electron accepting materials in OPVs.3, 5, 34, 36 The increased redox reversibility is likely related to how the electron-withdrawing peripheral substituents have been theoretically shown to

Page 21 of 37 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

shorten the boron-to-axial substituent bond distance as well as significantly lower the dipole moment, thus, possibly leading to more electrochemically reversible species.50 The newly formed irreversible oxidation peak at ~ -0.3 to -0.5 V that appears after the second reduction process for the axially-substituted BsubPcs is not observed for PhO-Cl6BsubPc (9) or PhO-Cl12BsubPc (10). This is another indication for an improved stability of the compound or the anionic species generated with peripheral substitution. The result of peripheral substituents having a greater effect on the redox processes also correlates with the theoretically shown location of the macrocyclic HOMO orbital.47, 49 As seen in Figure 4, similar redox behavior is observed for Cl-BsubPc (13) and ClCl6BsubPc (14) as the potentials for the first oxidation and reduction peaks are shifted more positively for Cl-Cl6BsubPc (14) by 230 and 270 mV, respectively. The cyclic voltammograms of both BsubPcs have ill-defined peaks and Cl-Cl6BsubPc (14) also has an irreversible reduction at 0.72 V, followed by three additional irreversible peaks of smaller current at -0.87, -1.22, and -1.76 V. As stated previously, the unclear and inconclusive peaks of the chlorine axially-substituted BsubPcs could be due to their known lower solubility.7, 31 However, none of the redox processes of Cl-BsubPc (13) or Cl-Cl6BsubPc (14) are reversible, which could be due to the high dipole moment associated with the chlorine axial substituent.

Page 22 of 37 ACS Paragon Plus Environment

Page 22 of 44

Page 23 of 44 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

The Journal of Physical Chemistry

Figure 4. Cyclic voltammetry of PhO-BsubPc (1), PhO-Cl6BsubPc (9), PhO-Cl12BsubPc (10), Cl-BsubPc (13), and Cl-Cl6BsubPc (14), in CH2Cl2 containing 0.1 M TBAP as the electrolyte.

Page 23 of 37 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

Spectroelectrochemistry The peripheral chlorine substituents on compounds PhO-Cl6BsubPc (9), PhO-Cl12BsubPc (10), Cl-Cl6BsubPc (14) also affect the UV-visible spectra for the neutral, reduced, and oxidized species. As depicted in Figure 5, the successive addition of six chlorine atoms to the periphery of the BsubPc macrocycle shifts the main absorption bands from 307, 513, and 561 nm in the case of PhO-BsubPc (1) to 318, 520, and 568 nm for PhO-Cl6BsubPc (9) and 328, 540, and 586 nm for PhO-Cl12BsubPc (10). The UV-visible spectra of the anion radicals also follow a similar trend. The anion radicals of these three compounds have UV-visible absorption spectra consisting of three major, equal-intensity bands, which are red shifted with increasing number of chlorine peripheral substituents. The formation of three equal-intensity bands is consistent with what is seen for most of the axially-substituted BsubPcs described above. The absorption bands shift from 483, 513, and 566 nm for PhO-BsubPc (1) to 490, 534, and 606 nm for PhOCl6BsubPc (9) and then to 505, 540, and 613 nm for PhO-Cl12BsubPc (10). A broad band in the region of 650 – 800 nm is also formed for the reduced species.

Page 24 of 37 ACS Paragon Plus Environment

Page 24 of 44

Page 25 of 44 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 5. UV-visible spectral changes during the first reduction of PhO-BsubPc (1), PhOCl6BsubPc (9), PhO-Cl12BsubPc (10), Cl-BsubPc (13), and Cl-Cl6BsubPc (14), in CH2Cl2 containing 0.1 M TBAP as the electrolyte.

In a similar fashion, the absorption profile shifts to higher wavelengths upon going from Cl-BsubPc (13) to Cl-Cl6BsubPc (14), with the main peaks moving from 302, 502, and 564 nm to 314, 525, and 570 nm, respectively, for the two neutral species. However, upon reduction, the maximum BsubPc absorbance peak flattens out and is characterized by a single broad band at 500 or 552 nm as opposed to the three equal-intensity bands typical of other derivatives. The broad band in the 600 – 800 nm region also becomes more pronounced with peaks at 657 and 705 nm for Cl-BsubPc (13) and Cl-Cl6BsubPc (14), respectively. Again, the flattening of the major peaks upon reduction correlates with the irreversible and unclear reduction processes in Page 25 of 37 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

the electrochemical measurements. This is likely due to the lower solubility and the large dipole moment of the boron-to-chlorine axial substituent.7, 31, 47 Figure 6 depicts the oxidation processes for these five compounds and reveals how the spectra of the generated oxidized products are flattened just like the axially-substituted BsubPcs. For the phenoxy-axial substituted BsubPcs with peripheral chlorination, the ~300 nm band broadens with peaks at 350 and 400 nm for PhO-Cl6BsubPc (9) and PhO-Cl12BsubPc (10). However, the major peak between 561 – 586 nm completely flattens to zero absorbance. The same disappearance of the major 564 – 570 nm band occurs with Cl-BsubPc (13) and ClCl6BsubPc (14), but the ~300 nm band decreases in intensity, shifts to 322 nm, and does not broaden as significantly as for the other peripherally-chlorinated BsubPcs. Future exploration of the reversibility of the reduction and oxidation processes via spectroelectrochemical analysis will give a better idea of radical cation stability and whether the typical BsubPc absorbance profile can be regenerated from the flattened absorbance spectra of the oxidized species.

Page 26 of 37 ACS Paragon Plus Environment

Page 26 of 44

Page 27 of 44 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 6. UV-visible spectral changes during the first oxidation of PhO-BsubPc (1), PhOCl6BsubPc (9), PhO-Cl12BsubPc (10), Cl-BsubPc (13), and Cl-Cl6BsubPc (14), in CH2Cl2 containing 0.1 M TBAP as the electrolyte.

CONCLUSION In summary, a series of 14 BsubPcs with different axial and/or peripheral substituents were synthesized and examined by electrochemistry and UV-visible spectroelectrochemistry in CH2Cl2 containing 0.1 M TBAP. Each compound undergoes two or more one-electron reductions and a single one-electron oxidation within the potential limit of the electrochemical

Page 27 of 37 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 28 of 44

solvent (+1.6 to -1.8 V vs. SCE). In general, the peripheral substituents shift all of the redox potentials more positively, while the reversibility of these processes seems to depend upon solubility and upon both the peripheral and axial substituents, with peripheral chlorines and phenoxy axial ligand resulting in the most reversible processes. The reversibility of the redox reactions appears to correlate with theoretically calculated dipole moments of the boron-to-axial ligand bond and the lengthening and shortening of this bond upon oxidation or reduction, respectively, as indicated by the fact that the phenyl axial substituent has the smallest dipole moment and most reversible electrochemistry. Spectroscopic properties of the singly-reduced and singly-oxidized BsubPcs can be tuned with changes in axial and peripheral substituents. For the BsubPcs without chlorine or fluorine on the axial ligand, the UV-visible spectra of the anion radical are characterized by three equalintensity peaks in the 450 to 650 nm region, a reduced intensity peak at ~300 nm, and a broad band in the region of 600 and 800 nm. The absorption profiles of the cationic species are considerably flattened, compared to the anionic species. This is true for all of the BsubPcs which show only a broad peak between 300 and 400 nm. These trends are important in the design of BsubPc

derivatives

for

use

in

high

performance,

organic

electronic

devices.

Spectroelectrochemistry provides simultaneous characterization of photochemical and electronic properties, similar to the function of an OPV. Future work on the spectroelectrochemical reversibility for the redox processes of these BsubPcs, both in solution and in the solid-state, will further enhance our understanding of their material and device stability and knowledge as to their use as better electron acceptor (stable anion) or electron donor (stable cation) materials in organic electronics.

Page 28 of 37 ACS Paragon Plus Environment

Page 29 of 44 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

AUTHOR INFORMATION The authors declare no competing financial interest.

ACKNOWLEDGEMENTS This work was supported by a Natural Sciences and Engineering Research Council (NSERC) Alexander Graham Bell Canada Graduate Scholarship (Doctoral Level) and Ontario government Queen Elizabeth II Graduate Scholarship in Science and Technology (QEII-GSST) to K.L.S., an NSERC Alexander Graham Bell Canada Graduate Scholarship (Masters Level) to K.L.S, and the Robert A. Welch Foundation (K. M. K. Grant E-680). Support was also received through the NSERC via a Discovery Grant to T.P.B..

SUPPORTING INFORMATION The Supporting Information is available and contains extinction coefficients of investigated BsubPcs, spectroelectrochemical data of neutral and anion radicals of the fluorophenoxy-BsubPcs, cyclic voltammograms of the axial phenoxy-substituted BsubPcs, UVvisible spectra of neutral and reduced fluorophenoxy-BsubPcs, and UV-visible spectral changes of F5BsubPc and 3-MePhO-BsubPc during first reduction and first oxidation processes.

REFERENCES 1.

Morse, G. E.; Bender, T. P., Boron Subphthalocyanines as Organic Electronic Materials.

ACS Appl Mater Interfaces 2012, 4 (10), 5055-5068.

Page 29 of 37 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

2.

Helander, M. G.; Morse, G. E.; Qiu, J.; Castrucci, J. S.; Bender, T. P.; Lu, Z.-H.,

Pentafluorophenoxy Boron Subphthalocyanine As a Fluorescent Dopant Emitter in Organic Light Emitting Diodes. ACS Appl. Mater. Interfaces 2010, 2 (11), 3147-3152. 3.

Josey, D. S.; Castrucci, J. S.; Dang, J. D.; Lessard, B. H.; Bender, T. P., Evaluating

Thiophene Electron-Donor Layers for the Rapid Assessment of Boron Subphthalocyanines as Electron Acceptors in Organic Photovoltaics: Solution or Vacuum Deposition? ChemPhysChem 2015, 16 (6), 1245-1250. 4.

New, E.; Howells, T.; Sullivan, P.; Jones, T. S., Small Molecule Tandem Organic

Photovoltaic Cells Incorporating an α-NPD Optical Spacer Layer. Org. Electron. 2013, 14 (9), 2353-2359. 5.

Cnops, K.; Zango, G.; Genoe, J.; Heremans, P.; Martinez-Diaz, M. V.; Torres, T.; Cheyns,

D., Energy Level Tuning of Non-Fullerene Acceptors in Organic Solar Cells. J. Am. Chem. Soc. 2015, 137 (28), 8991-8997. 6.

Chandran, H. T.; Ng, T.-W.; Foo, Y.; Li, H.-W.; Qing, J.; Liu, X.-K.; Chan, C.-Y.; Wong,

F.-L.; Zapien, J. A.; Tsang, S.-W., et al., Direct Free Carrier Photogeneration in Single Layer and Stacked Organic Photovoltaic Devices. Adv. Mater. 2017, 29 (22), 1606909. 7.

Morse, G. E.; Paton, A. S.; Lough, A.; Bender, T. P., Chloro boron Subphthalocyanine

and its Derivatives: Dyes, Pigments or Somewhere in Between? Dalton Trans. 2010, 39 (16), 3915-3922. 8.

Menke, S. M.; Mullenbach, T. K.; Holmes, R. J., Directing Energy Transport in Organic

Photovoltaic Cells Using Interfacial Exciton Gates. ACS Nano 2015, 9 (4), 4543-4552.

Page 30 of 37 ACS Paragon Plus Environment

Page 30 of 44

Page 31 of 44 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

9.

Gommans, H.; Schols, S.; Kadashchuk, A.; Heremans, P.; Meskers, S. C. J., Exciton

Diffusion Length and Lifetime in Subphthalocyanine Films. J. Phys. Chem. C 2009, 113 (7), 2974-2979. 10.

Cnops, K.; Rand, B. P.; Cheyns, D.; Verreet, B.; Empl, M. A.; Heremans, P., 8.4%

Efficient Fullerene-Free Organic Solar Cells Exploiting Long-Range Exciton Energy Transfer. Nat. Commun. 2014, 5, 3406. 11.

Morse, G. E.; Gantz, J. L.; Steirer, K. X.; Armstrong, N. R.; Bender, T. P.,

Pentafluorophenoxy Boron Subphthalocyanine (F5BsubPc) as a Multifunctional Material for Organic Photovoltaics. ACS Appl Mater Interfaces 2014, 6 (3), 1515-1524. 12.

Deibel, C.; Strobel, T.; Dyakonov, V., Role of the Charge Transfer State in Organic

Donor-Acceptor Solar Cells. Adv. Mater. (Weinheim, Ger.) 2010, 22 (37), 4097-4111. 13.

Shoaee, S.; Clarke, T. M.; Huang, C.; Barlow, S.; Marder, S. R.; Heeney, M.; McCulloch,

I.; Durrant, J. R., Acceptor Energy Level Control of Charge Photogeneration in Organic Donor/Acceptor Blends. J. Am. Chem. Soc. 2010, 132 (37), 12919-12926. 14.

Morse, G. E.; Helander, M. G.; Stanwick, J.; Sauks, J. M.; Paton, A. S.; Lu, Z.-H.;

Bender, T. P., Experimentally Validated Model for the Prediction of the HOMO and LUMO Energy Levels of Boronsubphthalocyanines. J. Phys. Chem. C 2011, 115 (23), 11709-11718. 15.

Del Rey, B.; Keller, U.; Torres, T.; Rojo, G.; Agullo-Lopez, F.; Nonell, S.; Marti, C.;

Brasselet, S.; Ledoux, I.; Zyss, J., Synthesis and Nonlinear Optical, Photophysical, and Electrochemical Properties of Subphthalocyanines. J. Am. Chem. Soc. 1998, 120 (49), 1280812817. 16.

Gonzalez-Rodriguez, D.; Torres, T., Peripheral Functionalization of Subphthalocyanines.

Eur. J. Org. Chem. 2009, (12), 1871-1879.

Page 31 of 37 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

17.

Paton, A. S.; Lough, A. J.; Bender, T. P., One Well-Placed Methyl Group Increases the

Solubility of Phenoxy Boronsubphthalocyanine Two Orders of Magnitude. Ind. Eng. Chem. Res. 2012, 51 (18), 6290-6296. 18.

Koca, A.; Oezkaya, A. R.; Selcukoglu, M.; Hamuryudan, E., Electrochemical and

Spectroelectrochemical Characterization of the Phthalocyanines with Pentafluorobenzyloxy Substituents. Electrochim. Acta 2007, 52 (7), 2683-2690. 19.

Koca, A., Spectroelectrochemistry of Phthalocyanines. In Electrochemistry of N4

Macrocyclic Metal Complexes: Volume 2: Biomimesis, Electroanalysis and Electrosynthesis of MN4 Metal Complexes, Zagal, J. H.; Bedioui, F., Eds. Springer International Publishing: Cham, 2016; pp 135-200. 20.

Kobayashi, N.; Lam, H.; Nevin, W. A.; Janda, P.; Leznoff, C. C.; Koyama, T.; Monden,

A.; Shirai, H., Synthesis, Spectroscopy, Electrochemistry, Spectroelectrochemistry, LangmuirBlodgett Film Formation, and Molecular Orbital Calculations of Planar Binuclear Phthalocyanines. J. Am. Chem. Soc. 1994, 116 (3), 879-890. 21.

Nevin, W. A.; Hempstead, M. R.; Liu, W.; Leznoff, C. C.; Lever, A. B. P.,

Electrochemistry and Spectroelectrochemistry of Mononuclear and Binuclear Cobalt Phthalocyanines. Inorg. Chem. 1987, 26 (4), 570-578. 22.

Arici, M.; Arican, D.; Ugur, A. L.; Erdogmus, A.; Koca, A., Electrochemical and

Spectroelectrochemical Characterization of Newly Synthesized Manganese, Cobalt, Iron and Copper Phthalocyanines. Electrochim. Acta 2013, 87, 554-566. 23.

Koca, A.; Ozcesmeci, M.; Hamuryudan, E., Substituents Effects to the Electrochemical,

and In Situ Spectroelectrochemical Behavior of Metallophthalocyanines: Electrocatalytic Application for Hydrogen Evolution Reaction. Electroanalysis 2010, 22 (14), 1623-1633.

Page 32 of 37 ACS Paragon Plus Environment

Page 32 of 44

Page 33 of 44 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

24.

Morse, G. E.; Castrucci, J. S.; Helander, M. G.; Lu, Z. H.; Bender, T. P., Phthalimido-

Boronsubphthalocyanines:

New

Derivatives

of

Boronsubphthalocyanine

with

Bipolar

Electrochemistry and Functionality in OLEDs. ACS Appl. Mater. Interfaces 2011, 3 (9), 35383544. 25.

González-Rodriguez, D.; Torres, T.; Olmstead, M. M.; Rivera, J.; Angeles Herranz, M.;

Echegoyen, L.; Atienza Castellanos, C.; Guldi, D. M., Photoinduced Charge-Transfer States in Subphthalocyanine-Ferrocene Dyads. J. Am. Chem. Soc. 2006, 128 (33), 10680-10681. 26.

González-Rodriguez, D.; Torres, T.; Herranz, M. A.; Echegoyen, L.; Carbonell, E.; Guldi,

D. M., Screening Electronic Communication through ortho-, meta- and para-Substituted Linkers Separating Subphthalocyanines and C60. Chem. - Eur. J. 2008, 14 (25), 7670-7679. 27.

González-Rodríguez, D.; Torres, T.; Guldi, D. M.; Rivera, J.; Herranz, M. Á.; Echegoyen,

L., Subphthalocyanines:  Tuneable Molecular Scaffolds for Intramolecular Electron and Energy Transfer Processes. J. Am. Chem. Soc. 2004, 126 (20), 6301-6313. 28.

Shimizu,

S.;

Yamazaki,

Y.;

Kobayashi,

N.,

Tetrathiafulvalene-Annulated

Subphthalocyanines. Chem. - Eur. J. 2013, 19 (23), 7324-7327. 29.

Claessens, C. G.; Gonzalez-Rodriguez, D.; Rodriguez-Morgade, M. S.; Medina, A.;

Torres, T., Subphthalocyanines, Subporphyrazines, and Subporphyrins: Singular Nonplanar Aromatic Systems. Chem. Rev. 2014, 114 (4), 2192-2277. 30.

González-Rodríguez, D.; Torres, T.; Guldi, D. M.; Rivera, J.; Echegoyen, L., Energy

Transfer Processes in Novel Subphthalocyanine−Fullerene Ensembles. Org. Lett. 2002, 4 (3), 335-338. 31.

Fulford, M. V.; Jaidka, D.; Paton, A. S.; Morse, G. E.; Brisson, E. R. L.; Lough, A. J.;

Bender, T. P., Crystal Structures, Reaction Rates, and Selected Physical Properties of Halo-

Page 33 of 37 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

Boronsubphthalocyanines (Halo = Fluoride, Chloride, and Bromide). J. Chem. Eng. Data 2012, 57 (10), 2756-2765. 32.

Potz, R.; Goldner, M.; Huckstadt, H.; Cornelissen, U.; Tutass, A.; Homborg, H.,

Synthesis and Structural Characterization of Boron Subphthalocyaninates. Z. Anorg. Allg. Chem. 2000, 626 (2), 588-596. 33.

Rodriguez-Morgade, M. S.; Claessens, C. G.; Medina, A.; Gonzalez-Rodriguez, D.;

Gutierrez-Puebla, E.; Monge, A.; Alkorta, I.; Elguero, J.; Torres, T., Synthesis, Characterization, Molecular Structure and Theoretical Studies of Axially Fluoro-Substituted Subazaporphyrins. Chem. - Eur. J. 2008, 14 (4), 1342-1350. 34.

Ebenhoch, B.; Prasetya, N. B. A.; Rotello, V. M.; Cooke, G.; Samuel, I. D. W., Solution-

Processed Boron Subphthalocyanine Derivatives as Acceptors for Organic Bulk-Heterojunction Solar Cells. J. Mater. Chem. A 2015, 3 (14), 7345-7352. 35.

Duan, C.; Zango, G.; Garcia Iglesias, M.; Colberts, F. J. M.; Wienk, M. M.; Martinez-

Diaz, M. V.; Janssen, R. A. J.; Torres, T., The Role of the Axial Substituent in Subphthalocyanine Acceptors for Bulk-Heterojunction Solar Cells. Angew. Chem., Int. Ed. 2017, 56 (1), 148-152. 36.

Beaumont, N.; Castrucci, J. S.; Sullivan, P.; Morse, G. E.; Paton, A. S.; Lu, Z.-H.; Bender,

T. P.; Jones, T. S., Acceptor Properties of Boron Subphthalocyanines in Fullerene Free Photovoltaics. J. Phys. Chem. C 2014, 118 (27), 14813-14823. 37.

Castrucci, J. S.; Josey, D. S.; Thibau, E.; Lu, Z.-H.; Bender, T. P., Boron

Subphthalocyanines as Triplet Harvesting Materials within Organic Photovoltaics. J. Phys. Chem. Lett. 2015, 6 (15), 3121-3125.

Page 34 of 37 ACS Paragon Plus Environment

Page 34 of 44

Page 35 of 44 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

38.

Sullivan, P.; Duraud, A.; Hancox, I.; Beaumont, N.; Mirri, G.; Tucker, J. H. R.; Hatton, R.

A.; Shipman, M.; Jones, T. S., Halogenated Boron Subphthalocyanines as Light Harvesting Electron Acceptors in Organic Photovoltaics. Adv. Energy Mater. 2011, 1 (3), 352-355. 39.

Lingane, J. J., Interpretation of the Polarographic Waves of Complex Metal Ions. Chem.

Rev. 1941, 29 (1), 1-35. 40.

Kadish, K. M.; Mu, X. H.; Lin, X. Q., The Construction and Utilization of a Simple

Light-Transparent FTIR Spectroelectrochemical Cell with Thin-Layer Chamber. Electroanalysis 1989, 1 (1), 35-41. 41.

Morse, G. E.; Helander, M. G.; Maka, J. F.; Lu, Z.-H.; Bender, T. P., Fluorinated

Phenoxy Boron Subphthalocyanines in Organic Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2010, 2 (7), 1934-1944. 42.

Sampson, K. L.; Josey, D. S.; Li, Y.; Virdo, J. D.; Lu, Z.-H.; Bender, T. P., Ability To

Fine-Tune

the

Electronic

Properties

and

Open-Circuit

Voltage

of

Phenoxy-Boron

Subphthalocyanines through Meta-Fluorination of the Axial Substituent. J. Phys. Chem. C 2018, 122 (2), 1091-1102. 43.

Bonnier, C.; Josey, D. S.; Bender, T. P., Aryl-Substituted Boron Subphthalocyanines and

their Application in Organic Photovoltaics. Aust. J. Chem. 2015, 68 (11), 1750-1758. 44.

Dang, J. D.; Virdo, J. D.; Lessard, B. H.; Bultz, E.; Paton, A. S.; Bender, T. P., A Boron

Subphthalocyanine Polymer: Poly(4-methylstyrene)-co-poly(phenoxy boron subphthalocyanine). Macromolecules 2012, 45 (19), 7791-7798. 45.

Dang, J. D.; Fulford, M. V.; Kamino, B. A.; Paton, A. S.; Bender, T. P., Process for the

Synthesis

of

Symmetric

and

Unsymmetric

Oxygen

Bridged

Dimers

Subphthalocyanines (µ-oxo-(BsubPc)2s). Dalton Trans. 2015, 44 (9), 4280-4288.

Page 35 of 37 ACS Paragon Plus Environment

of

Boron

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

46.

Morse,

G.

E.;

Bender,

T.

P.,

Aluminum

Chloride Activation

Page 36 of 44

of Chloro-

Boronsubphthalocyanine: a Rapid and Flexible Method for Axial Functionalization with an Expanded Set of Nucleophiles. Inorg. Chem. 2012, 51 (12), 6460-6467. 47.

Ferro, V. R.; De La Vega, J. M. G.; Claessens, C. G.; Poveda, L. A.; Gonzalez-Jonte, R.

H., The Axial Coordination in Subphthalocyanines. Geometrical and Electronic Aspects. J. Porphyrins Phthalocyanines 2001, 5 (6), 491-499. 48.

Ferro, V. R.; Poveda, L. A.; Claessens, C. G.; González-Jonte, R. H.; García de la Vega, J.

M., Density Functional Study of the Redox Processes in Subphthalocyanines. Int. J. Quantum Chem 2003, 91 (3), 369-375. 49.

Ferro, V. R.; Poveda, L. A.; González-Jonte, R. H.; De La Vega, J. M. G.; Torres, T.; Rey,

B. D., Molecular Electronic Structure of Subphthalocyanine Macrocycles. J. Porphyrins Phthalocyanines 2000, 04 (06), 611-620. 50.

Ferro, V. R.; Garcia de la Vega, J. M.; Gonzalez-Jonte, R. H.; Poveda, L. A., A

Theoretical Study of Subphthalocyanine and its nitro- and tert-butyl Derivatives. J. Mol. Struct.: THEOCHEM 2001, 537, 223-234. 51.

Virdo, J. D.; Lough, A. J.; Bender, T. P., Redetermination of the Crystal Structure of

Boron Subphthalocyanine Chloride (Cl-BsubPc) Enabled by Slow Train Sublimation. Acta Crystallographica Section C 2016, 72 (4), 297-307.

Page 36 of 37 ACS Paragon Plus Environment

Page 37 of 44 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

TOC Graphic

Page 37 of 37 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 Figure 81x44mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 38 of 44

Page 39 of 44 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 155x87mm (300 x 300 DPI)

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 170x197mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 40 of 44

Page 41 of 44 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 3 195x184mm (300 x 300 DPI)

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 4 202x90mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 42 of 44

Page 43 of 44 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 5 211x159mm (300 x 300 DPI)

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 6 195x165mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 44 of 44