Ask Not How Many, But Where They Are ... - ACS Publications

Jan 17, 2018 - Ask Not How Many, But Where They Are: Substituents Control Energetic Ordering of Frontier Orbitals/Electronic Structures in Isomeric ...
0 downloads 0 Views 11MB Size
Subscriber access provided by READING UNIV

Article

Ask Not How Many, But Where They Are: Substituents Control Energetic Ordering of Frontier Orbitals/Electronic Structures in Isomeric Methoxy- Substituted Dibenzochrysenes Maxim V. Ivanov, Marat R. Talipov, Tushar Shivram Navale, and Rajendra Rathore J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b11232 • Publication Date (Web): 17 Jan 2018 Downloaded from http://pubs.acs.org on January 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 free 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 accessible to all readers and 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.

The Journal of Physical Chemistry C 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

Ask Not How Many, But Where They Are: Substituents Control Energetic Ordering of Frontier Orbitals/Electronic Structures in Isomeric MethoxySubstituted Dibenzochrysenes Maxim V. Ivanov, Marat R. Talipov, Tushar S. Navale and Rajendra Rathore* Department of Chemistry, Marquette University, P.O. Box 1881, Milwaukee, WI 53201-1881.

ABSTRACT. Redox properties of polycyclic aromatic hydrocarbons (PAHs) can be modulated by substitution with electron-rich groups. Here we show, using the example of dibenzo[g,p]chrysene (DBC), that substitution position (i.e., meta vs para) alters the energetic ordering of frontier molecular orbitals (FMOs), leading to cation radicals with altered electronic structures and thereby redox/optical properties. We demonstrate that a straightforward analysis of FMOs provides a valuable insight towards the rational design of novel PAHs with tailored redox properties.

ACS Paragon Plus Environment

1

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 2 of 24

Introduction Extensive efforts continue towards the design and synthesis of new and improved (derivatized) polycyclic aromatic hydrocarbons (PAHs) as building blocks for the preparation of functional electronic and optoelectronic materials for modern photovoltaic applications.1-3 As an example, molecules based on dibenzo[g,p]chrysene4 (DBC), a twisted PAH, have been extensively explored for potential applications in the areas of sensors, non-linear optical and liquid-crystalline materials.5-8 para

O

meta

O

O

O

O

O

O

O

m-DBC

O

O

O O

DBC

O

O

m/p-DBC

O

O

p-DBC

Chart 1. Structure and naming scheme of DBCs. It is known that oxidation potentials of PAHs can be modulated by appropriate substitutions with electron rich alkyl, alkoxy, or dialkylamino groups.8-10 In larger PAHs such as DBC, it is not immediately obvious whether the substitution of para (i.e. p-DBC) or meta (i.e. m-DBC) or both meta/para (m/p-DBC) positions (see Chart 1) would achieve maximum modulation of its redox properties. In this report, we show experimentally (via generation of cation radicals and electrochemical and spectroscopic measurements) and computationally (DFT) for parent and methoxy-substituted DBCs (Chart 1) that the position of substitution can have dramatic impact on redox and optical properties and alter the electronic structures of the cation radicals. For example, the oxidation potentials of isomeric m-DBC vs p-DBC, with four methoxy substituents, surprisingly differ by some 300 mV, while the m/p-DBC, with eight methoxy groups, was found to have a higher redox potential (by 60 mV) than m-DBC.

ACS Paragon Plus Environment

2

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

An analysis of frontier molecular orbitals (FMOs) can help explain these surprising trends. In particular, the position of substituents impacts the energetic ordering of the filled FMOs (i.e. HOMO/HOMO-1) and can also lead to cation radicals with altered electronic structures, and, in turn, dramatically altered redox/optical properties.11,12 In general, the spatial hole distribution in PAH cation radicals can be ascertained by inspecting the nodal structure and distribution of HOMO densities in corresponding neutral PAH.11 Accordingly, the example of DBCs presented herein will serve to demonstrate that understanding the substitutional positiondependent redox potential modulation of PAHs, based on the analysis of FMOs, can serve as a valuable tool for informed design and synthesis of novel PAH derivatives with desired redox properties. Results and discussion Synthesis and electrochemistry of DBCs and electronic spectroscopy of DBC cation radicals. Synthesis of various methoxy-substituted DBCs was carried out by adaptation of literature procedures.4-8

For

example,

an

oxidative

cyclodehydrogenation

of

the

tetrakis(3-

methoxyphenyl)ethylene and tetrakis(3,4-dimethoxyphenyl)ethylene, using either FeCl3 or DDQ/CH3SO3H as oxidants,13 afforded p-DBC and m/p-DBC, respectively, in excellent yields. Unfortunately, a similar oxidative cyclodehydrogenation of tetrakis(4-methoxyphenyl)ethylene did not produce any m-DBC;8,9 and therefore, it was accessed via an oxidative cyclization of bis(bianisyl)acetylene,

which,

in

turn,

was

prepared

by

Sonogashira

coupling

of

trimethylsilyacetylene with 2-iodobianisyl (Scheme 1 and the Supporting Information for complete experimental details).

ACS Paragon Plus Environment

3

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

X

X

y

y

DDQ/acid

DDQ/acid

X = OMe Y = OMe

X = OMe Y=H

m/p-DBC

y

y X

X

X y

I y

y

X X

SiMe3

X

PdCl2(PPh3)2

FeCl3

X=H Y = OMe

X=H Y = OMe

y

p-DBC

DDQ/acid X = H, Y = OMe

y

X

X

Page 4 of 24

y

m-DBC

X

Scheme 1. Synthesis of methoxy-substituted DBCs All DBCs were characterized by 1H/13C NMR spectroscopy and their molecular structures were established by X-ray crystallography. The X-ray structures of various DBCs (Scheme 1), including parent DBC, showed similar twist angles of 35°-37° around the central double bond (Figure S2 in the Supporting Information) irrespective of the number or position of the methoxy substituents. The cyclic voltammograms of all DBCs met the criteria of reversibility over scan rates of 50 to 1000 mV/s, when the potential window was limited in the range of their first oxidation potentials (solid curves in Figure 1).

ACS Paragon Plus Environment

4

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

0.15

DBC O

0.48

O

O

O

p-DBC O

O

O

O

O

m-DBC

0.42 O

O

O

O

O O

O

m/p-DBC

1.55

1.15 0.75 + 0.35 V vs Fc/Fc

Figure 1. Cyclic voltammograms of 2 mM DBCs in CH2Cl2 (0.2 M n-Bu4NPF6) recorded at a platinum electrode at a scan rate of 200 mV/s and 22 oC. The oxidation potentials were referenced to ferrocene. The CV data show that the first oxidation potential of parent (unsubstituted) dibenzochrysene [Eox1(DBC) = 0.88 V vs Fc/Fc+] decreases by 0.42 V upon incorporation of the 8 methoxy groups in all of its meta- and para-positions [i.e. Eox1(m/p-DBC) = 0.46 V]. Surprisingly, however, a removal of 4 methoxy groups from the para-positions of m/p-DBC led to a further decrease in its oxidation potential by 60 mV [i.e. Eox1(m-DBC) = 0.40 V]. This would clearly suggest that the four para-methoxy groups in m/p-DBC do not play any role in stabilizing its cation radical. Based on this observation, one may expect that the oxidation potentials of parent DBC and tetramethoxy p-DBC should be similar. However, the oxidation potential of tetramethoxy p-DBC [Eox1(p-DBC) = 0.73 V] was found to be 0.15 V lower than the parent DBC.

ACS Paragon Plus Environment

5

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 6 of 24

This surprisingly divergent effect of the para-methoxy groups in m/p-DBC, where they appear to destabilize the cation radical by 60 mV, and in p-DBC, where they stabilize the cation radical by 150 mV, suggests that these cation radicals must possess different electronic structures. To investigate this effect, we next examine the electronic absorption spectra of the DBC cation radicals, generated by quantitative redox titrations14 using either a hindered naphthalene cation radical (NAP+•; Ered = 0.94 V) or tetrasubstituted p-hydroquinone ether (THEO+•; Ered = 0.67 V) as oxidants15 in CH2Cl2 at 22 oC; see Supporting Information for details. A comparison of the reproducible absorption spectra of DBC+• showed remarkable similarity between the spectra of m/p-DBC+• and m-DBC+• (Figure 2), as both spectra have (i) intense bands at ~800-900 nm (λmax = 818 and 886 nm, respectively) with comparable extinction coefficients (log εmax = 4.36 and 4.28, respectively) and (ii) a small broad band in the nearinfrared region of the spectrum at ~1300 nm. Moreover, the absorption spectrum of DBC+•, although somewhat blue-shifted and less intense, bears similar spectral features as m/p-DBC+• and m-DBC+• (Figure 2). However, the absorption spectrum of p-DBC+• is rather dissimilar from the absorption spectra of other DBC cation radicals as it has a much less intense band at 808 nm (log ε808 = 3.97) and an additional broad near-IR band that extends beyond ~3000 nm, thus suggesting a different electronic structure of p-DBC+• as compared to m-DBC+• or m/p-DBC+•.

ACS Paragon Plus Environment

6

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

DBC+•

m/p-DBC+•

m-DBC+•

p-DBC+•

1000

2000 Wavelength (nm)

3000

Figure 2. A comparison of the absorption spectra of various DBC cation radicals in CH2Cl2 at 22 oC. DFT modeling of the redox/optical properties of DBCs. Accurate description of the electronic structure of PAH cation radicals is challenging for many DFT functionals due to the inherent self-interaction error that causes artificial hole delocalization and leads to underestimated oxidation potentials and inaccurate identification of the nature of the excited states.16-18 These artifacts can be minimized using a hybrid density functional that contains a portion of the exact Hartree-Fock (HF) exchange.19,20 However, in many standard hybrid functionals amount of the Hartree-Fock exchange is either insufficient or excessive for a correct description of the extent of hole delocalization/stabilization.21,22 Recent benchmarking studies on a set of mixed-valence compounds20,21,23 and poly-p-phenylene wires24 have demonstrated that customization of a standard density functional by fine-tuning the amount of the exact exchange to reproduce experimental data can provide a reliable description of the electronic structures of the πconjugated cation radicals. For example, in our previous study24 we have used a one-parameter

ACS Paragon Plus Environment

7

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 8 of 24

B1LYP25 functional where HF exchange was varied in 20-50% range to accurately reproduce oxidation potentials and the cation radical excitation energies of the alkyl-capped poly-pphenylenes with increasing number of p-phenylenes, i.e., iAPPn, n = 2-7. It was shown that 40% of the HF exchange provides a balanced description of the electronic structure of the poly-pphenylene cation radicals similar to that shown by Kaupp and coworkers23 for a reliable description of the another class of organic mixed-valence systems. Indeed, usage of modified B1LYP-40/6-31G(d) functional performed exceptionally well in reproducing the experimental redox/optoelectronic properties of a variety of poly-p-phenylene-10,26 and polyfluorene-based27 wires and other PAHs.28,29 Accordingly, in this manuscript, we performed electronic structure calculations of neutral DBCs and their cation radicals using B1LYP-40/6-31G(d) level of theory. Due to the important role of the solvation in stabilization of charged molecules and the affect on the nature of the excited states,21,30,31 we also performed a comparative analysis of the electronic structure of DBC cation radicals in solvents with varied polarity using polarizable continuum model (PCM) in linear response formalism (vide infra). Excited state calculations of the cation radicals of the DBC derivatives were performed using time-dependent DFT with B1LYP-40 functional, which has proven its reliability in reproducing correct nature of the first excited state and experimental excitation energies of a large variety of PAH cation radicals.10,27 To further validate the performance of TD-DFT in description of the excited state of π-conjugated cation radicals,32,33 we performed a benchmarking study on the performance of TD-B1LYP-40 to reproduce ab initio excitation energies of a set of several relevant to this study PAHs (Figure 3). As a reference method, we employed EOM-IPCCSD method due to its reported reliability in accurate description of the charge-(de)localized

ACS Paragon Plus Environment

8

Page 9 of 24

open-shell systems.34-36 Excitation energies of the PAH cation radicals were taken as the difference between first and second ionization energies (i.e., IE2-IE1) from EOM-IP-CCSD/631G(d) calculations performed at their equilibrium geometries optimized using B1LYP-40/631G(d) level of theory. Thus obtained excitation energies were then compared with the excitation energy (ν) calculated directly using TD-B1LYP-40/6-31G(d). This analysis showed that the TD-B1LYP-40/6-31G(d) level of theory correlates well with the results from EOM-IP-CCSD/6-31G(d) (R2 = 0.97, Figure 3) and reproduces ab initio energies with RMSD of 0.12 eV. Importantly, the orbitals involved in the excitation of the cation radicals obtained from TD-DFT calculations correspond to the orbitals that are involved in the first and second ionizations obtained using EOM-IP-CCSD method (see Supporting Information for additional details on the nature of transition in each molecule in Figure 3). 2.0 O O

O

O

O

O

O

O

O

O

O

O

ν (EOM-IP-CCSD), eV

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

1.1

O

O

O

O

O

O

O

0.2

1.1 ν (TD-DFT), eV

0.2 2.0

Figure 3. Left: Parent and methoxy-substituted PAH structures used in the benchmarking study. Right: Correlation plot between excitation energies of the cation radicals (structures on the left) calculated using EOM-IP-CCSD/6-31G(d) and TD-B1LYP-40/6-31G(d). Linear regression: y = 1.09x – 0.18, R2 = 0.97. Molecular orbital theory description of DBCs. It has been shown on a number of examples,12,37,38 that the electronic structures of the π-conjugated cation radicals are governed by

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry

their HOMO energies and nodal arrangements. Therefore, in order to investigate the origin of the remarkable differences in the redox/optical properties of p-DBC vs m-DBC vs m/p-DBC, below we examine their FMOs. As highlighted in Figure 4 below, the relative energies of FMOs in neutral DBCs display a strong dependence on the position of the methoxy substituents. DBC

m-DBC

0.17

-5.5

m/p-DBC

0.77

-5.7 -5.9 -6.1 -6.3

0.62

Orbital Energy (eV)

p-DBC

0.09

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 10 of 24

-6.5

Figure 4. Comparison of the HOMO/HOMO-1 energies of neutral DBCs calculated using B1LYP-40/6-31G(d)+PCM(CH2Cl2). The two filled FMOs of DBC in Figure 3 are hereafter referred to as phenanthrene-like and biphenyl-like orbitals, based on the resemblance of their electron density distributions to those of phenanthrene and biphenyl (Figure 5).

ACS Paragon Plus Environment

10

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

Figure 5. Schematic representations of phenanthrene-like (A) and biphenyl-like (B) FMOs of DBC and per-atom HOMO densities, obtained from the MO expansion coefficients of a single orbital. The orbital coefficients add up to 1.00 in phenanthrene-like orbital, while to 1.02 in biphenyl-like orbital due to the rounding error. A closer look at these FMOs suggests that the phenanthrene-like orbital bears significant electron density (or molecular orbital expansion coefficient squared, represented by yellow and blue circles in Figure 5A) at the meta carbons (0.05) and none at the para carbons. In contrast the biphenyl-like (Figure 5B) orbital bears significant electron density at both para (0.07) and meta (0.01) carbons, albeit in different amounts. It is emphasized that the energies of the FMOs in Figure 5 will respond to the substitution at meta and para carbons based on the amount of electron density coefficient, i.e., larger the coefficient at the substituent-bearing carbon, larger will be the increase in FMO energy. Based on this analysis, incorporation of the methoxy groups at meta position (i.e., mDBC) raises the energies of both FMOs (Figure 4). Addition of four more methoxy in m-DBC at the para positions (i.e., m/p-DBC) does not affect the energy of the phenanthrene-like HOMO, due to a complete lack of HOMO density at para carbons. The energy of HOMO-1 in m/p-DBC increases significantly as compared to m-DBC because the biphenyl-like orbital bears electron density at both meta and para carbons. Importantly, incorporation of the methoxy groups only at the para positions of DBC (i.e., p-DBC) does not affect the energy of its phenanthrene-like

ACS Paragon Plus Environment

11

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 24

orbital, but raises the energy of biphenyl-like orbital to a sufficient extent that it becomes HOMO (Figure 4).39 Relatively small HOMO/HOMO-1 energy gap of 0.09 eV in p-DBC may suggest that the solvent polarity could swap the order of these orbitals and thereby result in the cation radical of p-DBC with the alternate electronic structure similar to that of m-DBC. However, changing the solvent from toluene (ε = 2.4) to dichloromethane (8.9) to dichloroethane (10.1) to acetone (20.5) to acetonitrile (35.7) did not change the ordering of the FMOs in all DBCs examined, as their HOMO/HOMO-1 energy gap remains largely constant with changing solvent polarity (Figure 6).

Figure 6. Energies [B1LYP-40/6-31G(d)] of the HOMO/HOMO-1 of DBCs (biphenyl-like/red circles and phenanthrene-like/blue circles) against dielectric constant of the solvents of different polarity, i.e., toluene (ε = 2.4), dichloromethane (8.9), dichloroethane (10.1), acetone (20.5), and acetonitrile (35.7). Electronic structure of DBC cation radicals. The FMO analysis clearly demonstrates that the energetic ordering of HOMO/HOMO-1 can be controlled by positioning of the substituents and,

ACS Paragon Plus Environment

12

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

in turn, can lead to different electronic structures of the resulting cation radicals. The importance of this finding is recognized in light of prior works that have shown, through X-ray structural analysis of PAH cation radicals, that contraction/elongation of the C-C bonds upon 1-e- oxidation track in accordance with the nodal structure of HOMO (i.e., bonds with bonding HOMO lobes undergo

elongations

whereas

the

bonds

with

antibonding

HOMO

lobes

undergo

contractions).28,29,38 Moreover, the experimental bond length changes in various PAH cation radicals can be fully reproduced by DFT calculations and the calculated spin/charge distributions in these cation radicals closely resemble the HOMO density distributions.28,29,38 Indeed, the calculated spin-density distributions (Figure 7A) correspond to phenanthrene-like orbital in DBC+•, m-DBC+• and m/p-DBC+• and biphenyl-like orbital in p-DBC+•, in accordance with the nodal arrangement of HOMOs displayed in Figure 5.

ACS Paragon Plus Environment

13

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

DBC+•

m/p-DBC+•

m-DBC+•

Page 14 of 24

p-DBC+•

B DBC+•

m/p-DBC+•

m-DBC+•

p-DBC+•

1000

2000 Wavelength (nm)

3000

Figure 7. A. Isovalue spin-density distribution plots of parent and methoxy-substituted DBC cation radicals calculated using B1LYP-40/6-31G(d)+PCM(CH2Cl2). B. Experimental and TDDFT calculated stick spectra of parent and methoxy-substituted DBC cation radicals. Orbitals involved in the lowest energy transitions are shown as insets. The spectrum of 4,4’dimethoxybiphenyl cation radical is shown as dashed curve. While spin-density distributions in DBC+•, m-DBC+• and m/p-DBC+• are visually identical to the corresponding distribution of HOMO densities in neutral DBCs, it is curious that the

ACS Paragon Plus Environment

14

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

observed spin-density distribution in p-DBC+• occupies only one of the two biphenyl moieties linked via an ethylenic linkage and does not fully correspond to the biphenyl-like HOMO distribution (compare Figures 7 and 4). Such a discrepancy between spin/charge and HOMO distributions is known to arise in bichromophoric systems, in which interchromophoric electronic coupling (i.e., Hab) is far less than the reorganization energy (i.e., λ), i.e. 2Hab < λ.40-42 In such systems, the hole is known to be stabilized via dynamic hopping amongst the equivalent chromophores.43,44 Consistent with the drastically different spin-density distribution of p-DBC+• as compared to those of DBC+•, m-DBC+• and m/p-DBC+•, its electronic absorption spectrum is quite different (Figure 7B). Absorption band of p-DBC+• in the ~800-900 nm region closely resembles the absorption spectrum of 4,4’-dimethoxybiphenyl cation radical24 and therefore corresponds to the local excitation of a single biphenyl moiety of p-DBC+•. Additional (weak) broad absorption band in the near-IR region extending beyond 3000 nm then is readily assigned to as a charge transfer excitation where the hole hops between two biphenyl-like moieties. This analysis of the spectrum of p-DBC+• is further supported by the TD-DFT calculations and the charge-transfer nature of the transition displayed in Figure 7B. In contrast, (weak) broad band in the 1000-1500 nm region in the spectra of DBC+•, m-DBC+• and m/p-DBC+• involve relatively higher energy excitation of the cationic charge that is delocalized over the entire DBC core (Figure 7B). We next probed whether the nature of the transitions in DBC cation radicals changes upon changing the solvent polarity with the aid of TD-DFT calculations. As shown in Figure 8, changing the solvent from least polar toluene (ε = 2.4) to highly polar acetonitrile (35.7) did not affect the nature of the transitions in DBC cation radicals. In particular, the lowest-energy transitions in DBC+•, m-DBC+• and m/p-DBC+• remain invariant to the polarity of the solvent,

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry

while the charge-transfer transition observed in p-DBC+• experiences expected change from 0.6 to 0.8 eV (Figure 8).

CH2Cl2

1.20

.

m-DBC+ .

DBC+ . m/p-DBC+ ν , eV

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 16 of 24

0.85 .

p -DBC+

0.50

0

20

40

Figure 8. Excitation energy of parent and methoxy-substituted DBC cation radicals calculated using B1LYP-40/6-31G(d) in PCM of solvents with varying polarity. Conclusions In summary, we show that the positioning of the substituents controls the energetic ordering of the FMOs in DBCs and thereby results in cation radicals with different electronic structures and in turn different redox/optical properties. The variations in electronic structures of PAH cation radicals with positioning of the substituents can be readily ascertained by an inspection of the nodal arrangements of the FMOs and their energetic ordering and remains unchanged with varied solvent. The FMO analysis presented herein will serve as a valuable tool for the informed design and synthesis of novel PAH derivatives with desired (enhanced) redox properties.

ACS Paragon Plus Environment

16

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

ASSOCIATED CONTENT Supporting Information. Experimental and computational details are included in the Supporting Information. The following files are available free of charge. Supporting Information (PDF) Optimized equilibrium geometries of DBCs at neutral and cation radical states (TXT) AUTHOR INFORMATION Corresponding Author *[email protected] Notes Any additional relevant notes should be placed here. The authors declare no competing financial interests. ACKNOWLEDGMENT We thank the NSF (CHE-1508677) and NIH (R01-HL112639-04) for financial support, Professor Scott A. Reid for helpful discussions. The calculations were performed on the highperformance computing cluster Père at Marquette University and the Extreme Science and Engineering Discovery Environment (XSEDE). REFERENCES (1) Segawa, Y.; Ito, H.; Itami, K. Structurally Uniform and Atomically Precise Carbon Nanostructures. Nat. Rev. Mat. 2016, 1, 15002.

ACS Paragon Plus Environment

17

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 18 of 24

(2) Avouris, P.; Chen, Z.; Perebeinos, V. Carbon-based Electronics. Nat. Nanotechnol. 2007, 2, 605-615. (3) Heath, J. R. Molecular Electronics. Annu. Rev. Mater. Res. 2009, 39, 1-23. (4) Buckles, R. E.; Serianz, A.; Naffziger, D. Dibenzo (gp) Chrysene: A Challenging Experiment in Organic Synthesis. Proc.. Iowa. Acad.. Sci. 1973, 80, 45-49. (5) Yamaguchi, S.; Swager, T. M. Oxidative Cyclization of Bis (biaryl) Acetylenes: Synthesis and Photophysics of Dibenzo [g, P] Chrysene-based Fluorescent Polymers. J. Am. Chem. Soc. 2001, 123, 12087-12088. (6) Toal, S. J.; Trogler, W. C. Polymer Sensors for Nitroaromatic Explosives Detection. J. Mater. Chem. 2006, 16, 2871-2883. (7) Li, C. -W.; Wang, C. -I.; Liao, H. -Y.; Chaudhuri, R.; Liu, R. -S. Synthesis of Dibenzo[g,p]chrysenes From Bis(biaryl)acetylenes Via Sequential ICl-induced Cyclization and Mizoroki-Heck Coupling. J. Org. Chem. 2007, 72, 9203-9207. (8) Navale, T. S.; Thakur, K.; Rathore, R. Sequential Oxidative Transformation of Tetraarylethylenes to 9,10-diarylphenanthrenes and Dibenzo[g,p]chrysenes Using DDQ As An Oxidant. Org. Lett. 2011, 13, 1634-1637. (9) Navale, T. S.; Zhai, L.; Lindeman, S. V.; Rathore, R. Octamethoxydibenzochrysene: Isolation and X-ray Crystallographic Characterization of a Twisted Polyaromatic Cation Radical. Chem. Commun. (Camb). 2009, 2857-2859.

ACS Paragon Plus Environment

18

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

(10) Wang, D.; Talipov, M. R.; Ivanov, M. V.; Rathore, R. Energy Gap Between the Poly-pphenylene Bridge and Donor Groups Controls the Hole Delocalization in Donor--Bridge--Donor Wires. J. Am. Chem. Soc. 2016, 138, 16337-16344. (11) Talipov, M. R.; Navale, T. S.; Rathore, R. Nodal Arrangement of HOMOs in Polychromophoric Molecules and Assemblies Controls Interchromophoric Electronic Coupling. Angew. Chem. Int. Ed. 2015, 54, 14468-14472. (12) Merz, J.; Fink, J.; Friedrich, A.; Krummenacher, I.; Al Mamari, H.; Lorenzen, S.; Hähnel, M.; Eichhorn, A.; Moos, M.; Holzapfel, M.; et al. Pyrene MO Shuffle - Controlling Excited State and Redox Properties by Changing the Nature of the Frontier Orbitals. Chemistry. 2017, . (13) Zhai, L.; Shukla, R.; Rathore, R. Oxidative C- C Bond Formation (Scholl Reaction) with DDQ As An Efficient and Easily Recyclable Oxidant. Org. Lett. 2009, 11, 3474-3477. (14) Talipov, M. R.; Boddeda, A.; Hossain, M. M.; Rathore, R. Quantitative Generation of Cation Radicals and Dications Using Aromatic Oxidants: Effect of Added Electrolyte on the Redox Potentials of Aromatic Electron Donors. J. Phys. Org. Chem. 2015, 29, 227-233. (15) Talipov, M. R.; Rathore, R. Robust Aromatic Cation Radicals as Redox Tunable Oxidants. In Organic Redox Systems: Synthesis, Properties, and Applications John Wiley & Sons: Hoboken, NJ, 2015; pp 131. (16) Ranasinghe, D. S.; Margraf, J. T.; Jin, Y.; Bartlett, R. J. Does the Ionization Potential Condition Employed in QTP Functionals Mitigate the Self-interaction Error? J. Chem. Phys. 2017, 146, 034102.

ACS Paragon Plus Environment

19

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

(17) Lundberg, M.; Siegbahn, P. E. M. Quantifying the Effects of the Self-interaction Error in DFT: When Do the Delocalized States Appear? J. Chem. Phys. 2005, 122, 224103. (18) Zhang, Y.; Yang, W. A Challenge for Density Functionals: Self-interaction Error Increases for Systems with a Noninteger Number of Electrons. J. Chem. Phys. 1998, 109, 2604-2608. (19) Félix, M.; Voityuk, A. A. DFT Performance for the Hole Transfer Parameters in DNA Π Stacks. Int. J. Quantum. Chem. 2011, 111, 191-201. (20) Kaupp, M.; Renz, M.; Parthey, M.; Stolte, M.; Würthner, F.; Lambert, C. Computational and Spectroscopic Studies of Organic Mixed-valence Compounds: Where Is the Charge? Phys. Chem. Chem. Phys. 2011, 13, 16973-16986. (21) Renz, M.; Kess, M.; Diedenhofen, M.; Klamt, A.; Kaupp, M. Reliable Quantum Chemical Prediction of the Localized/Delocalized Character of Organic Mixed-Valence Radical Anions. From Continuum Solvent Models to Direct-COSMO-RS. J. Chem. Theory. Comput. 2012, 8, 4189-4203. (22) Yang, J.; Zhang, W.; Si, Y.; Zhao, Y. Intramolecular Electronic Couplings in Class II/III Organic Mixed-valence Systems of Bis(1,4-dimethoxybenzene). J. Phys. Chem. B. 2012, 116, 14126-14135. (23) Renz, M.; Theilacker, K.; Lambert, C.; Kaupp, M. A Reliable Quantum-chemical Protocol for the Characterization of Organic Mixed-valence Compounds. J. Am. Chem. Soc. 2009, 131, 16292-16302.

ACS Paragon Plus Environment

20

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

(24) Talipov, M. R.; Boddeda, A.; Timerghazin, Q. K.; Rathore, R. Key Role of End-capping Groups in Optoelectronic Properties of Poly-p-phenylene Cation Radicals. J. Phys. Chem. C. 2014, 118, 21400-21408. (25) Adamo, C.; Barone, V. Toward Reliable Adiabatic Connection Models Free From Adjustable Parameters. Chem. Phys. Lett. 1997, 274, 242-250. (26) Ivanov, M. V.; Chebny, V. J.; Talipov, M. R.; Rathore, R. Poly-p-hydroquinone Ethers: Isoenergetic Molecular Wires with Length-Invariant Oxidation Potentials and Cation Radical Excitation Energies. J. Am. Chem. Soc. 2017, 139, 4334-4337. (27) Ivanov, M. V.; Talipov, M. R.; Boddeda, A.; Abdelwahed, S. H.; Rathore, R. Hückel Theory+ Reorganization Energy= Marcus-Hush TheoryBreakdown of the 1/n Trend in ΠConjugated Poly-p-phenylene Cation Radicals Is Explained. J. Phys. Chem. C. 2017, 121, 15521561. (28) Ivanov, M. V.; Thakur, K.; Boddeda, A.; Wang, D.; Rathore, R. Nodal Arrangement of HOMO Controls the Turning On/Off the Electronic Coupling in Isomeric Polypyrene Wires. J. Phys. Chem. C. 2017, 121, 9202-9208. (29) Ivanov, M. V.; Thakur, K.; Bhatnagar, A.; Rathore, R. Isolation of a Chiral Anthracene Cation Radical: X-ray Crystallography and Computational Interrogation of Its Racemization. Chem. Commun. (Camb). 2017, 53, 2748-2751. (30) Chakraborty, R.; Bose, S.; Ghosh, D. Effect of Solvation on the Ionization of Guanine Nucleotide: A Hybrid QM/EFP Study. J. Comput. Chem. 2017, 38, 2528-2537.

ACS Paragon Plus Environment

21

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 22 of 24

(31) Bernini, C.; Zani, L.; Calamante, M.; Reginato, G.; Mordini, A.; Taddei, M.; Basosi, R.; Sinicropi, A. Excited State Geometries and Vertical Emission Energies of Solvated Dyes for DSSC: A PCM/TD-DFT Benchmark Study. J. Chem. Theory. Comput. 2014, 10, 3925-3933. (32) Kronik, L.; Stein, T.; Refaely-Abramson, S.; Baer, R. Excitation Gaps of Finite-Sized Systems From Optimally Tuned Range-Separated Hybrid Functionals. J. Chem. Theory. Comput. 2012, 8, 1515-1531. (33) Dreuw, A.; Head-Gordon, M. Failure of Time-dependent Density Functional Theory for Long-range Charge-transfer Excited States: The Zincbacteriochlorin-bacteriochlorin and Bacteriochlorophyll-spheroidene Complexes. J. Am. Chem. Soc. 2004, 126, 4007-4016. (34) Krylov, A. I. Equation-of-motion Coupled-cluster Methods for Open-shell and Electronically Excited Species: The Hitchhiker's Guide to Fock Space. Annu. Rev. Phys. Chem. 2008, 59, 433-462. (35) Pieniazek, P. A.; Arnstein, S. A.; Bradforth, S. E.; Krylov, A. I.; Sherrill, C. D. Benchmark Full Configuration Interaction and Equation-of-motion Coupled-cluster Model with Single and Double Substitutions for Ionized Systems Results for Prototypical Charge Transfer Systems: Noncovalent Ionized Dimers. J. Chem. Phys. 2007, 127, 164110. (36) Pieniazek, P. A.; Bradforth, S. E.; Krylov, A. I. Charge Localization and Jahn-Teller Distortions in the Benzene Dimer Cation. J. Chem. Phys. 2008, 129, 074104. (37) Ivanov, M. V.; Wang, D.; Wadumethridge, S. H.; Rathore, R. When Substituents Don't Matter: Frontier Orbitals Explain Unusually High and Invariant Oxidation Potential in Alkoxy-, Alkyl- and H-substituted Iptycenes. J. Phys. Chem. Lett. 2017, .

ACS Paragon Plus Environment

22

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

(38) Talipov, M. R.; Boddeda, A.; Lindeman, S. V.; Rathore, R. Does Koopmans' Paradigm for 1-Electron Oxidation Always Hold? Breakdown of IP/Eox Relationship for P-Hydroquinone Ethers and the Role of Methoxy Group Rotation. J. Phys. Chem. Lett. 2015, 6, 3373-3378. (39) Note that the small HOMO/HOMO-1 energy gap (i.e., 0.1 eV) in p-DBC may allow the coexistence of two distinct electronic structures of p-DBC+•, a subject matter which is under investigation. (40) Brunschwig, B. S.; Creutz, C.; Sutin, N. Optical Transitions of Symmetrical Mixed-valence Systems in the Class II--III Transition Regime. Chem. Soc. Rev. 2002, 31, 168-184. (41) Lambert, C.; Nöll, G. The Class II/III Transition in Triarylamine Redox Systems. J. Am. Chem. Soc. 1999, 121, 8434-8442. (42) Heckmann, A.; Lambert, C. Organic Mixed-valence Compounds: A Playground for Electrons and Holes. Angew. Chem. Int. Ed. Engl. 2012, 51, 326-392. (43) Lambert, C.; Nöll, G.; Schelter, J. Bridge-mediated Hopping or Superexchange Electrontransfer Processes in Bis(triarylamine) Systems. Nat. Mater. 2002, 1, 69-73. (44) Talipov, M. R.; Navale, T. S.; Hossain, M. M.; Shukla, R.; Ivanov, M. V.; Rathore, R. Dihedral Angle-Controlled Crossover From Static Hole Delocalization to Dynamic Hopping in Biaryl Cation Radicals. Angew. Chem. Int. Ed. 2017, 129, 272-275.

ACS Paragon Plus Environment

23

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 24 of 24

TOC GRAPHICS

ACS Paragon Plus Environment

24