Influence of push–pull configuration on the electro-optical and charge transport properties of novel naphtho-difuran derivatives: a DFT study

Aijaz Rasool Chaudhry*ab, R. Ahmed*a, Ahmad Irfanc, Shabbir Muhammadb, A. Shaaria and Abdullah G. Al-Sehemicde
aDepartment of Physics, Faculty of Science, Universiti Teknologi Malaysia, UTM Skudai, 81310 Johor, Malaysia. E-mail: aijaz_bwp27@hotmail.com; rashidahmed@utm.my; Tel: +60 17 734 1953
bDepartment of Physics, Faculty of Science, King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia
cDepartment of Chemistry, Faculty of Science, King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia
dUnit of Science and Technology, Faculty of Science, King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia
eCenter of Excellence for Advanced Materials Research, King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia

Received 17th June 2014 , Accepted 19th September 2014

First published on 19th September 2014


Abstract

We present a density functional theory (DFT) study pertaining to electro-optical and charge transport properties of two novel derivatives of diphenyl-naphtho[2,1-b:6,5-b′]difuran (DPNDF) as investigated based on push-pull configuration. Both molecular structures of the designed derivatives were optimized, in ground state (S0) as well as excited state (S1), using DFT and time-dependent DFT (TD-DFT) respectively. The push-pull configuration effect was studied meticulously for different electro-optical properties including adiabatic/vertical electron affinity (EAa/EAv), adiabatic/vertical ionization potential (IPa/IPv) and hole/electron reorganization energies (λh/λe), hole/electron transfer integrals (Vh/Ve), hole/electron mobility and photostability. We observed smaller λe, improved Ve and higher electron mobility for compound 1 compared with the parent molecule. Our calculated value of the electron mobility for compound 1 (2.43 cm2 V−1 s−1) revealed it to be an efficient electron transport material. Moreover, the influence of the push-pull on the electronic structure was also investigated by calculating the total and partial density of states (DOS). Taking advantage of the strong push-pull configurations effect on other properties, the study of the designed chemical systems was extended to their nonlinear optical (NLO) properties. Our designed novel derivatives (1 & 2) exhibited significantly larger amplitude values for first hyperpolarizability with βtot equal to 209.420 × 10−30 esu for compound 1 and 333.830 × 10−30 esu for compound 2. It was found that the first hyperpolarizability and HOMO–LUMO energy gap are in an inverse relationship for compounds 1 and 2.


1. Introduction

Push–pull is an important strategy widely used in organic semiconductor materials (OSMs) to tune the photophysical properties.1–4 More recently, the push–pull strategy has been a good approach for enhancing the electronic and charge transport properties.5–8 A stronger push–pull effect represents a significant charge separation on molecular geometry and dipole moment.6 A push–pull configuration usually consists of an electron donating group (EDG), a π conjugation bridge and an electron withdrawing group (EWG), which is expressed as EDGs-π-EWGs. Such a type of push–pull configuration with strong EDGs and EWG results in lowering the HOMO–LUMO energy gap Eg, which leads to efficient intramolecular charge transfer9 and material performance that is beneficial for designing excellent OSMs. OSMs with push–pull configuration10–12 are widely studied at both levels (theoretical and experimental) due to their light weight and their low fabrication cost on flexible substrates and large area bendy displays. These advantages of OSMs give them an edge over the silicon-based traditional inorganic semiconductors and have attracted massive interest of academic researchers as well as their industrial partners because of their potential applications in photonics and electronic devices, such as organic field effect transistors (OFETs),13–15 organic light emitting diodes (OLEDs),16,17 organic photo-voltaics (OPVs)17,18 and organic light emitting transistors (OLETs).19

OFETs are fabricated and produced by adding an electron system with π-conjugation,20 or an aromatic compound21 which helps in the orbital wave functions delocalization22 and establishes a good relationship between the geometric and electronic structure.23–28 Several experimental and theoretical research reports are available on thiophene containing materials for use in OFETs and OLEDs.29–38 The push–pull effect on the electronic, optical and charge transport properties of the benzo[2,3-b]thiophene derivatives has been studied theoretically.6 However, in the literature only a very small number of investigations has been reported about the furan containing OFETs and OLEDs materials.39–42 Considerable attention is currently being given to furan as a basic building block for organic π-conjugated materials that are more stable and that have given an indication of their potential applications, especially in OFETs and OLEDs.19,43–48 Binaphtha-furanyl has been reported as an OLET.19 Recently, diphenyl-naphtho[2,1-b:6,5-b′]difuran (DPNDF) has been experimentally synthesized and reported as a good hole transport OSM for OFETs.49 In our previous study,50 it was also found that the furan ring is one of the best electron transport materials because it demonstrates very low reorganization energy for electron λe.

No study on DPNDF with push–pull strategy has been found in the literature so far. In our present work, as a starting point, the experimental crystal of DPNDF49 has been used as the parent molecule and two new structures were derived by employing a push–pull approach where EDGs were attached on one side of the naphtho-difuran (NDF) ring and EWGs on the other side (EDGs-π-EWGs). In these structures, three CH3/OCH3 groups were attached as EDGs at position X, whereas three CN groups were attached as EWGs at position Y for structures 1/2, respectively (named compound 1 and compound 2) (see Scheme 1 and Fig. S1 of the ESI).


image file: c4ra05850j-s1.tif
Scheme 1 Labeled diagram of DPNDF showing the positions of the attached groups.

The derived geometries were optimized in ground state (S0) and first excited states (S1) at the level of DFT and TD-DFT, respectively. Also, the other properties such as highest occupied molecular orbitals (HOMOs), lowest unoccupied molecular orbitals (LUMOs), HOMO–LUMO energy gap (Eg), adiabatic and vertical electron affinities (EAa/EAv), reorganization energies for hole (λh)/electron (λe), adiabatic/vertical ionization potentials (IPa/IPv), total/partial density of states (TDOS/PDOS), nonlinear optical properties (NLO), hole extraction potential (HEP), electron extraction potential (EEP) and electronegativity (χ) were computed and are discussed in detail. In addition, the transfer integrals, mobility and photostability of these compounds were evaluated. Moreover, the push–pull effect was investigated on the above mentioned properties. A comparison of our obtained results with experimental data is also made where available.

2. Computational methodology

DFT was used to optimize the initial molecular structures for S0 by applying the hybrid exchange correlation functional B3LYP51–53 with 6-31G** basis sets.30,39–42,54–56 For S1 TD-DFT,56–59 the hybrid functional TD-B3LYP60–63 with the same basis set was used to optimize the geometries of the analogues. The electronic and photophysical properties including absorption (λabs), and emission (λemis) wavelengths were calculated at the same level of theory. Reorganization energy (λ) represents the geometric relaxation energy of a molecule from the charged (cation/anion) to the neutral state and from neutral to the charged (cation/anion) state (for details see the computational method expressed in the ESI). The reorganization energy for hole (λh) and electron (λe) was evaluated as:
 
λh = λ+ + λ1 and λe = λ + λ2 (1)
where the energy of geometry relaxation from neutral to charged (cation/anion) state is λ+ and λ, and the relaxation energy of a molecule from charged (cation/anion) state to neutral is λ1 and λ2, respectively.64,65 These two terms were calculated directly from the adiabatic potential energy surfaces for λh and λe.1,66,67 In the next step, the calculations related to term transfer integrals were performed. To calculate the transfer integrals, inter-molecular nearest-neighboring hopping pathways were generated using the single-crystal structure. There are two widely employed approaches to obtain transfer integrals; one is a Koopmans' theorem based method68 and the other one is a direct evaluation method for the frontier molecular orbitals (FMOs).69,70 We used the direct approach69,70 to investigate the charge transport properties in this study. The hole/electron transfer integrals in this approach can be expressed as:
 
th/e = 〈ϕ0,site1LUMO/HOMO|F0|ϕ0,site2LUMO/HOMO (2)
where th/e is the hole/electron transfer integrals, ϕ0,site1LUMO/HOMO and ϕ0,site2LUMO/HOMO correspond to the HOMOs and LUMOs of the two consecutive molecules when there is no contact between the adjacent molecules, and F0 is the Fock operator with unperturbed molecular orbitals for the dimer of a fixed pathway.

The carrier mobility μ can be evaluated with the help of the Einstein relation as:

 
μ = eD/kBT (3)
where D represents a charge diffusion constant, e is the electronic charge, T is the temperature and kB denotes the Boltzmann constant. Further details related to transfer integrals and mobility can be seen in the computational method expressed in the ESI. All these first-principles calculations were carried out using the GAUSSIAN 09 package.71

For investigation of NLO response, we calculated the static first hyperpolarizability (βtot) and its components by a finite field (FF) method. The FF method is broadly applied to investigate NLO because this methodology can be used with the electronic structure method to compute β values.72–81 In some very recent reports, βtot calculated by this method was found to be substantiated with the experimental structure–property relationship.10,82 In the FF method, a molecule is subjected to a static electric field (F) and the energy (E) of the molecule is expressed as:

 
image file: c4ra05850j-t1.tif(4)
where E(0) is the energy of the molecule in the absence of an electronic field, μ is the component of the dipole moment vector, α is the linear polarizability tensor, β and γ are the first and second hyperpolarizability tensors respectively, and i, j and k label the x, y and z components respectively. It is clear from eqn (4) that the values of μ, α, β, and γ can be obtained by differentiating E with respect to F. In our present investigation, we calculated the molecular first hyperpolarizability. For a molecule, the components of the first hyperpolarizability can be calculated using the following:
 
image file: c4ra05850j-t2.tif(5)

Using the x, y and z components, the magnitude of first hyperpolarizability (βtot) can be calculated by the following:

 
image file: c4ra05850j-t3.tif(6)
where
βx = (βxxx + βxxy + βxyy),

βy = (βyyy + βxxz + βyyz),

βz = (βxzz + βyzz + βzzz)

Therefore, the complete equation for calculating the magnitude of the total first static hyperpolarizability from GAUSSIAN 09 outputs is given by:

 
βtot = [(βxxx + βxxy + βxyy)2 + (βyyy + βxxz + βyyz)2 + (βxzz + βyzz + βzzz)2]1/2 (7)

Since these βtot values of the GAUSSIAN 09 output files are reported in atomic units (a.u.), the calculated βtot values were converted into electrostatic units (esu) (1 a.u. = 8.6393 × 10−33 esu). First hyperpolarizability is a third rank tensor that can be described by a 3 × 3 × 3 matrix. The 27 components of the 3D matrix can be reduced to 10 components due to the Kleinman symmetry (βxyy = βyxy = βyyx, βyyz = βyzy = βzyy,… likewise other permutations also take same value).

3. Results and discussion

3.1. Ground and excited state geometries

The optimized values of bond lengths and bond/dihedral angles for the neutral, cation and anion structures are tabulated in Table S1 of the ESI. For compound 1/2, the cation and anion structures differ from the neutral one. For compound 1, alterations in the bond lengths C19–C20, C20–C21, C21–C22, C22–C23, C23–C24 and C23–C28 were found; for the cation as 0.021 Å, −0.032 Å, 0.37 Å, −0.033 Å, 0.026 Å and 0.023 Å; for the anion as −0.003 Å, 0.00 Å, 0.001 Å, −0.001 Å, 0.004 Å and 0.002 Å, respectively; the bond angles C10–C9–C14, O18–C22–C21 and O18–C22–C23 were altered as 1.02°, −1.08° and 1.14° for the cation and as −1.86°, −0.24° and −0.04° for the anion, respectively. Similarly the distortion in the dihedral angles C20–C21–C22–C23, O18–C22–C23–C24, O18–C22–C23–C28, C21–C22–C23–C24 and C21–C22–C23–C28 was found as −1.13°, 23.23°, −23.53°, −22.09° and 22.40° for the cation and as −0.69°, 4.25°, −4.34°, −3.64° and 3.74° for the anion, respectively, as compared with the neutral ones. A graphical representation of bond lengths in angstrom (Å) for compound 1 (left) and compound 2 (right) is shown in Fig. 1(a) for a more clear understanding of the bond length alteration. The optimized coordinates of the neutral, cation and anion structures for both the compounds are tabulated in Tables S3–S5, respectively, in the ESI.
image file: c4ra05850j-f1.tif
Fig. 1 (a) Selected optimized bond lengths in angstrom (Å) for compound 1 (left) and compound 2 (right) of ground state (neutral, cation and anion) optimized at B3LYP/6-31G** level of theory. (b) Selected optimized bond angles (degree) for compound 1 (left) and compound 2 (right) of ground state (neutral, cation and anion) optimized at B3LYP/6-31G** level of theory.

For compound 2, the bond lengths C9–C10, C9–C14 and C11–C12, varied; for the cation as −0.007 Å, −0.007 Å and 0.00 Å; for the anion as −0.034 Å, 0.031 Å and 0.013 Å, respectively; the bond angles O1–C8–C7, C8–C9–C14, C10–C9–C14, C10–C11–C12, C11–C12–C13 and C12–C13–C14 were altered as 0.67°, −0.34°, 1.02°, −0.37°, 0.58° and −0.47° for the cation and as −1.19°, 1.08°, −1.83°, 1.32°, −1.50° and 1.09° for the anion, respectively; similarly the distortions in the dihedral angles C5–C1–C8–C9, C6–C7–C8–C9, C1–C8–C9–C10, C1–C8–C9–C14, C7–C8–C9–C10 and C7–C8–C9–C14 were found as 0.37°, 0.66°, −10.12°, 10.12°, 9.39° and −9.40° for the cation and as 1.33°, 1.65°, 17.01°, −16.75°, −18.69° and 18.43° for the anion, respectively, as compared with the neutral ones.

The bond/dihedral angles (degree) are represented graphically for compound 1 (left) and compound 2 (right) in Fig. 1(b), for a more clear understanding of the bond/dihedral angle distortion. The relaxation in the geometric parameters of the compound 1 cation structure was found to be more than that of the anion, whereas for compound 2 the anion had more distortion as compared with the cation. This high distortion especially in bond/dihedral angles83 might increase the reorganization energy of the compounds due to the increased polarization caused by this distortion. Generally, it has been stated that more relaxation in geometric parameters from neutral to anion/cation can increase the reorganization energy.2,83

3.2. Electronic properties

3.2.1 Frontier molecular orbitals (ground and excited states). HOMO and LUMO formation patterns for both the compounds at S0 and S1 were formed at isosurface values of 0.02 and are presented in Fig. 2(a) and (b). In compound 1, for HOMO formation, it was found that the charge is delocalized on C7–C8, C5–C6, C3–C4, C2–C20–C19, C16–C17, C21–C22, C23–C24 and C25–C26 while on C10, C12 and C14 the charge is localized (lone-pair). Neither of the O atoms take part in the formation of the HOMO, whereas a lone-pair is formed on all N atoms. For LUMO formation, the delocalization of charge was found on C2–C3, C5–C6–C15, C7–C8, C8–C9, C10–C11, and C13–C14. The charge is localized (lone-pair) on C17, C20 and on all N atoms. Similar patterns of HOMO and LUMO formation were found for compound 2 at S0. Charge delocalization and localization behavior followed the same trend for compound 1/2 at S1. It is clear from Fig. 2(a) that in the formation of HOMOs all the charge density is distributed on EDGs and the central core; for LUMOs the charge density is shifted on EWGs, revealing good intramolecular charge transfer for both the compounds.
image file: c4ra05850j-f2.tif
Fig. 2 (a) HOMOs and LUMOs formation patterns at ground state. (b) HOMOs and LUMOs spreading patterns at excited state.

The HOMO energies (EHOMO), LUMO energies (ELUMO) and HOMO–LUMO energy gaps (Eg) at S0 and S1 (in the brackets) states for both the compounds are tabulated in Table 1. A graphical representation of EHOMO and ELUMO for S0/S1 is shown in Fig. 3 left and right, respectively, for a more clear understanding of the Eg. EHOMO of compound 1 and 2 are −5.65 eV and −5.18 eV, respectively, and are in good agreement with the experimental EHOMO (−5.48 eV)49 and the computational EHOMO (−5.10 eV)84 of the parent molecule DPNDF. The trend of EHOMO and ELUMO, respectively, is compound 1 (−5.65 eV, −2.93 eV) > compound 2 (−5.18 eV, −2.78 eV). A similar trend of EHOMO and ELUMO was found at S1 for both the compounds.

Table 1 The EHOMOa,b, ELUMO and Eg for S0 and S1 states (in the brackets) at the B3LYP/6-31G** and TD-B3LYP/6-31G** levels of theory
Molecule EHOMO ELUMO Eg
a Experimental data (EHOMO = −5.48 eV) from ref. 49.b Computed value (EHOMO = −5.10 eV) from ref. 84.
Compound 1 −5.65 −2.93 2.72
(−5.38) (−3.05) (2.33)
Compound 2 −5.18 −2.78 2.40
(−4.86) (−2.86) (2.00)



image file: c4ra05850j-f3.tif
Fig. 3 Comparison of EHOMO and ELUMO for ground (S0) state (left) and excited (S1) state (right) at the B3LYP/6-31G** and TD-B3LYP/6-31G** levels of theory.

The energy gap is theoretically expressed as the difference of the orbital energies between HOMO and LUMO whereas experimentally it is the lowest energy transition from the S0 to the S1 state, and termed as the band gap, which can be obtained from the absorption spectra. When electron promotion takes place from HOMO to LUMO, quantitatively Eg can be approximately the same as the optical band gap.1,85–87 The trend in the Eg is as compound 1 (2.72 eV) > compound 2 (2.40 eV) for S0 and for S1 it is compound 1 (2.33 eV) > compound 2 (2.00 eV). The smaller Eg of compound 2 reveals the red shift in the absorption and emission wavelengths in comparison with compound 1. Similarly, the lower Eg of compound 2 illuminates the high charge transfer interaction9,88 within the compound. A high amplitude of first hyperpolarizability (β) correlates9,88 with a lower Eg of a compound. We anticipated that compound 2 might show larger first hyperpolarizability (β) as compared with compound 1. It has been reported earlier that an electronic system with a smaller Eg might be more reactive83,84 than one with a larger Eg, so compound 2 may be more reactive than compound 1, revealing that the latter one would be more stable.

It is expected that compounds having a low-lying LUMO energy level might be thermodynamically more stable and charge transfer could not be quenched by electron loss. Moreover, according to Koopman's theorem the LUMO energy is directly proportional to the EA. The higher LUMO energy level is illuminating that the electron injection barrier would be small resulting in the improvement of charge injection ability. It can be seen from Table 1 that the value of ELUMO is increased in compound 1 and 2 as compared with the computed value of ELUMO (−2.17 eV)84 for the parent molecule DPNDF, which would decrease the electron injection barrier resulting in an improvement of the electron injection. Therefore it is expected that the new compounds might be better materials as electron transporters.

3.3. Photophysical properties

The calculated absorption (λabs) and emission wavelengths (λemis), oscillator strengths (f) and HOMO–LUMO (H → L) contribution were evaluated and are tabulated in Table 2. The λabs and λemis against f are represented graphically in Fig. 4(a) and (b), respectively. Table 2 shows the maximum H → L contribution at the S0, which is 99% from H → L for both compounds 1 and 2. The maximum contribution of H → L for S1 is H → L (99%) and H → L (100%) for compounds 1 and 2, respectively. The λabs/λemis have a red shift of 131/180 nm for compound 1 and 207/294 nm for compound 2, respectively, as compared with the parent molecule of DPNDF (λabs/λemis 381/427 nm) as evaluated computationally,84 whereas compound 2 has a red shift of 76/114 nm as compared with compound 1 for λabs/λemis, respectively. This might be due to the strong EDGs and EWGs attached to compound 2. The structure–property relationship revealed that by substituting the EDGs and EWGs, the λabs and the λemis have shown red shifted behavior. Compound 1 would be an orange light emitter while compound 2 would be a red light emitter.
Table 2 Calculated absorption (λabs) and emission (λemis) wavelengthsa (nm), oscillator strength (f) and HOMO–LUMO contribution for S0 and S1 states at the TD-DFT level of theories
Molecule λabs f Contribution λemis f Contribution
a Computed values (λabs = 381 nm; λemis = 427 nm) for comparison with ref. 84.
Compound 1 512 0.467 H → L (99%) 607 0.373 H → L (99%)
Compound 2 588 0.342 H → L (99%) 721 0.278 H → L (100%)



image file: c4ra05850j-f4.tif
Fig. 4 (a) Computed absorption spectra and (b) computed emission spectra of compounds 1 and 2 at TD-B3LYP/6-31G** levels of theory.

3.4. Density of states

As remarkable electro-optical properties are attributed to a push–pull configuration in designed chemical systems, we also calculated explicit contributions for the individual parts in the form of their PDOS as shown in Fig. 5. We define three fragments for each compound: fragment one contains the phenyl ring with EWGs; fragment two is the central core (CC); and fragment three consists of the phenyl ring with EDGs. The individual fragment represents its contribution to the TDOS of the whole molecule as shown in Fig. 5 with different curves. As shown in Fig. 5, for compound 1, the peaks from −15.0 to −8.0 eV for the valence band and from 0.0 to 7.0 eV for the conduction band are due to the similar contributions from EWGs, CC and EDGs. At the HOMO between −6.0 and −4.0 eV, the major contribution is from CC, whereas EDGs have a minor contribution. On the other hand, the EWGs take maximum part in the conduction band (−3.0 and −2.0 eV) while the CC has minimum contribution. EDGs have no contribution in the lower region of the conduction band while they have significant contribution between −1.0 to 0.0 eV and 2.0 to 7.0 eV. In TDOS the EDG contribution dominates in the lower valence bands from −13.0 to −9.0 eV and the higher conduction bands from 2.0 to 7.0 eV. The contribution of EDGs is more in the lower energy bonding molecular orbitals (−5.65 to −7.0 eV) while the EWG contribution is larger in the higher energy anti-bonding molecular orbitals (−2.93 to −3.5 eV), which facilitates as easy charge transfer during the transition process of this push–pull configuration. This contribution of TDOS/PDOS from valence and conduction bands revealed good intramolecular charge transport from EDGs to EWGs. The high intramolecular charge transport from EDGs to EWGs leads to a very large value of β. A similar trend for TDOS and PDOS was found in compound 2.
image file: c4ra05850j-f5.tif
Fig. 5 Graphical representation of TDOS and PDOS for compound 1 (left) and compound 2 (right) computed at the B3LYP/6-31G** level of theory.

3.5. Charge transfer properties

The EA and IP are the most essential properties to calculate the charge transport barriers, and these were evaluated at the DFT/B3LYP/6-31G** level of theory. In OSMs, a lower IP and higher EA is very crucial to enhance the charge transport ability for electron and hole, respectively. The adiabatic/vertical IP (IPa/IPv) and adiabatic/vertical EA (EAa/EAv) of all derivatives were calculated and are tabulated in Table 3. A graphical comparison of the IPv, electronegativity and EAv is shown in Fig. 6 (left) to represent the results more clearly. In OFETs, the OSMs having high EAv and small IPv might be better for n-type and p-type charge injection, respectively.89 From Table 3, it can be seen that compounds 1 and 2 have EAv of 1.63 and 1.49 eV, respectively, which are higher than that of the parent molecule DPNDF (0.29 eV).84 Thus it is expected that the new designed compounds might be much better electron transport materials as compared with DPNDF. The EAv follow the same trend as ELUMO for both compounds as it has been observed that the molecule with high ELUMO has the higher EAv. It can be seen from Tables 1 and 3 that compound 1 has the highest ELUMO (−2.93 eV) among the two compounds and hence has the highest EAv (1.63 eV).
Table 3 The vertical/adiabatic ionization potential and electron affinity, hole extraction potential, electron extraction potential, electronegativity and reorganization energy for holea λh/electron λe at the B3LYP/6-31G** level of theory. All values are in eV
  Compound 1 Compound 2
a Data for comparison (λh = 0.17 eV) from ref. 49.
IP (vertical) 6.85 6.37
IP (adiabatic) 6.68 6.15
HEP 6.53 5.95
EA (vertical) 1.63 1.49
EA (adiabatic) 1.76 1.61
EEP 1.87 3.36
χ 4.24 3.93
λh 0.32 0.42
λe 0.24 1.87



image file: c4ra05850j-f6.tif
Fig. 6 Graphical representation of IPv, electronegativity and EAv (left) and of λh and λe (right) calculated at the B3LYP/6-31G** level of theory.

image file: c4ra05850j-f7.tif
Fig. 7 The definition of the Cartesian axis for optimized structures of compounds 1 and 2.

The reorganization energy is a quantity which is very important for estimation of the ability to carry the charge in a solid.90,91 The reorganization energies at the B3LYP/6-31G** level of theory for electron λe/hole λh are given in Table 3. A graphical representation of hole λh and λe is shown in Fig. 6 (right) to represent the trend for further clarity. The calculated λh of the DPNDF is 0.17 eV (ref. 84) at the same level of theory and is in good agreement with the already computed value.49 From Table 3, it can be seen that compounds 1 and 2 have λh of 0.32 and 0.42 eV, respectively, and λe of 0.24 and 1.87 eV, respectively. For compound 1 λh is higher than λe, whereas for compound 2 λh is lower than λe. The alteration and distortion in the bond/dihedral angles of the cation is more than that of the anion for compound 1, resulting in more polarization2,83 so λe for compound 1 is less than λh. On the other hand, for compound 2, the bond/dihedral angle distortion in the anion is higher than in the cation; it might be due to this that the λe of compound 2 is much higher than λh. From this trend it is predicted that compound 1 would be good as an electron-transport material; and compound 2 might be good as a hole-transport material. The value of λe for compound 1 is smaller than those of diphenyl-naphtho-dithiophene (0.34 eV)92 and oligofuran (0.40 eV),93 which indicates that the new designed compound 1 might be more efficient as an electron transport material.

The electronegativities (χ) of the two compounds are given in Table 3. The electronegativity is the power of an atom in a molecule to attract electrons towards itself. A molecule with high electronegativity might be more efficient as an electron transport material because it can pull more electrons towards itself, resulting in high electron charge transfer.57,71–73,94 The trend of electronegativities in compound 1 > compound 2 revealed that the former might be better as an electron transport material as compared with the latter. The reorganization energies decreased with the increase in electronegativities of the compounds. It is the same trend as for ELUMO and EAv of the two compounds.

3.6. First hyperpolarizability

It is well known that push–pull chemical configurations usually show remarkable NLO responses. In the present investigation taking advantage of strong push–pull configurations, we have also spotlighted the NLO responses of our designed chemical systems by calculating their static first hyperpolarizabilities. The calculated values of hyperpolarizability (β) along with their individual tensor components are shown in Table 4. A well-established electronic communication of two different parts of a push–pull molecule usually accompanies a larger amplitude of its first hyperpolarizability, which is perhaps the case in our present designed compounds 1 and 2 Fig. 7. In Table 4, it can be seen that the calculated amplitudes of first hyperpolarizability (βtot) for compound 1 and 2 are significantly larger with βtot values of 209.420 × 10−30 esu and 333.830 × 10−30 esu, respectively. These values of first hyperpolarizability of compound 1 and 2 are much larger than that of proto-type urea molecule [β for urea is 0.3728 × 10−30 esu]. These total hyperpolarizability values are dominated by their diagonal components (components along the dipole moment axis) of βxxx. This is because there is significant charge transfer from EDGs to EWGs along the x-axis. It can be seen from Tables 1 and 4 that the first hyperpolarizability and Eg are in inverse relationship for both compounds, which supports our prediction on the basis of Eg. Thus our designed systems have significant potential for NLO applications with good stability and large first hyperpolarizability amplitudes.
Table 4 The calculated values of polarizability (α) and hyperpolarizability (β) along their individual tensor components
Compound 1 Compound 2
Component a.u. (×10−30) esu Component a.u. (×10−30) esu
βxxx −24 305 200.100 βxxx −38 827 −335.500
βxxy −676 −5.841 βxxy −881 −7.613
βxyy −59 −0.509 βxyy 119 1.028
βyyy 7 0.0604 βyyy 17 0.146
βxxz 102 0.881 βxxz −362 −3.128
βxyz −90 −0.777 βxyz −103 −0.890
βyyz −30 −0.259 βyyz −8 −0.069
βxzz 137 1.183 βxzz 86 0.743
βyzz 11 0.095 βyzz −3 −0.025
βzzz 0 0.000 βzzz 18 0.155
βtot 24 236 209.420 βtot 38 633 333.830


3.7. Molecular simulation

In our previous study,95 we optimized the initial geometry of the parent molecule at S0 by a hybrid functional B3LYP along with the 6-31G** basis set using the GAUSSIAN 09 package. The crystal structure was simulated using facilities provided within the Materials Studio (MS) package using the same lattice parameters as used for the experimental crystal of DPNDF.49 The crystal was simulated using a Molecular Mechanics (MM) simulation approach, and the energy of the crystal was minimized by the FORCITE module96 with the P21/c space group as available in the MS package, which is considered to be a good tool for this purpose. The DREIDING force field97 was used, which is suitable for these kinds of OMs with C, H, O, and N atoms. The simulated crystal structure was found to be in good agreement with the experimentally synthesized structure (see Fig. 8). In that study we described the four pathways to compute the transfer integrals and mobility. Previously, the main computed transfer integral for the parent molecule DPNDF was 36.9 meV (ref. 95) using the GAUSSIAN package at B3LYP/6-31G** level and was closer to the available data of the same crystal 39.9 meV evaluated by the ADF program.49 Similarly, the computed mobility was (1.1 cm2 V−1 s−1), which shows good agreement with the experimentally measured value (1.30 cm2 V−1 s−1). These results revealed that our adopted approach was reliable to build the crystal, and to compute the transfer integrals and mobility. In the current study, the same approach has been used to simulate the crystal structures for the new designed compounds.
image file: c4ra05850j-f8.tif
Fig. 8 Experimental (a) and simulated (b) crystal structure of DPNDF along (aoc) direction.

3.8. Transfer integrals

We have also evaluated four discrete nearest neighboring hopping pathways for the two compounds. Transfer integrals for electron and hole have been evaluated using the method expressed in eqn (2) and presented in Table 5. A graphical comparison of hole and electron transfer integrals is shown in Fig. 9(a) (left) for a more clear representation. It can be seen that some transfer integrals have negligibly small values so they are not discussed further here. The strongest hole/electron transfer integrals for compound 1 are 65.1/118.4 meV and for compound 2 are 12.9/−92 meV, respectively. Compound 1 has higher electron transfer integrals than compound 2, revealing that compound 1 is a better electron transport material than compound 2.
Table 5 The transfer integrals (meV), mass centers (Å) and mobilities (cm2 V−1 s−1) for hole and electron for compound 1 and 2, computed with DFT
Molecules Pathways Transfer integralsc Mass centers Mobilityc
Vha Ve Holeb Electron
a Vh (39.9 meV) by other method from ref. 49.b Experimental hole mobility (1.30 cm2 V−1 s−1) from ref. 49.c Computed values from ref. 95.
Compound 1 i 65.1 114.1 5.071 0.49 2.09
ii −20.6 118.4 5.078 4.89 × 10−3 2.43
iii −2.9 6.2 7.661 4.37 × 10−6 4.15 × 10−5
iv −15.9 5.7 19.106 2.46 × 10−2 1.85 × 10−4
Compound 2 i 12.9 67 5.078 5.89 × 10−3 2.66 × 10−8
ii −6.5 −92 7.661 8.64 × 10−4 2.15 × 10−7
iii 2.65 × 10−3 6.64 × 10−3 11.969 5.84 × 10−17 1.43 × 10−23
iv −0.15738 0.21216 19.108 1.85 × 10−9 3.79 × 10−17



image file: c4ra05850j-f9.tif
Fig. 9 (a) Graphical representation hole/electron transfer integrals (left) and mobilities (right) computed with DFT. (b) The dimers investigated in the present study to calculate the transfer integrals and mobilities.

3.9. Mobility

Hole and electron mobilities of both compounds for four pathways were calculated and are tabulated in Table 5. A graphical representation of hole and electron mobilities is shown in Fig. 9(a) (right) for a more clear comparison. In our previous study95 we computed the mobility of the parent crystal DPNDF using the direct method and found the hole mobility to be 1.1 cm2 V−1 s−1, which is in good agreement with the experimentally calculated mobility of 1.30 cm2 V−1 s−1 of the DPNDF.49 We have used the same method for our current study for the four nearest neighboring molecules and found that some of the pathways have very low mobility for hole and electron; hence not discussed in the text; only the highest hole/electron mobilities for two pathways of each compound are discussed in detail here. The hole mobilities of the two pathways were found as 0.49 and 4.89 × 10−3 cm2 V−1 s−1 for compound 1 and as 5.89 × 10−3 and 8.64 × 10−4 cm2 V−1 s−1 for compound 2, whereas the electron mobilities of the two pathways were found as 2.09 and 2.43 cm2 V−1 s−1, for compound 1 and as 2.66 × 10−8 and 2.15 × 10−7 cm2 V−1 s−1 for compound 2. It can be seen that compound 1 exhibits the highest electron mobility of 2.43 cm2 V−1 s−1 for specific 2nd pathway, which is higher than the already computed electron mobility of 1.10 cm2 V−1 s−1 of the parent molecule DPNDF.

The four pathways are shown in Fig. 9(b) for a more clear understanding of dimers packing which may affect the mobilities.30,49 These highest electron mobility values of the 1st and 2nd pathways are in the stacking direction and might be due to the smallest distance between the two molecules of the dimers. This packing and distance allowed maximum overlapping of the orbitals ensuring enhanced mobilities, whereas the 3rd and 4th pathways have side to side packing and greater distance between two molecules, which might be the reason for the lowest mobilities.

The highest electron mobilities for compound 1 are several times higher than DPNDF49 and α-oligofuran93 0.0134 cm2 V−1 s−1, hence the former is predicted to be a good electron transport material in comparison with DPNDF and α-oligofuran. It might be due to the attached EWGs (-CN), the push–pull effect and the comprehensive intramolecular charge transfer from donor to acceptor moieties. From the highest average electron/hole intrinsic mobilities (1.13/0.13) and (6.08 × 10−8/1.69 × 10−3) cm2 V−1 s−1 for compound 1 and 2, respectively, we anticipate that compound 1 is a good electron transport material; compound 2 may be a hole transport material. These results support our prediction about the same materials in terms of ELUMO, IPv, EAv and reorganization energies for hole and electron.

3.10. Photostability

Molecular electrostatic surface potentials of all compounds were mapped onto a total electron density surface as shown in Fig. 10. High electron density regions and low electron density regions are shown in indigo and green, respectively. High electron density is distributed on O and N atoms in the studied systems. Previously, the photostability of organic materials has been explained on the basis of molecular electrostatic surface potentials.56 Recently, we pointed out that more electron density distributed on the system would favor enhanced photostability.98 In our last study we observed that in the parent molecule DPNDF the high electron density is distributed on oxygen atoms95 (see Fig. S2 of the ESI). By substituting the EDGs and EWGs, the photostability was augmented in our designed molecules. The substitution of –CH3 and –CN at the outer phenyl rings in compound 1 improved the photostability compared with the parent DPNDF molecule. High electron density distribution in compound 2 covered more area due to the –OCH3 and –CN than in compound 1, revealing that the former would be more photostable than the latter, that is due to strong EDGs (–OCH3) attached with compound 2. We noticed that by increasing the size and strength of EDGs, the photostability might be enhanced. Higher electron density in compound 2 would decrease the oxidation resulting in improved photostability, which is in good agreement with our previous study.99
image file: c4ra05850j-f10.tif
Fig. 10 Molecular electrostatic potential surfaces of the two compounds.

4. Conclusions

Push–pull configurations have shown interesting effects for tuning the electro-optical properties of compounds 1 and 2. In the light of our present DFT investigation, we can draw the following interesting conclusions:

1. The two designed novel compounds have higher EAv values as compared with their parent molecule DPNDF.

2. The absorption and emission spectra of designed compounds 1 and 2 have a red shift as compared with their parent molecule. This is because of well-established communication between donor and acceptor parts. The HOMOs and LUMOs in both the studied compounds are delocalized as well as localized on the central core and EWGs, respectively. The EWGs take part in the establishment of LUMOs only. The HOMO energies are in good agreement with the experimentally estimated HOMO of the parent molecule DPNDF.

3. The influence of the push–pull parts has been investigated by calculating their total and partial density of states (DOS).

4. Taking advantage of strong push–pull configurations, our designed chemical systems have also been rationalized as efficient NLO materials with significantly larger amplitudes of first hyperpolarizability for compounds 1 and 2.

5. The first hyperpolarizability and HOMO–LUMO energy gap are in inverse relationship for compounds 1 and 2.

6. The electron transfer integrals and electron mobility have been enhanced significantly in compound 1 by introducing the –CH3 and –CN groups, respectively. So it is predicted that compound 1 would be a good electron transport material as compared to compound 2 and the parent molecule DPNDF.

7. The photostability has been enhanced significantly in compound 2 by introducing the –OCH3 and –CN groups, because of enriched electron density distributed on these groups. As a result it is predicted that compound 2 would be more stable than compound 1.

Hence, we expect that these compounds would serve as excellent candidates for OFET, OLET and OLED applications with enhanced photostability.

Acknowledgements

The authors are grateful to the Ministry of Education/Universiti Teknologi Malaysia (UTM) for providing funding via project Q.J130000.2526.06H15 for the successful execution of this project and to King Khalid University (KKU) for providing the support and facilities to complete the research study.

References

  1. Y. Li, L.-Y. Zou, A.-M. Ren and J.-K. Feng, Comput. Theor. Chem., 2012, 981, 14 CrossRef CAS PubMed.
  2. A. Irfan, R. Cui, J. Zhang and L. Hao, Chem. Phys., 2009, 364, 39 CrossRef CAS PubMed.
  3. A. É. Masunov and P. M. Zorkii, J. Struct. Chem., 1992, 33, 423 CrossRef.
  4. S. Ananthavel and M. Manoharan, Chem. Phys., 2001, 269, 49 CrossRef CAS.
  5. A. Irfan, A. G. Al-Sehemi and S. Muhammad, Synth. Met., 2014, 190, 27 CrossRef CAS PubMed.
  6. A. Irfan, A. G. Al-Sehemi and M. S. Al-Assiri, Comput. Theor. Chem., 2014, 1031, 76 CrossRef CAS PubMed.
  7. A. Irfan, A. G. Al-Sehemi and M. S. Al-Assiri, J. Fluorine Chem., 2014, 157, 52 CrossRef CAS PubMed.
  8. A. Irfan, A. G. Al-Sehemi and M. S. Al-Assiri, J. Mol. Graphics Modell., 2013, 44, 168 CrossRef CAS PubMed.
  9. H. Alyar, Rev. Adv. Mater. Sci., 2013, 34, 79 CAS.
  10. P. Srinivasan, T. Kanagasekaran and R. Gopalakrishnan, Cryst. Growth Des., 2008, 8, 2340 CAS.
  11. M. Rao, R. Ponce Ortiz, A. Facchetti, T. J. Marks and K. S. Narayan, J. Phys. Chem. C, 2010, 114, 20609 CAS.
  12. I. Osaka, T. Abe, M. Shimawaki, T. Koganezawa and K. Takimiya, ACS Macro Lett., 2012, 1, 437 CrossRef CAS.
  13. H. E. Katz, J. Mater. Chem., 1997, 7, 369 RSC.
  14. G. Horowitz and M. E. Hajlaoui, Adv. Mater., 2000, 12, 1046 CrossRef CAS.
  15. C. R. Newman, C. D. Frisbie, D. A. da Silva Filho, J.-L. Brédas, P. C. Ewbank and K. R. Mann, Chem. Mater., 2004, 16, 4436 CrossRef CAS.
  16. C. W. Tang and S. A. VanSlyke, Appl. Phys. Lett., 1987, 51, 913 CrossRef CAS PubMed.
  17. P. K. H. Ho, J.-S. Kim, J. H. Burroughes, H. Becker, S. F. Y. Li, T. M. Brown, F. Cacialli and R. H. Friend, Nature, 2000, 404, 481 CrossRef CAS PubMed.
  18. F. Padinger, R. S. Rittberger and N. S. Sariciftci, Adv. Funct. Mater., 2003, 13, 85 CrossRef CAS.
  19. K. Niimi, H. Mori, E. Miyazaki, I. Osaka, H. Kakizoe, K. Takimiya and C. Adachi, Chem. Commun., 2012, 48, 5892 RSC.
  20. A. W. Hofmann, Proc. R. Soc. London, 1856, 8, 1 CrossRef.
  21. IUPAC, Compendium of Chemical Terminology (the Gold Book), ed. A. D. McNaught and A. Wilkinson, 2nd edn, 1997 Search PubMed.
  22. P. v. R. Schleyer, Chem. Rev., 2005, 105, 3433 CrossRef CAS.
  23. J.-L. Brédas, D. Beljonne, V. Coropceanu and J. Cornil, Chem. Rev., 2004, 104, 4971 CrossRef PubMed.
  24. J.-L. Brédas, J. Cornil, D. Beljonne, D. A. dos Santos and Z. Shuai, Acc. Chem. Res., 1999, 32, 267 CrossRef.
  25. J. Cornil, D. A. dos Santos, X. Crispin, R. Silbey and J. L. Brédas, J. Am. Chem. Soc., 1998, 120, 1289 CrossRef CAS.
  26. J. L. Bredas and G. B. Street, Acc. Chem. Res., 1985, 18, 309 CrossRef CAS.
  27. A. Warshel and M. Karplus, J. Am. Chem. Soc., 1974, 96, 5677 CrossRef CAS.
  28. S.-C. Yang, W. Graupner, S. Guha, P. Puschnig, C. Martin, H. R. Chandrasekhar, M. Chandrasekhar, G. Leising, C. Ambrosch-Draxl and U. Scherf, Phys. Rev. Lett., 2000, 85, 2388 CrossRef CAS.
  29. A. Irfan, M. Nadeem, M. Athar, F. Kanwal and J. Zhang, Comput. Theor. Chem., 2011, 968, 8 CrossRef CAS PubMed.
  30. A. Irfan, A. G. Al-Sehemi, S. Muhammad and J. Zhang, Aust. J. Chem., 2011, 64, 1587 CrossRef CAS.
  31. M. S. Wrackmeyer, M. Hein, A. Petrich, J. Meiss, M. Hummert, M. K. Riede and K. Leo, Sol. Energy Mater. Sol. Cells, 2011, 95, 3171 CrossRef CAS PubMed.
  32. S. Das, S. P. Senanayak, A. Bedi, K. S. Narayan and S. S. Zade, Polymer, 2011, 52, 5780 CrossRef CAS PubMed.
  33. J. A. Letizia, S. Cronin, R. P. Ortiz, A. Facchetti, M. A. Ratner and T. J. Marks, Chem.–Eur. J., 2010, 16, 1911 CrossRef CAS PubMed.
  34. F. Buonocore and A. Matteo, Theor. Chem. Acc., 2009, 124, 217 CrossRef CAS PubMed.
  35. Q.-X. Wu, Y. Geng, Y. Liao, X.-D. Tang, G.-C. Yang and Z.-M. Su, Theor. Chem. Acc., 2012, 131, 1 CAS.
  36. K. N. N. Unni, S. Dabos-Seignon and J.-M. Nunzi, J. Mater. Sci., 2006, 41, 317 CrossRef CAS.
  37. P. Pingel, A. Zen, D. Neher, I. Lieberwirth, G. Wegner, S. Allard and U. Scherf, Appl. Phys. A, 2009, 95, 67 CrossRef CAS.
  38. H. Koezuka, A. Tsumura and T. Ando, Synth. Met., 1987, 18, 699 CrossRef CAS.
  39. Y. Miyata, T. Nishinaga and K. Komatsu, J. Org. Chem., 2005, 70, 1147 CrossRef CAS PubMed.
  40. Y. Miyata, M. Terayama, T. Minari, T. Nishinaga, T. Nemoto, S. Isoda and K. Komatsu, Chem.–Asian J., 2007, 2, 1492 CrossRef CAS PubMed.
  41. O. Gidron, A. Dadvand, Y. Sheynin, M. Bendikov and D. F. Perepichka, Chem. Commun., 2011, 47, 1976 RSC.
  42. C.-C. Wu, W.-Y. Hung, T.-L. Liu, L.-Z. Zhang and T.-Y. Luh, J. Appl. Phys., 2003, 93, 5465 CrossRef CAS PubMed.
  43. K. Kadac, M. J. Bosiak and J. Nowaczyk, Synth. Met., 2012, 162, 1981 CrossRef CAS PubMed.
  44. M. Watanabe, W.-T. Su, Y. J. Chang, T.-H. Chao, Y.-S. Wen and T. J. Chow, Chem.–Asian J., 2013, 8, 60 CrossRef CAS PubMed.
  45. M. Nakano, S. Shinamura, Y. Houchin, I. Osaka, E. Miyazaki and K. Takimiya, Chem. Commun., 2012, 48, 5671 RSC.
  46. M. Nakano, K. Niimi, E. Miyazaki, I. Osaka and K. Takimiya, J. Org. Chem., 2012, 77, 8099 CrossRef CAS PubMed.
  47. K. Mitsudo, J. Harada, Y. Tanaka, H. Mandai, C. Nishioka, H. Tanaka, A. Wakamiya, Y. Murata and S. Suga, J. Org. Chem., 2013, 78, 2763 CrossRef CAS PubMed.
  48. H. Chen, W. Delaunay, J. Li, Z. Wang, P.-A. Bouit, D. Tondelier, B. Geffroy, F. Mathey, Z. Duan, R. Réau and M. Hissler, Org. Lett., 2013, 15, 330 CrossRef CAS PubMed.
  49. C. Mitsui, J. Soeda, K. Miwa, H. Tsuji, J. Takeya and E. Nakamura, J. Am. Chem. Soc., 2012, 134, 5448 CrossRef CAS PubMed.
  50. A. R. Chaudhry, R. Ahmed, A. Irfan, A. Shaari and A. G. Al-Sehemi, Mater. Chem. Phys., 2013, 138, 468 CrossRef CAS PubMed.
  51. P. J. Stephens, F. J. Devlin, C. F. Chabalowski and M. J. Frisch, J. Phys. Chem., 1994, 98, 11623 CrossRef CAS.
  52. A. D. Becke, J. Chem. Phys., 1993, 98, 5648 CrossRef CAS PubMed.
  53. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785 CrossRef CAS.
  54. E. Cho, C. Risko, D. Kim, R. Gysel, N. Cates Miller, D. W. Breiby, M. D. McGehee, M. F. Toney, R. J. Kline and J.-L. Bredas, J. Am. Chem. Soc., 2012, 134, 6177 CrossRef CAS PubMed.
  55. T. Sajoto, S. P. Tiwari, H. Li, C. Risko, S. Barlow, Q. Zhang, J.-Y. Cho, J.-L. Brédas, B. Kippelen and S. R. Marder, Polymer, 2012, 53, 1072 CrossRef CAS PubMed.
  56. A. Irfan, J. Zhang and Y. Chang, Theor. Chem. Acc., 2010, 127, 587 CrossRef CAS.
  57. R. Bauernschmitt and R. Ahlrichs, Chem. Phys. Lett., 1996, 256, 454 CrossRef CAS.
  58. C. Van Caillie and R. D. Amos, Chem. Phys. Lett., 2000, 317, 159 CrossRef CAS.
  59. F. Furche and R. Ahlrichs, J. Chem. Phys., 2002, 117, 7433 CrossRef CAS PubMed.
  60. G. Scalmani, M. J. Frisch, B. Mennucci, J. Tomasi, R. Cammi and V. Barone, J. Chem. Phys., 2006, 124, 094107 CrossRef PubMed.
  61. J. E. Lee, G. C. Choi, N. G. Park, Y. K. Ha and Y. S. Kim, Curr. Appl. Phys., 2004, 4, 675 CrossRef PubMed.
  62. M. C. Zwier, J. M. Shorb and B. P. Krueger, J. Comput. Chem., 2007, 28, 1572 CrossRef CAS PubMed.
  63. E. Li, A. Kim, L. Zhang, Journal of Student Computational Chemistry, Spring, 2007, vol. 1, North Carolina School of Science and Mathematics Search PubMed.
  64. N. E. Gruhn, D. A. da Silva Filho, T. G. Bill, M. Malagoli, V. Coropceanu, A. Kahn and J.-L. Brédas, J. Am. Chem. Soc., 2002, 124, 7918 CrossRef CAS PubMed.
  65. J. R. Reimers, J. Chem. Phys., 2001, 115, 9103 CrossRef CAS PubMed.
  66. A. Irfan, R. Cui and J. Zhang, Theor. Chem. Acc., 2009, 122, 275 CrossRef CAS PubMed.
  67. V. Coropceanu, T. Nakano, N. E. Gruhn, O. Kwon, T. Yade, K.-i. Katsukawa and J.-L. Brédas, J. Phys. Chem. B, 2006, 110, 9482 CrossRef CAS PubMed.
  68. B. C. Lin, C. P. Cheng, Z.-Q. You and C.-P. Hsu, J. Am. Chem. Soc., 2004, 127, 66 CrossRef PubMed.
  69. A. Troisi and G. Orlandi, Chem. Phys. Lett., 2001, 344, 509 CrossRef CAS.
  70. S. Yin, Y. Yi, Q. Li, G. Yu, Y. Liu and Z. Shuai, J. Phys. Chem. A, 2006, 110, 7138 CrossRef CAS PubMed.
  71. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09 Revision A.02, Gaussian Inc., Wallingford, CT, 2009 Search PubMed.
  72. C. Dehu, F. Meyers and J. L. Bredas, J. Am. Chem. Soc., 1993, 115, 6198 CrossRef CAS.
  73. S. Muhammad, M. R. S. A. Janjua and Z. Su, J. Phys. Chem. C, 2009, 113, 12551 CAS.
  74. S. Muhammad, C. Liu, L. Zhao, S. Wu and Z. Su, Theor. Chem. Acc., 2009, 122, 77 CrossRef CAS.
  75. S. Muhammad, H. Xu, Y. Liao, Y. Kan and Z. Su, J. Am. Chem. Soc., 2009, 131, 11833 CrossRef CAS PubMed.
  76. H.-L. Xu, Z.-R. Li, Z.-M. Su, S. Muhammad, F. L. Gu and K. Harigaya, J. Phys. Chem. C, 2009, 113, 15380 CAS.
  77. S. Muhammad, H. Xu, M. R. S. A. Janjua, Z. Su and M. Nadeem, Phys. Chem. Chem. Phys., 2010, 12, 4791 RSC.
  78. S. Muhammad, T. Minami, H. Fukui, K. Yoneda, R. Kishi, Y. Shigeta and M. Nakano, J. Phys. Chem. A, 2012, 116, 1417 CrossRef CAS PubMed.
  79. D. L. Wang, H. L. Xu, Z. M. Su, S. Muhammad and D. Y. Hou, ChemPhysChem, 2012, 13, 1232 CrossRef CAS PubMed.
  80. R.-L. Zhong, H.-L. Xu, S. Muhammad, J. Zhang and Z.-M. Su, J. Mater. Chem., 2012, 22, 2196 RSC.
  81. S. Muhammad, H. Xu, Z. Su, K. Fukuda, R. Kishi, Y. Shigeta and M. Nakano, Dalton Trans., 2013, 15053 RSC.
  82. A. B. Ahmed, H. Feki, Y. Abid, H. Boughzala and A. Mlayah, J. Mol. Struct., 2008, 888, 180 CrossRef PubMed.
  83. Y. Xiaodi, L. Qikai and S. Zhigang, Nanotechnology, 2007, 18, 424029 CrossRef PubMed.
  84. A. R. Chaudhry, R. Ahmed, A. Irfan, A. Shaari, H. Maarof and A. G. Al-Sehemi, Sains Malays., 2014, 43, 867 Search PubMed.
  85. P. J. Hay, J. Phys. Chem. A, 2002, 106, 1634 CrossRef CAS.
  86. A. Curioni, W. Andreoni, R. Treusch, F. J. Himpsel, E. Haskal, P. Seidler, C. Heske, S. Kakar, T. Van Buuren and L. J. Terminello, Appl. Phys. Lett., 1998, 72, 1575 CrossRef CAS PubMed.
  87. S. Y. Hong, D. Y. Kim, C. Y. Kim and R. Hoffmann, Macromolecules, 2001, 34, 6474 CrossRef CAS.
  88. V. Arjunan, P. S. Balamourougane, C. V. Mythili and S. Mohan, J. Mol. Struct., 2011, 1003, 92 CrossRef CAS PubMed.
  89. Y. Zhang, X. Cai, Y. Bian, X. Li and J. Jiang, J. Phys. Chem. C, 2008, 112, 5148 CAS.
  90. R. A. Marcus, Rev. Mod. Phys., 1993, 65, 599 CrossRef CAS.
  91. J. L. Brédas, J. P. Calbert, D. A. da Silva Filho and J. Cornil, Proc. Natl. Acad. Sci. U. S. A., 2002, 99, 5804 CrossRef PubMed.
  92. Y. Yi, L. Zhu and J.-L. Brédas, J. Phys. Chem. C, 2012, 116, 5215 CAS.
  93. S. Mohakud, A. P. Alex and S. K. Pati, J. Phys. Chem. C, 2010, 114, 20436 CAS.
  94. Y. Ie, Y. Umemoto, M. Okabe, T. Kusunoki, K.-i. Nakayama, Y.-J. Pu, J. Kido, H. Tada and Y. Aso, Org. Lett., 2008, 10, 833 CrossRef CAS PubMed.
  95. A. R. Chaudhry, R. Ahmed, A. Irfan, A. Shaari and A. G. Al-Sehemi, Sci. Adv. Mater., 2014, 6, 1727 CrossRef CAS PubMed.
  96. W. H. Baur and D. Kassner, Acta Crystallogr., Sect. B: Struct. Sci., 1992, 48, 356 CrossRef.
  97. S. L. Mayo, B. D. Olafson and W. A. Goddard, J. Phys. Chem., 1990, 94, 8897 CrossRef CAS.
  98. A. Irfan and J. Zhang, Theor. Chem. Acc., 2009, 124, 339 CrossRef CAS.
  99. A. Irfan, R. Cui, J. Zhang and M. Nadeem, Aust. J. Chem., 2010, 63, 1283 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra05850j

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.