Abstract

Solar energy is receiving considerable attention worldwide. Our contribution here focuses on fabricating p-N,N-(dimethylamino) benzoic acid (4-DMABA) donor-π-acceptor derivatives for use in dye-sensitized solar cells (DSSCs). The gas-phase and solvated 4-DMABA and some of its electron donating or withdrawing ortho or meta derivatives were studied theoretically. Density functional theory (DFT) and time-dependent DFT (TD-DFT) were applied to visualize their structural, molecular, photoelectrical, electronic, and photophysical parameters. The parameters for monitoring DSSC efficacies include HOMOs, LUMOs, energy gaps, wavelengths, oscillator strengths, light harvesting efficiencies (LHE), electron injection driving forces (ΔGinject), regeneration driving forces (ΔGregen), open circuit voltages (VOC), and short-circuit current densities (Jsc).

1. Introduction

The rapid development and utilization of renewable energy are vital measures to provide clean energy sources. The steadily increasing global demand for energy, coupled with the depletion of harmful environmental effects, which are associated with the combustion of fossil fuels, has necessitated the rapid search for renewable energy sources [1]. At the present time, the global energy sources are mainly dependent on fossil fuels. This energy source is the main cause of global warming that brings about devastating climate changes as a result of the emitted carbon dioxide (CO2) [2]. However, the alarming rate of depletion of the major conventional energy resources such as coal, petroleum, and natural gas, coupled with the environmental degradation caused by the process of harnessing these energy sources, has necessitated the investment in renewable energy resources that could provide sufficient power without degrading the environment through greenhouse gas emissions. These renewable energy sources include wind power, solar energy, hydropower, bioenergy, and geothermal energy.

Dye-sensitized solar cells (DSSCs) have received considerable attention due to their low-cost conversion of photovoltaic energy compared to the silicon-based semiconductor ones [3]. In particular, DSSCs are characterized by acquiring a wide bandgap semiconductor, typically titanium dioxide (TiO2) that can be sensitized by molecular organic dyes. This architectural design, together with a transparent conductive oxide layer (TCO) and a redox electrolyte, typically iodide/triiodide, is capable of capturing light in the visible region of the spectrum [4].

Solar energy is considered the most promising renewable energy source for supplying the future with energy [5]. Consequently, dye-sensitized solar cells (DSSCs) have attracted ever-increasing attention in scientific research and in practical applications since the first report by O’Regan and Grätzel in 1991 [6] due to their low cost and high efficiency in converting sunlight into electricity. In DSSCs, the incoming light photons excite the dye electrons, which are injected into the conduction band of the nanocrystalline metal oxide. This step is followed by the regaining of electrons by the dye from the redox couple in the electrolyte solution [7]. Generally, a power-conversion efficient dye sensitizer has the following characteristics: The highest occupied molecular orbital (HOMO) energy must be below the conduction band minimum (CBM) of the semiconductor (TiO2) and the lowest unoccupied molecular orbital (LUMO) energy should be higher than the energy of the redox electrolyte pair (I/I−3).

The sensitizing dyes play important roles in DSSCs through maximizing the solar-to-electricity conversion efficiency [8]. In contrast to experimental results on metal-free organic dyes, the theoretical investigations are still lagging behind. Only a few research groups have studied the electronic structures and photophysical properties of the dye sensitizers [9] and the intramolecular electron dynamic process between dyes and TiO2 nanocrystals [10, 11]. Wang et al. [11] have thoroughly studied theoretically 3-(10-butyl-8-(methylthio)-10H-phenothiazin-3-yl)-2-cyanoacrylic acid coupled with diketopyrrolopyrrole derivatives. In this context, the theoretical study of the structures of several benzoic acid derivatives has been conducted recently with considerable accuracy [12]. The structures of p-dimethylamino benzonitrile (pDMABN) [13], p-aminobenzoic acid (pANA) [14],and Ethyl-4-(dimethylamino) benzoate (4-EDMAB) were taken as references to study the effect of para-substituents on the aromatic-ring moiety. The structure of p-(dimethylamino) benzoic acid (hereinafter-DMABA), with the dimethylamino group as a para-substituent having relatively intermediate behavior as compared with the nitro and amino groups, is expected to yield more information about the cooperative electronic interaction in the para-substituted compounds.

In this study, we endeavor to monitor the characteristics of several donor-π-acceptor (D-π-A) dyes having –N(Me)2 and –COOH moieties as donors and acceptors, respectively. The electronic structure and optical absorption properties of eight dye sensitizers (see Scheme 1) in the gas-phase and dimethyl sulphoxide (DMSO) were investigated by using density functional theory (DFT) and time-dependent density functional theory (TD-DFT). Based on the calculated results, we analyzed the role of the different electron-donor or–acceptor groups in tuning the geometries, electronic structures, and optical properties. In addition, we aim to see the effects of the donors or the acceptors of the sensitizers on the open circuit photovoltage (Voc) and the short-circuit current density (Jsc) of the cell by discussing the key factors affecting Voc and Jsc with the goal of finding potential sensitizers for use in solar cells.

2. Computational Details

The ab initio molecular orbital calculations were performed using the Gaussian09 suites [15] and viewed by GaussView software [16]. All calculations for structural optimization were done by density functional theory (DFT) [17] using the B3LYP (Beck three parameter Lee-Yang-Parr) [17], the Coulomb-attenuating Beck three parameter Lee-Yang-Parr (CAM-B3LYP) [18], the pure functional, Perdew, Burke, and Ernzerhof (PBEPBE) [19], and the long-range corrected hybrid ωB97XD [20] functional. All the geometrical structures of the substrates yielded global minima in their potential energy surfaces. This is achieved by performing vibrational calculations that gave no imaginary frequencies. The time-dependent density functional theory (TD-DFT) [21] was used to simulate the absorption and emission spectra of the substrates which were monitored by the Chemissian Software [22]. The effects of basis sets were investigated by using the triple-zeta without and with polarization and diffuse functions on hydrogen and heavy atom basis sets: 6-311G, 6–311++G, 6–311G, 6–311+ G, and 6–311++G [23]. The HOMO and LUMO, electron absorption and emission spectra, oscillator strengths (f), and light harvesting efficiency of the optimized substrates were investigated using the same methods and basis set in the gas phase. The dimethyl sulphoxide (DMSO) solvated substrates were studied using the polarizable continuum model (PCM) [24] for simulating their absorption and emission spectra.

3. Result and Discussion

3.1. The Geometry

Table 1 lists some selected bond lengths, angles, and dihedral angles of the neat 4-DMABA and Ti+4-bound complex computed by using PBEPBE, B3LYP, CAM-B3LYP, and WB97XD functionals with 6–311++G basis sets in the gas phase. The atom numbering of these two substrates are shown in Figure 1, while their standard orientations are depicted in Tables S2 and S3. A quick look at Table 1 shows that the neat 4-DMABA estimated geometrical parameters using all DFT functionals are in excellent agreement with each other. The calculated exocyclic C1–N15 and C4–C11 bond lengths of 4-DMABA of ca. 1.377 and 1.471 Å, respectively, are in excellent agreement with those of its solid-state crystal structure; while the C11–O12 and C11–O13 bonds of the former are longer (ca. 0.04 Å) and shorter (0.09 Å), respectively, compared to those of the latter as a result of dimerization in the solid-state [25]. All DFT functionals have predicted completely planar gas-phase geometries, but the crystal structures have indicated deviations from a planarity of ca. 2–4° [25]. The exocyclic C1–N15 bond length is shorter than a typical C–N single bond by 0.094 Å [26] and longer than an exemplary C=N double bond by 0.105 Å [27]. Likewise, the exocyclic C4–C11 bond length is shorter than a model C–C bond length by 0.068 Å [28] and longer by 0.134 Å compared to a representative C=C double bond [29]. The multiple bond characters of these bonds are quite important in dictating the passage of charge from donors (N–N-dimethyl amine group) to acceptors (carboxyl moiety).

In comparison, the calculated gas-phase geometrical parameters of the Ti+4-bound complex using all DFT functionals are in complete coincidence with each other and agree satisfactorily with those of a DMABA-Zn bond complex [30] as shown in Table 1. Noticeably, the planarity of the neat dye is hardly lifted upon chelation with the titanium hydroxide. Here, again, the exocyclic C1–N15 and C4–C11 bonds have witnessed a slight shortening as a result of the bidentate binding of 4-DMABA with Ti(OH)4. In addition, the parent dye O12–C11–O13 bond angle is closed by ca. 5° to facilitate the bidentate complexation. These observations could be harnessed as indicators for the existence of more charge transfer from the donor to the acceptor. As the charge transfer is a cornerstone in the functioning of solar cells; this parameter will be investigated more thoroughly in the upcoming sections.

3.2. Frontier Molecular Orbitals (FMO)

The Frontier molecular orbitals (FMOs) include some of the lowest unoccupied molecular orbitals (LUMOs) and a few of the highest occupied molecular orbitals (HOMOs). The HOMO-LUMO energy gap is an ideal quantum mechanical descriptor that plays a major role in governing a wide range of chemical interactions [31]. In Figure 2 are depicted the FMOs of the parent dye and its Ti+4-bound complex were computed by using the B3LYP/6311++G level of theory. As shown in Table 2, all the FMOs of the parent dye are stabilized upon chelation with Ti(OH)4. The LUMO+2 for the neat dye is a π-antibonding orbital spreading all over the dye skeleton, while that of the Ti+4-bound complex is crowding around the Ti-atom. The LUMO+1 is 0.07 and 0.24 eV below the LUMO+2 of the parent dye and its Ti+4-boud complex, respectively. Both of them are benzene ring π-antibonding orbitals. The LUMOs are concentrated around the carboxylate chelating acceptor moiety as π-antibonding orbitals. The LUMO of the neat dye is stabilized by 0.74 eV upon chelation. The HOMOs of both species are π-bonding orbitals engulfing the donor dimethylamino group. The complexation effect has decreased the energy gap of the neat dye by 0.6 to 0.7 eV, which means more charge flow [32]. The HOMO-1 and HOMO-2 of the neat dye are oxygen atom’s nonbonding orbitals and benzene π-bonding orbitals, respectively. Upon chelation, both of them characterize the π-bonding orbitals on the benzene moiety. They are 1.428 and 1.800 eV below the HOMO of the parent species, respectively: and 1.409 and 2.215 eV underneath the HOMO of the Ti+4-bound complex, respectively.

As a conclusion, all the as-investigated functionals have yielded narrower energy gaps and larger dipole moments upon chelation, in excellent agreement with that of Wang et al. [11]. This means that an easy charge flow occurs when the dye preys on the Ti (OH)4 molecule. It is apparent that the pure PBEPBE functional has given the straight energy gap and larger dipole moment; while the dispersion ωB97XD functional produced the colossal energy gap and smaller dipole moments. The order of decreasing energy gaps among the applied functionals is as follows: ωB97XD > CAM-B3LYP > B3LYP > PBEPBE [33]. This trend is inverted by the calculated dipole moments. The pure PBEPBE functional yielded the narrowest energy gap by stabilizing the LUMO and destabilizing the HOMO; while the ωB97XD functional gave the widest energy gaps through destabilizing the LUMO and stabilizing the HOMO [34]. As the hybrid B3LYP functional gave energy gap suitable for reasonable charge flow; it was adopted for simulating the absorption and emission spectra and the theoretical testing of DSSCs [35].

3.3. Electronic Excitations of 4-DMABA Derivatives

The ππ and nπ electronic transitions in π-conjugated organic compounds are recognized for leading to UV-visible spectra [36, 37]. Table 2 lists the maximum absorption wavelengths of gas-phase 4-DMABA dye (D1) using the as-investigated functionals with 6–311++G basis set. The parent (D1) absorption UV-visible bands calculated by B3LYP functional are in excellent agreement with its experimental ones [38]. In Table S1 (Supplementary Information), the electronic transitions of D1 are registered using B3LYP and PBEPBE functionals with different basis sets. The polarization functions have a propensity to underestimate the maximum absorption wavelengths [39], while the diffuse functions are liable to overestimating them [40]. That is, when the polarization [6-311G] and the diffuse [6–311++G] basis sets were applied with B3LYP and PBEPBE functionals, the maximum absorption wavelengths were blue-shifted and red-shifted, respectively, compared with the values obtained by using the 6-311G basis set with these two functionals (See Table S1). Hence, it is recommended to use both the polarization and diffuse functions in simulating UV-Vis spectra [41].

Normally, molecules with large dipole moments have strong asymmetric electronic charge distribution and thus become more reactive and sensitive to external electric fields [42]. The theoretical UV-visible spectra of 4-DMABA using TD-B3LYP and TD-PBEPBE functionals with a 6–311++G basis set of theory are shown in Figures S1 and S2. TD-B3LYP/6-311++G predicted intense electronic transition bands at λmax = 290 nm (f = 0.5297) which is red-shifted by ca. 11 nm compared to the experimental value of 310 nm [11, 38]. The calculations showed that these bands originate mainly from a HOMO- > LUMO transition (See Table S1). Moreover, the computed dipole moment using this level of theory of 5.164 Debye is in line with the values of the predicted electronic bands (See Table 2).

We have also investigated some D1 derivatives (D1.1-D1.7) by replacing the hydrogen atoms at the ortho or meta positions by two or four CH3, OH, NH2, F, CN, and NO2 as electron donating or withdrawing moieties (cf. Scheme 1). Figure 3 depicts their optimized geometries using the B3LYP/6-311++G level of theory. Their standard orientations are registered in Tables S4S10. Figure 4(a) depicts the absorption spectra of the DMSO-solvated D1-D1.7 dyes which have been simulated by using TD-B3LYP/6-311++G level of theory.

Table 3 lists the calculated HOMO, LUMO, bandgaps, absorption wavelengths, oscillator strengths, light harvesting efficiencies (LHE), and electronic transition assignments of the gas-phase D1-D1.7 derivatives using the B3LYP/6-311++G level of theory. On the one hand, the electron donating groups (CH3, NH2, and OH) at the ortho positions have widened the band gaps and hence blueshifted the wavelengths compared to those of the parent dye (D1). On the other hand, the electron withdrawing groups (F, CN, and NO2) at the meta positions have straitened the bandgaps and accordingly redshifted the wavelengths [11]. When both electron donating (CH3) and withdrawing (CN) ligands are used (D1.7); moderate results are obtained for both the band gap and absorption wavelength. These UV spectra originate mostly from HOMO to LUMO transitions.

In Table 4 are registered the HOMO, LUMO, bandgaps (∆E), absorption and emission wavelengths, the emission oscillator strengths of D1.1–D1.7 solvated (DMSO (ε = 46)) derivatives which were obtained by using TD-B3LYP functional with 6–311++G basis set. Apart from D1.4, all the energy gaps of the elected dyes (D1.1-D1.7) are smaller than that of the parent dye(D1). The derivatives D1.6, D1.7 and D1.5 gave the smallest energy gaps amongst them. The absorption wavelength (λabs/nm) of D1.1 to D1.7 derivatives are slightly shifted compared to that of D1. The absorption wavelength of the derivatives having–I (inductive) with +R (resonance) effects (NH2, OH, and F) were generally blueshifted compared to that of the parent D1; while those for the ones with–I effect (CN) were redshifted. In the methyl derivative (D1.1), the methyl groups with +I effect have redshifted the peaks. Similar trends have been shown by the emission spectra.

Furthermore, the emission spectra of the DMSO-solvated dyes D1.1–D1.7 are depicted in Figure 4(b), while the corresponding values for the orbitals responsible for the first excited singlet state (HOMO-LUMO contribution), excitation energy (cm−1), maximum emission wavelength (λem), emission oscillator strength (f), excited-state lifetime (τ), and Stokes shift (SS) are presented in Table 4. The emission spectra arising from the S1–S0 transitions were assigned ππ character for all the studied dyes. Similar to the absorption spectrum, the emission bands were found to be related to the HOMO-to-LUMO transitions. It is noteworthy that the small values of the Stokes shift (SS) were obtained for all dyes. It is known that dyes with weak Stokes shifts present minimal conformational reorganization between the ground and excited state [43]. Our results showed that the spectral properties of the derivatives D1-D1.7 are weakly affected by these structural changes.

3.4. The Life Time of the Singlet Excited State (τ)

The first excited state (S1) to the ground--state (S0) decay step is extremely important for injecting the excited electron into the conduction band minimum (CBM) of the TiO2 semiconductor. Thus, the lifetime (τ) of the excited state is an extremely crucial parameter for evaluating the charge transfer efficiency [44]. A dye sensitizer having an excited state with a longer τ value is expected to facilitate the charge transfer [45]. In Table 4 are listed the decay time (τ/nm) of the derivatives. They are obtained using where is the emission energy in cm−1 and f is the oscillator strength of the transition state [46]. As Table 4 shows, apart from D1.3 and D1.4, the τ values for all the derivatives are much longer than that for the parent D1. They increase in the order D1<D1.2<D1.1<D1.5D1.7<D1.6. Therefore, generally the introduction of the electron withdrawing, apart from F atoms, or donating groups into DMABA, increases the lifetime of the excited state, expedites the energy transfer, and boosts the efficacy of DSSCs. Accordingly, the derivatives D1.6, D1.7, and D1.5, acquiring NO2, CH3 with CN, and CN groups, respectively, display substantially dynamic charge transfer and electron admission to the conduction band minimum of the TiO2 semiconductor substrate. In addition, these three dyes could hinder the charge recombination process and boost the performance of DSSCs in comparison to D1. It is worth noting that the excited states of derivatives, in the more polar solvent (DMSO), live longer compared to the less polar ones [47].

3.5. The Overall Performance

The power conversion efficiency (PCE) of DSSCs is investigated by the short-circuit current density (JSC), open-circuit photovoltage (VOC), and fill factor (FF) values, as well as the intensity of the incident light (PIN). It is estimated by applying Equation (2) [48]:where FF is the ratio of the maximum power of the DSSC and the product of VOC, JSC, and PIN [49]. As (2) shows, the impact of the electron withdrawing or donating groups on the product of VOC and JSC should be optimized to enhance the performance. VOC can be computed by applying (3) [50],where kBT is the thermal energy, q is the unit charge, and nC is the number of electrons in the conduction band, NCB is the conduction band attainable density of states, and Eredox is the electrolyte oxidation potential. VOC is computed as an energy difference between the conduction band minimum of the TiO2 semiconductor and the iodide/triiodide () electrolyte redox potential, which is assumed to be constant.

It is known that the value of JSC in a DSSC can be computed by applying (4) [51]where LHE denotes the light-harvesting efficiency with respect to a given wavelength (λmax), Φinject gives the electron insertion efficacy, and ηcollect is related to the charge collection performance. In our proposed DSSCs, the electrode () is kept constant while the dye sensitizers are varied. Accordingly, ηcollect turns out to be a constant.

However, LHE(λ) can be estimated by (5) [52]:where f denotes the oscillator strength of the absorption wavelength (λmax). It is known that the LHE and the free energy for electron injection (∆Ginject) are the two important elements that impact JSC. The values of LHE of the elected DMABA derivatives should be high to maximize the performance of the DSSCs. Higher oscillator strengths enhance the light harvesting efficiencies and maximize the overlap with the solar light, especially over the whole UV–visible spectral region. The light harvesting efficiencies (LHE) using TD-B3LYP/6-311++G listed in Table 4 of the different DMSO-solvated DMABA derivatives (D1.1-D1.7) range between 0.168 and 0.795, indicating that these derivatives provide different photocurrent performances. JSC can also be increased through enhancing the electron injection free energy () which can be computed by Equation (6) [53]where denotes the reduction potential energy level of the TiO2 conduction band and gives the oxidation potential energy of the excited state DMABA derivatives, and which can be estimated by (7) [5456]:where denotes the ground-state energy of the DMABA derivatives in the oxidation potential energy, while E is the lowest vertical electronic excitation energy, related to λmax.

As registered in Table 5, all the elected DMABA derivatives give negative ∆Ginject values, where D1.2 gives the most negative value; while D1.7 shows the least negative one. In addition, the electron donating groups CH3, OH, and NH2 represented by D1.1, D1,2, and D1.3, respectively, yield more negative values compared to that of the parent D1; while the electron-withdrawing moieties F, CN, NO2, and CN with CH3 represented by D1.4, D1.5, D1.6, and D1.17, respectively, give less negative ones relative to that of their predecessor D1 (See Table 5). The absolute values of ∆Ginject for all DMABA derivatives understudy are much greater than the 0.2 eV threshold necessary for an efficient electron injection [57]. It is noteworthy that the large energy difference between the LUMOs of the DMABA derivatives and the CBM of TiO2 of −4.0 eV is large enough to secure efficient electron injections (See Figure 5).

J SC of the DSSC is affected by the regeneration efficiency (ηreg) of the DMABA derivatives, which can be computed from the regeneration force represented by ∆Greg shown by (8) [58],

Table 5 lists the computed values of for the derivatives D1.1–D1.7 of 0.829, 0.717, 1.085, 1.289, 1.542, 1.514, and 1.286 eV, respectively. It is clear that D1.1 and D1.2 offer relatively smaller driving forces for dye regeneration, which may enhance ηreg compared with the parent D1 (0.965 eV). In particular, derivative D1.2 could present the fastest regeneration, promoting the DSSC efficacy. In addition, VOC is computed as an energy difference between ELUMO of the DMABA derivative and ECB of the TiO2 semiconductor [59]:

Our theoretical protocol has yielded VOC values for the DMSO-solvated derivatives that spread between 0.699 and 2.931 eV (see Table 5). These values are quite appropriate for enabling efficacious electron injection into the LUMO of the carboxylate moiety as an electron acceptor. Thus, all the elected derivatives could be applied as sensitizers. This is because the carboxylate (COOH) acceptor group is quite propitious for electron injection from the dye into the CBM of the TiO2 semiconductor, a situation that results in improved VOC values. The derivative D1.2 characterized by the NH2 moiety as an electron donating group, showed the highest VOC value and hence could acquire an outstanding potential for application in DSSCs.

3.6. The Binding Energy (Eb)

For a high power conversion efficiency (PCE), the electron-hole entities must be kept separate as positive and negative charges to avoid recombination as a result of the attractive Coulombic forces. To achieve this crucial process, the binding energy (Eb) should be overcome, i.e., the derivatives should have lower binding energy to yield high PCE values. The binding energy (Eb) could be computed by (10) [60, 61],where is the HOMO–LUMO energy gap and Ex is the optical energy, calculated with reference to λmax. Table 5 lists the exciton computed binding energies for the parent D1 and its derivatives D1.1-D1.7. It reveals the lowest value for D1.6 compared with the remaining derivatives. These findings reconfirm that the derivative D1.6 characterized by the NO2 group is the most preferable for fabricating DSSCs.

4. Conclusion

We attempted to explore the optimized geometry, electronic structure, and their related optical absorption and emission properties of eight D–π–A-type chromophores with different ortho or para electron withdrawing or donating moieties. Our findings demonstrate that the ortho electron donating NH2, the meta electron withdrawing CN, and the ortho electron donating CH3 with meta electron withdrawing CN groups applied in D1.2, D1.5, and D1.7, respectively, are favorable for implication in the D–π–A design. The coplanar environment of the dimethylamine donor, benzene ring, and carboxylate acceptor together with the ortho or meta groups boosts efficacious injection of electrons from the donor [(CH3)2N-] to the acceptor [-COOH] of the elected dyes. Our findings show that the electron- withdrawing or donating moieties implicated in D1.2, D1.5 and D1.7 have generated relatively strong absorptions for maintaining stable charge transfer pathways for prompt electron grouting and dye regeneration, first singlet excited-state lifetime, and exciton binding energy. In addition, the elected dyes D1.5, D1.6, and D1.7 show small HOMO-LUMO energy gaps and clear red shifts in absorption and emission, compared with the parent dye (D1). It could be concluded that the elected dyes D1.2, D1.5, D1.6, and D1.7 are more favorable nominees for applications in relatively potent organic DSSCs.

Data Availability

The authors confirm that the data supporting the findings of this study are available within the article and the accompanying materials.

Conflicts of Interest

The authors declare that they have no conflicts of interest.

Authors’ Contributions

O.I.O. planned the research. S.A.E. and S.G.A. provided the necessary literature, and M.A.H. performed some calculations and proofread the manuscript. M.Y.A reset of the calculation, analyzed the results, and wrote the first version of the manuscript.

Acknowledgments

This work was supported by the Deanship of Scientific Research (DSR) King Abdulaziz University, Jeddah, under grant no. (G:84–130-1441). The authors, therefore, gratefully acknowledge with thanks the DSR’s technical and financial support. The authors are highly indebted to the High-Performance Computing (HPC) Center at King Abdulaziz University for their technical support.

Supplementary Materials

Supplemental Table 1 shows electronic absorption spectra of gas-phase free p-N,N-(Dimethylamino) benzoic acid. Supplemental Table 2 lists the total energy, number of imaginary frequencies, and Standard Orientation of 4-DMABA. Supplemental Table 3 lists the total energy, number of imaginary frequencies, and Standard Orientation of Ti+4-DMABA. Supplemental Tables 4–10 list the total energies, number of imaginary frequencies, and Standard Orientations of D1.1- D1.7 substrates. Supplemental Figure 1 shows the electronic absorption spectrum of the free 4-(dimethylamino) benzoic acid using the B3LYB/6–311++ G level of theory. Supplemental Figure 2 shows the electronic absorption spectrum of the free 4-(dimethylamino) benzoic acid using the PBEPBE/6–311++ G level of theory. (Supplementary Materials)