Comparison of selected dihedral angles (
porphyrin (TSPP) with their protonated structures (H4FBP, H4TPP, H4TSPP and H8TSPP), calculated in water as solvent at the B3LYP/6-311G(d,p) level of DFT.
Abstract
In this chapter, we discuss protonation and substitution effects on the absorption spectra of porphyrin molecules based on density functional theory (DFT) and time-dependent DFT calculations. The results of the calculations are compared with experimental data. The calculations show that protonation of core nitrogen atoms of porphyrin and meso-substituted porphyrins produces a substantial shift in Soret and Q-absorption bands, relative to their positions in corresponding nonprotonated and nonsubstituted chromophores. A relaxed potential energy surface (RPES) scan has been utilized to calculate ground and excited state potential energy surface (PES) curves as functions of the rotation of one of the meso-substituted sulfonatophenyl groups about dihedral angles θ (corresponding to Cα─Cm─Cϕ─C) ranging from 40 to 130°, using 10° increments. The ground state RPES curve indicates that when the molecule transitions from the lowest ground state to a local state, the calculated highest potential energy barrier at the dihedral angle of 90° is only 177 cm−1. This finding suggests that the meso-sulfonatophenyl substitution groups are able to rotate around Cm─Cϕ bond at room temperature because the thermal energy (kBT) at 298 K is 207.2 cm−1. Furthermore, the calculations show that the geometric structure of the porphyrin is strongly dependent on protonation and the nature of the meso-substituted functional groups.
Keywords
- porphyrins
- protonation
- absorption
- PES
- DFT calculation
1. Introduction: overview of molecular spectroscopy and quantum calculations
Spectroscopy is the branch of science dealing with the interaction of electromagnetic and other forms of radiated energy with matter. The earliest prospect of making spectroscopic measurements came with the observation that visible light can be dispersed by an optical prism, and the concomitant recognition that matter could be intimately investigated through its response to optical radiative energy as a function of frequency, defining what is referred to as optical spectroscopy. As it turns out, optical spectroscopy is a useful approach for both qualitative and quantitative studies of physical and chemical processes involving matter in most of its states by measurement of absorption, emission, or scattering of electromagnetic radiation; moreover, optical spectroscopic measurements can be very sensitive, nondestructive, and typically require only small amounts of material for analysis.
Absorption spectra are usually acquired for analytes dissolved in nonabsorbing solvents. And, ideally the absorbance of a dissolved analyte depends linearly on concentration, thereby resulting in an absorption spectrum providing quantitative measurement of the analyte's concentration in solution, arrived at by applying the Beer-Lambert Law. In particular, since absorption spectra of molecules depend on their energy level structure, absorption spectra are not only useful for identifying isolated molecules, but also can be used to probe intermolecular interactions (e.g., effects of aggregation) that affect energy level structure.
It is to be noted that molecules that are excited to higher energy than the lowest excited state above the ground electronic state can relax to lower excited levels by a range of intrinsic processes. Included among such deactivation processes are emission of radiation, more popularly referred to as luminescence, as well as processes that are nonradiative in nature, where lower energy states can be directly populated without the emission of photons. Luminescence from such intermediate states can be defined as fluorescence or phosphorescence, where fluorescence is a process by which electronically excited molecules return to a lower electronic state of the same spin multiplicity (which is often the electronic ground state) by emitting a photon; phosphorescence, on the other hand, is the corresponding transition between states with different spin multiplicities. While fluorescence is a spin-allowed process and generally occurs rapidly, phosphorescence is spin forbidden and is typically a slower relaxation process.
Paths by which nonradiative relaxation can occur include, but are not limited to, such phenomena as collisional energy transfer, electron or proton transfer processes, change of molecular conformation, photochemistry, formation of excited state complexes (e.g., excimers or exciplexes), as well as the classic processes of internal conversion (IC) (e.g., vibrational relaxation) and intersystem crossing (ISC) (e.g., singlet-triplet conversion).
It is to be noted that transient intermediates are likely to form during IC and ISC radiationless processes, and detection of such species, if at all possible, often necessitates the use of highly sensitive ultrafast optical (or other) techniques.
The aforementioned phenomena are depicted more fully in Figure 1 that shows a combined Perrin‐Jablonski diagram illustrating the different processes involved in the interaction of a molecule with photons in the spectral region between 300 and 1500 nm. Photophysical processes for an isolated molecule would occur via transitions between the different internal energy states shown in Figure 1.

Figure 1.
A general Perrin-Jablonski diagram, where S and T stand for singlet and triplet electronic states, respectively. IC and ISC represent “internal conversion” and “intersystem crossing,” respectively.
Using Figure 1 for discussion, in the gaseous or solution phase at room temperature a molecular system is generally in its ground state (S0). The transition from the ground state to an upper vibroelectronic state by absorption of a photon would take place within ca. 10−15 s, which is much faster than the emission of the photon from an excited electronic state (
During the past two decades, there has been very intense theoretical research on the physical and chemical properties of molecular structures. Computational chemistry is a powerful tool for investigation of molecules, surfaces, and interfaces at the electronic structure level. Various molecular properties directly comparable with experiment such as structural parameters, thermodynamic data, and vibrational spectra can be obtained by solving quantum mechanical equations. When the result of a theoretical prediction is consistent with an experimental measurement, one can more confidently interpret the experimental result. Computational studies are not only carried out in order to provide an understanding of experimental data, such as the position and source of spectroscopic peaks, but also can be used to predict the existence of unobserved molecules, intermediates, or to explore reaction mechanisms that are not readily studied experimentally.
The particular computational method chosen depends critically on the desired accuracy (qualitative vs. quantitative) sought, the size of the system, and available computational capacities. At a qualitative level, especially for large systems, molecules can be treated by classical mechanics using a class of methods called molecular mechanics. The structure of a protein containing hundreds of atoms might be calculated this way. Somewhat more quantitatively accurate are the semiempirical methods. These methods (such as PM3) use experimentally measured parameters that approximate parts of a quantum mechanical system. These latter methods can be fast, and give good results if the molecule of interest is very similar to those used to determine the parameters. However, many molecules of interest (i.e., transition metal complexes) do not have sufficiently good parameter sets to be accurately calculated using such methods.
The fourth type of computational methods are the density functional theory (DFT) methods. With a few exceptions, DFT is the most cost-effective method to achieve a given level of quantitative accuracy. It incorporates electron correlation and is computationally less expense.
In addition to choosing a method, one must also choose a basis set. A basis set is a set of functions that substitute for the “real” atomic orbitals (AOs) of a system and should approximate the real wave functions well enough to give chemically meaningful and close approximations to the correct values of measurable quantities being considered (e.g., geometry and energy). Using more complex basis sets improves results at the cost of increased computer time to make a calculation (i.e., increased computational expense). Basis sets, in order to allow electron-electron correlation to be taken into account, must incorporate polarization terms to allow distortions in orbital shapes; they must also incorporate diffuse functions (especially necessary when you have a molecule with weakly bound electrons) (as in the case of some anions and for some transition states); and they must account for relativistic effects (for heavier atoms).
In this chapter, we will discuss protonation and
2. Porphyrin macrocycle
Porphyrin and its derivatives have received extensive attention from both experimentalists and theoreticians since they have been found to have many important potential applications in a broad variety of high technology and biomedical fields. Indeed, in recent years analyses of geometric and spectroscopic properties of molecular systems incorporating porphyrins have produced a substantial body of information that has greatly expanded our knowledge of high efficiency utilization of solar energy [1–5] and the use of such synthetic molecular analogs as active agents in molecular electronic devices [6, 7]. Also, a great deal of interest has been shown for the use of porphyrin-like molecular systems as therapeutic drugs and photosensitizers in photodynamic therapy of cancer [8] and their possible use in the treatment of nonmalignant conditions such as psoriasis, treatment of blocked arteries, and for the treatment of pathological and bacterial viruses [9] and HIV [10]. The biological importance of porphyrins essentially derives from their physicochemical properties that basically determine their photophysical behavior. Additionally, aggregation and axial ligation lead to significant changes in absorption spectra as well as quantum yield, fluorescence lifetime, and triplet state lifetime [11–13]. More detailed information about porphyrins can be obtained in the
Of particular note is the observation that optical properties of porphyrin can be altered by the protonation or metallation of nitrogen atoms in its core structure, with electronic changes as a result of structural alterations such as flattening and distortion from planarity of the macrocycle, interactions between porphyrins (aggregation), redox reactions, and solvent effects. A few porphyrins have been found to form aggregates; a requirement of being zwitterionic character upon protonation of macrocycle core nitrogen atoms. It has also been suggested that aggregation is facilitated by interaction with proteins [16, 17] and surfactants [18].
Indeed, aggregation of the anionic porphyrin
The developments in computing facilities and the sophisticated computation programs, with increasingly efficient algorithms, especially the fundamental improvements in the treatment of electron correlation based on density-functional theory (DFT) [22], have combined to allow quantum chemical methods to routinely handle molecular systems containing hundreds of atoms. As a result, DFT has become one of the most important techniques used by theoreticians to provide deep insight into spectroscopic and structural properties, even for complex molecular systems, especially those of large sizes such as the porphyrinoids [23–28].
In this chapter, we discuss the effect of
3. Structures of porphyrin and its derivatives

Figure 2.
Optimized geometric structures of unsubstituted porphyrin (FBP),
DFT theory at the B3LYP/6-311G(d,p) level was performed to predict the geometric parameters of the ground state of the parent porphyrin and its derivatives in water used as a solvent. The optimized ground state geometry of these compounds is provided in Figure 2. The selected bond angles and dihedral angles are given in Table 1. Results of the calculations show that while the
FBP | TPP | TSPP | H4FBP | H4TPP | H4TSPP | H8TSPP | |
---|---|---|---|---|---|---|---|
|
N.A | 3.6 | 4.0 | N.A | 19.4 | 19.6 | 19.0 |
|
180.0 | −176.5 | −176.4 | −169.5 | −166.6 | −160.5 | −161.6 |
|
0.0 | 2.3 | 2.4 | 10.3 | 20.6 | 20.7 | 20.0 |
|
N.A. | 72.1 | 71.0 | N.A | 47.8 | 47.2 | 49.6 |
|
180.0 | −177.2 | −176.9 | −169.5 | −160.6 | −160.5 | −160.8 |
|
180.0 | 179.8 | 179.7 | −178.8 | −176.4 | −174.4 | −176.9 |
|
0.0 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 |
|
180.0 | −179.8 | −179.7 | 178.8 | 176.4 | 176.4 | 176.9 |
|
N.A | −2.6 | −2.6 | N.A | −19.4 | −19.4 | −18.9 |
|
N.A | −72.1 | −70.9 | N.A | −47.8 | −47.5 | −49.9 |
|
0.0 | 0.4 | 0.4 | 2.2 | 4.2 | 4.2 | 4.1 |
|
127.9 | 123.0 | 123.1 | 127.7 | 128.0 | 128.0 | 128.0 |
|
N.A | 118.2 | 118.2 | N.A | 118.3 | 118.3 | 118.2 |
|
N.A | 116.6 | 116.5 | N.A | 118.3 | 118.4 | 118.4 |
|
127.0 | 125.2 | 125.3 | 127.4 | 123.4 | 123.4 | 123.9 |
|
123.4 | 126.9 | 126.9 | 127.7 | 128.0 | 128.0 | 127.9 |
|
125.6 | 126.3 | 126.2 | 125.5 | 125.4 | 125.3 | 125.2 |
|
125.7 | 126.6 | 126.6 | 125.5 | 125.4 | 125.3 | 125.3 |
Table 1.
It is ascertained that the calculated bond lengths are consistent with X-ray data within ca. ± 0.01 Å. Hence, one can conclude that protonation of the porphyrin core, in addition to causing deviation from planarity of the macrocycle, also simply has an effect on the tilt angles of the phenyl and
In Section 3.1, protonation and
3.1. Calculated electronic spectra of porphyrin and its derivatives
Porphyrin and its derivatives find use in a myriad of important natural and biomimetic processes, with the major focus in the latter case on processes such as conversion of solar energy into chemical energy, photodynamic therapy, and as active agents in optical sensors. The excited states of porphyrins play fundamental roles in essentially all processes involving porphyrin and its derivatives.
In this section, we discuss the

Figure 3.
Comparison of calculated dipole-allowed electronic transitions of the parent porphyrin (FBP),
The calculations mainly produce a strong electronic absorption band in the 360–450 nm range and a few weak or very weak electronic transitions below as well as above the strong bands (Figure 3). The strongest band is known as the Soret band (also referred to as the B-band), and weaker bands at longer wavelength, in the range 500–750 nm, are known as Q-bands that are usually quite weak. The results of calculations indicate that (1) the electronic bands in the parent porphyrin (FBP, neutral) are slightly blue-shifted in diprotonated-FBP (H4FBP, dicationic) structure; (2) the bands in neutral TPP molecule (
3.2. The electronic spectra of FBP and protonated-FBP (H4FBP)
The electronic spectrum of the FBP molecule exhibited two weak electronic bands at wavelength longer than that of the Soret band (B): one of the bands corresponds to S0 → S1 (B1u, at 540 nm with oscillator strength
There are also two strong electronic bands in the spectral range of the Soret band (B): one of the bands corresponds to S0 → S4 (B2u, at 367 nm and
A relatively strong band at 330 nm is also calculated as existing. One of these is due to the S0 → S7 transition (B1u, at 330 nm and
The experimental absorption spectrum of FBP [30] exhibits absorption bands at about 372 and 340 nm in the Soret-band region. In the Q-band region, bands at about 512 and 626 nm are observed. The measured bands in the FBP spectrum are in good agreement with calculated values for the B-bands at 380, 367, and 340 nm, and for the Q-bands at 540 and 506 nm, but not for the weak band at 626 nm. In these band regions, the calculation did not produce any dipole-allowed or forbidden singlet-singlet transition. Therefore, the free-base porphin (FBP) sample may contain the free-base aza-porphin(s) (as aza substitution at the
|
|
|
|
|
|
|
|
|
|
||
---|---|---|---|---|---|---|---|---|---|---|---|
1 | 2.30 | 540 | 0.0005 | B1U | H − 1 → L + 1 (40%), H → L (59%) |
1 | 1.51 | 822 | B2U | H − 1 → L (21%), H → L + 1 (79%) |
|
2 | 2.45 | 506 | 0.0003 | B2U | H − 1 → L (47%), H → L + 1 (53%) |
2 | 1.82 | 682 | B1U | H → L (94%) | |
3 | 3.26 | 380 | 0.8144 | B1U | H − 3 → L (22%), H − 1 → L + 1 (48%), H → L (29%) |
3 | 2.04 | 608 | B2U | H − 1 → L (78%), H → L + 1 (22%) |
|
4 | 3.38 | 367 | 1.1911 | B2U | H − 1 → L (50%), H → L + 1 (47%) |
4 | 2.07 | 598 | B1U | H − 1 → L + 1 (94%) | |
5 | 3.44 | 360 | B3G | H − 2 → L (98%) | 7 | 2.90 | 428 | B3G | H − 2 → L (88%) | ||
6 | 3.66 | 339 | AG | H − 2 → L + 1 (99%) | 8 | 2.96 | 419 | B1U | H − 3 → L (86%) | ||
7 | 3.76 | 330 | 0.6934 | B1U | H − 3 → L (76%), H − 1 → L + 1 (12%), H → L (12%) |
9 | 3.15 | 393 | AG | H − 2 → L + 1 (93%) | |
8 | 3.76 | 330 | 0.2479 | B2U | H − 3 → L + 1 (93%) | 11 | 3.33 | 373 | B3G | H − 8 → L + 1 (16%), H → L + 2 (72%) |
|
16 | 4.33 | 287 | 0.0914 | B2U | H − 5 → L + 1 (97%) | 13 | 3.39 | 366 | B2U | H − 3 → L + 1 (96%) | |
18 | 4.41 | 281 | 0.1037 | B1U | H − 5 → L (99%) | 15 | 3.61 | 343 | AG | H − 8 → L (26%), H − 1 → L + 2 (68%) |
|
23 | 5.22 | 237 | 0.1338 | B1U | H − 2 → L + 2 (98%) | 16 | 3.64 | 340 | B3G | H − 4 → L + 1 (79%) | |
1 | 2.31 | 538 | 0.0007 | E | H − 1 → L + 1 (48%), H → L (52%) |
1 | 1.63 | 763 | E | H − 1 → L + 1 (30%), H → L (70%) |
|
2 | 2.31 | 538 | 0.0007 | E | H − 1 → L (48%), H → L + 1 (52%) |
2 | 1.63 | 763 | E | H − 1 → L (30%), H → L + 1 (70%) |
|
3 | 3.39 | 366 | 1.4554 | E | H − 1 → L + 1 (52%), H → L (48%) |
3 | 1.96 | 632 | E | H − 1 → L + 1 (69%), H → L (31%) |
|
4 | 3.39 | 366 | 1.4554 | E | H − 1 → L (52%), H → L + 1 (48%) |
4 | 1.96 | 632 | E | H − 1 → L (69%), H → L + 1 (31%) |
|
7 | 3.90 | 318 | 0.0597 | E | H − 5 → L + 1 (45%), H − 4 → L + 1 (53%) |
7 | 3.23 | 384 | B1 | H − 3 → L + 1 (28%), H − 2 → L (28%), H → L + 2 (31%) |
|
8 | 3.90 | 318 | 0.0597 | E | H − 5 → L (45%), H − 4 → L (53%) |
8 | 3.32 | 374 | E | H − 3 → L + 1 (44%), H − 2 → L (44%) |
|
11 | 4.05 | 306 | 0.0695 | E | H − 5 → L + 1 (54%), H − 4 → L + 1 (45%) |
9 | 3.37 | 367 | E | H − 5 → L + 1 (42%), H − 4 → L + 1 (48%) |
|
12 | 4.05 | 306 | 0.0695 | E | H − 5 → L (54%), H − 4 → L (45%) |
10 | 3.37 | 367 | E | H − 5 → L (42%), H − 4 → L (48%) |
Table 2.
The selected values of the calculated singlet-singlet (
The predicted electronic spectrum of protonated-FBP (H4FBP) exhibits Q- and B-bands for S0 → S1/2 (E at 538 nm with
Comparing the electronic spectrum of FBP with that of H4FBP: in FBP three strong bands were predicted at 380, 360, and 330 nm, see Figure 3 and Table 2, whereas the H4FBP spectrum exhibited only one strong band at 366 nm in the Soret-band region; in the Q-band region, H4FBP has a doubly degenerate band at 538 nm, while FBP has two very weak transitions at 540 and 506 nm. This reduction in the number of bands is due to the higher symmetry for H4FBP.
Also for FBP, the calculated electronic spectrum of diprotonated-FBP (i.e., H4FBP) molecule indicates two IC processes from the S3/4 (the strongest bands or B-band) at 366 nm (with the symmetry E) to the S1/2 (at 538 nm with symmetry E), as well as the ISC process between the S3/4 (at 366 nm) and T7/8 (at 367 nm), see Table 2.

Figure 4.
Plot of calculated electron densities in the desired HOMOs (H) and LUMOs (L) of parent porphyrin (FBP),
The electron density plots of the molecular orbitals (i.e., HOMOs (H) and LUMOs (L)), as seen in Figure 4 and Table 3, show that the H − m and L + m (m = 0, 1, 2, …) are not just pure π and π* molecular orbitals (MOs), in particular cases they also include nonbinding atomic orbitals (AOs).
FBP | H4FBP | ||
---|---|---|---|
H | π(Cβ─Cβ/Cm─Cα) + n(N) | H | π(Cβ─Cβ/Cα─Cm─Cα) + n(N) |
H − 1 | π(Cβ─Cα) | H − 1 | π(Cα─Cβ) |
H − 2 | π(Cα─N─Cα/Cβ─Cβ) | H − 2/H − 3 | π(Cα─Cβ) + n(N) |
H − 3 | π(Cα─N─Cα/Cβ─Cβ) + n(minor, N/Cm) | H − 4/H − 5 | π(Cβ─Cβ) + n(N) |
H − 4/H + 5 | π(Cβ─Cβ) + n(N) | H − 6/H − 7 | π(Cβ─Cα─Cm) + n(minor, N) |
H − 6/H − 7 | n(N) + σ(minor; Cβ─Cα) | H − 8 | π(Cα─Cm─Cα) |
L/L + 1 | π*(Cβ─Cβ/Cβ─Cα/Cm─Cα) + n(minor; N) | L/L + 1 | π(Cβ─Cα) + n(minor, N) |
L + 2 | π*(Cβ─Cα) + n(Cm) | L + 2 | π(Cβ─Cα) + n(Cm) |
L + 3 | π*(Cβ─Cβ/Cm─Cα) + n(N/Cα) | L + 3 | π(Cβ─Cβ) + n(N/Cα/Cm) |
L + 4/L + 5 | π*(Cβ─Cβ/Cm─Cα) + n(N/Cα/Cβ) | L + 4 | π(Cβ─Cβ and Cα─Cm) + n(N/Cα/Cm) |
Table 3.
Bond type of the highest occupied molecular orbitals (H - m) and the lowest unoccupied molecular orbitals (L + m), m = 0, 1, 2, …
3.3. The electronic spectra of TPP and H4TPP
While the calculated spectrum of the TPP molecule displayed two weak peaks at 571 (S0 → S1,
|
|
|||||||||
---|---|---|---|---|---|---|---|---|---|---|
|
|
|
|
|
|
|
|
|
|
|
1 | 2.17 | 571 | 0.0337 | B2 | H − 1 → L + 1 (32%), H → L (67%) |
1 | 1.40 | 884 | B1 | H − 1 → L (16%), H → L + 1 (84%) |
2 | 2.32 | 535 | 0.0359 | B1 | H − 1 → L (37%), H → L + 1 (63%) |
2 | 1.66 | 745 | B2 | H → L (98%) |
3 | 3.09 | 401 | 1.2834 | B2 | H − 3 → L (10%), H − 1 → L + 1 (62%), H → L (27%) |
3 | 1.99 | 623 | B1 | H − 1 → L (84%), H → L + 1 (15%) |
4 | 3.16 | 393 | 1.6972 | B1 | H − 1 → L (62%), H → L + 1 (37%) |
4 | 2.06 | 602 | B2 | H − 1 → L + 1 (97%) |
6 | 3.54 | 350 | 0.5462 | B2 | H − 3 → L (87%) | 5 | 2.84 | 436 | A2 | H − 2 → L (88%) |
8 | 3.62 | 343 | 0.0909 | B1 | H − 3 → L + 1 (98%) | 6 | 2.90 | 428 | B2 | H − 3 → L (82%) |
19 | 3.95 | 314 | 0.0267 | B1 | H − 10 → L (39%), H − 8 → L + 1 (57%) |
7 | 3.08 | 403 | A1 | H − 2 → L + 1 (91%) |
20 | 3.95 | 314 | 0.0216 | B2 | H − 14 → L (15%), H − 11 → L (56%), H − 10 → L + 1 (14%), H − 8 → L (10%) |
8 | 3.15 | 393 | A2 | H − 16 → L + 1 (10%), H → L + 2 (74%) |
|
|
|||||||||
|
|
|
|
|
|
|
|
|
|
|
1 | 1.92 | 645 | 0.304 | A′ | H − 1 → L + 1 (16%), H → L (84%) |
1 | 1.22 | 1020 | A″ | H → L + 1 (98%) |
2 | 1.92 | 645 | 0.3039 | A″ | H − 1 → L (16%), H → L + 1 (84%) |
2 | 1.22 | 1020 | A′ | H → L (98%) |
3 | 2.89 | 430 | 1.2029 | A′ | H − 1 → L + 1 (74%), H → L (14%) |
3 | 2.02 | 615 | A″ | H − 1 → L (95%) |
4 | 2.89 | 430 | 1.2026 | A″ | H − 1 → L (74%), H → L + 1 (14%) |
4 | 2.02 | 615 | A′ | H − 1 → L + 1 (95%) |
10 | 3.13 | 396 | 0.2053 | A″ | H − 5 → L (89%) | 5 | 2.67 | 464 | A″ | H − 3 → L (13%), H − 2 → L + 1 (13%), H → L + 2 (56%) |
11 | 3.13 | 396 | 0.2056 | A′ | H − 5 → L + 1 (89%) | 6 | 2.76 | 449 | A′ | H − 7 → L + 1 (16%), H − 6 → L (16%), H − 3 → L + 1 (29%), H − 2 → L (29%) |
15 | 3.20 | 387 | 0.078 | A″ | H − 8 → L (80%) | 7 | 2.88 | 431 | A′ | H − 3 → L + 1 (46%), H − 2 → L (46%) |
16 | 3.20 | 387 | 0.0778 | A′ | H − 8 → L + 1 (80%) | 8 | 2.88 | 430 | A″ | H − 3 → L (46%), H − 2 → L + 1 (46%) |
21 | 3.45 | 360 | 0.0351 | A′ | H − 10 → L (12%), H − 9 → L (80%) |
9 | 2.93 | 424 | A′ | H − 5 → L + 1 (82%) |
22 | 3.66 | 339 | 0.039 | A″ | H − 10 → L + 1 (13%), H − 9 → L + 1 (80%) |
10 | 2.93 | 424 | A″ | H − 5 → L (82%) |
Table 4.
Selected values of the calculated singlet-singlet (

Figure 5.
Calculated and measured absorption spectra of TPP and calculated absorption spectrum of protonated-TPP (H4TPP).
TPP | |||
---|---|---|---|
H | π(Cα─Cm─Cα/Cβ─Cβ) + n(N and Cϕ) | L/L + 1 | π*(Cβ─Cβ/Cβ─Cα/Cm─Cα) + n*(N(H)) |
H − 1 | π(Cα─Cβ) | L + 2 | π*(Cα─Cm) + n*(Cϕ) |
H − m m = 2, 3 |
π(Cβ─Cβ)/ π(N─Cα─Cm) | L + m m = 3–6 |
π*(C─C in phenyl) |
Table 5.
Bond type of the highest occupied molecular orbitals (H − m) and the lowest unoccupied molecular orbitals (L + m), m = 0, 1, 2, ….
Furthermore, Jiang
Additionally, the results of the calculations for TPP (in solution/water) indicate the possibility of an IC process from S6 (B2 at 350 nm)/S4 (B1 at 393 nm)/S3 (B2 at 401 nm) to S2 (B1 at 535 nm) and S1 (B2 at 571 nm), which are verified by experimental measurements of the fluorescence spectrum of the TPP in different environments. Moreover, based on theoretical predictions, there are strong surface crossings between the singlet-triplet excited states of the TPP: S4 (B1 at 393 nm) and T8 (A2 at 393 nm), and S3 (B2 at 401 nm) and T7 (A1 at 403 nm), which may cause an ISC process in the excited state.
The results of the calculated electronic energy states of diprotonated-TPP molecule (H4TPP) indicate the existence of an IC process from the S3/4 (A′ and A"″ at 430 nm) to S1/2 (A′ and A″ at 645 nm), in addition to possibility of an ISC process between the S3/4 (A′ and A″ at 430 nm) and T7/8 (at 431 and 430 nm, with symmetry A′ and A″, respectively).
3.4. Calculated electronic spectra of TSPP, H4TSPP, and H8TSPP
In the Q-band region, while the calculations indicate the presence of two weak transitions at 573 nm (S0 → S1 with symmetry B2 and
In the B-band (Soret band) region, while the calculated spectrum of the TSPP exhibited two strong bands at 403 nm (S0 → S3 with B2 symmetry and
Akins
TSPP:S0 →Sn | S0 →Tn | |||||||||
---|---|---|---|---|---|---|---|---|---|---|
Sn | (eV) | (nm) | f | Sym | Major contrib's | Tn | (eV) | (nm) | Sym | Major contrib's |
5 | 2.16 | 573 | 0.0419 | B2 | H − 1 → L + 1 (32%), H → L (67%) |
1 | 1.40 | 884 | B1 | H − 1 → L (16%), H → L + 1 (86%) |
6 | 2.31 | 536 | 0.0506 | B1 | H − 1 → L (36%), H → L + 1 (64%) |
2 | 1.67 | 744 | B2 | H → L (97%) |
10 | 3.07 | 403 | 1.4382 | B2 | H − 1 → L + 1 (62%), H → L (28%) |
3 | 1.99 | 624 | B1 | H − 1 → L (84%), H → L + 1 (15%) |
11 | 3.13 | 396 | 1.8378 | B1 | H − 1 → L (62%), H → L + 1 (36%) |
4 | 2.05 | 604 | B2 | H − 1 → L + 1 (97%) |
36 | 3.49 | 355 | 0.3924 | B2 | H − 10 → L (83%) | 5 | 2.84 | 437 | A2 | H − 9 → L (49%), H − 7 → L (39%) |
38 | 3.56 | 348 | 0.0392 | B1 | H − 10 → L + 1 (28%), H − 8 → L (69%) |
6 | 2.89 | 429 | B2 | H − 11 → L (36%), H − 10 → L (48%) |
47 | 3.64 | 341 | 0.2178 | B2 | H − 11 → L (83%) | 7 | 3.07 | 404 | A1 | H − 9 → L + 1 (39%), H − 7 → L + 1 (51%) |
48 | 3.77 | 329 | 0.0936 | B1 | H − 11 → L + 1 (82%), H − 10 → L + 1 (11%) |
8 | 3.14 | 395 | A2 | H → L + 2 (70%) |
f | ||||||||||
1 | 1.85 | 669 |
|
B2 | H − 5 → L + 1 (13%), H → L (87%) |
1 | 1.17 | 1056 | B2 | H → L (97%) |
2 | 1.85 | 669 |
|
B1 | H − 5 → L (13%), H → L + 1 (87%) |
2 | 2.01 | 618 | B1 | H − 5 → L (95%) |
4 | 2.35 | 528 |
|
A1 | H − 3 → L + 1 (47%), H − 2 → L (53%) |
3 | 2.34 | 530 | A1 | H − 3 → L + 1 (47%), H − 2 → L (52%) |
5 | 2.35 | 528 |
|
B2 | H − 4 → L (53%), H − 1 → L + 1 (47%) |
5 | 2.34 | 529 | B1 | H − 4 → L + 1 (47%), H − 1 → L (52%) |
10 | 2.40 | 518 |
|
B2 | H − 4 → L (47%), H − 1 → L + 1 (53%) |
9 | 2.39 | 518 | A1 | H − 3 → L + 1 (53%), H − 2 → L (47%) |
12 | 2.74 | 452 |
|
B1 | H − 6 → L (79%), H − 5 → L (15%) |
13 | 2.51 | 494 | A2 | H − 8 → L + 1 (34%), H − 7 → L (35%), H → L + 2 (23%) |
13 | 2.75 | 451 |
|
B2 | H − 6 → L + 1 (83%), H − 5 → L + 1 (12%) |
14 | 2.62 | 474 | B1 | H − 6 → L (90%) |
16 | 2.81 | 441 |
|
B2 | H − 9 → L (17%), H − 6 → L + 1 (16%), H − 5 → L + 1 (57%) |
17 | 2.65 | 467 | A1 | H − 8 → L (46%), H − 7 → L + 1 (46%) |
18 | 2.81 | 441 |
|
B1 | H − 9 → L + 1 (16%), H − 6 → L (19%), H − 5 → L (54%) |
18 | 2.69 | 462 | A2 | H − 8 → L + 1 (47%), H − 7 → L (47%) |
19 | 2.97 | 417 |
|
B1 | H − 13 → L (14%), H − 10 → L + 1 (74%) |
19 | 2.72 | 455 | A2 | H − 8 → L + 1 (13%), H − 7 → L (12%), H → L + 2 (63%) |
22 | 3.10 | 400 |
|
B2 | H − 13 → L + 1 (12%), H − 10 → L (82%) |
20 | 2.96 | 419 | A2 | H − 12 → L + 1 (43%), H − 11 → L (47%) |
1 | 2.00 | 620 | 0.2467 | B2 | H − 1 → L + 1 (23%), H → L (77%) |
1 | 1.31 | 946 | B2 | H → L (96%) |
2 | 2.00 | 619 | 0.2377 | B1 | H − 1 → L (23%), H → L + 1 (77%) |
2 | 1.31 | 945 | B1 | H → L + 1 (96%) |
3 | 2.92 | 424 | 1.7153 | B1 | H − 1 → L (74%), H → L + 1 (23%) |
3 | 1.94 | 639 | B1 | H − 1 → L (94%) |
4 | 2.92 | 424 | 1.7145 | B2 | H − 1 → L + 1 (75%), H → L (22%) |
4 | 1.94 | 638 | B2 | H − 1 → L + 1 (94%) |
7 | 3.32 | 374 | 0.0199 | B2 | H − 4 → L (92%) | 5 | 2.74 | 452 | A2 | H → L + 2 (72%) |
8 | 3.32 | 373 | 0.0240 | B1 | H − 4 → L + 1 (92%) | 6 | 3.00 | 413 | A1 | H − 7 → L (26%), H − 6 → L + 1 (26%), H − 3 → L (13%), H − 2 → L + 1 (18%) |
11 | 3.41 | 363 | 0.0560 | B1 | H − 5 → L (88%) | 7 | 3.08 | 403 | A2 | H − 3 → L + 1 (37%), H − 2 → L (40%) |
12 | 3.42 | 363 | 0.0614 | B2 | H − 5 → L + 1 (88%) | 8 | 3.09 | 401 | A1 | H − 3 → L (36%), H − 2 → L + 1 (34%), H − 1 → L + 2 (14%) |
15 | 3.50 | 355 | 0.0822 | B1 | H − 8 → L (94%) | 9 | 3.16 | 393 | B2 | H − 5 → L + 1 (16%), H − 4 → L (66%) |
16 | 3.50 | 355 | 0.0918 | B2 | H − 8 → L + 1 (94%) | 10 | 3.16 | 393 | B1 | H − 5 → L (17%), H − 4 → L + 1 (65%) |
20 | 3.59 | 346 | 0.0365 | B2 | H − 10 → L (10%), H − 9 → L (84%) |
11 | 3.20 | 387 | A2 | H − 12 → L + 1 (11%), H − 11 → L (12%), H − 7 → L + 1 (29%), H − 6 → L (30%) |
21 | 3.69 | 336 | 0.0363 | B1 | H − 10 → L + 1 12%), H − 9 → L + 1 (81%) |
12 | 3.23 | 384 | B1 | H − 8 → L (66%) |
Table 6.
The selected values of the calculated singlet-singlet (
Also, Akins and coworkers have measured fluorescence spectra of free-base TSPP (pH = 12), monomeric H4TSPP (pH = 4.5), and aggregate H4TSPP in highly acidic situation. The authors reported that the fluorescence spectrum of the TSPP at 412 nm (B-band region) excitation displayed a peak at 642 nm with a red degraded shoulder at 702 nm. The spectrum of H4TSPP upon excitation at 432 nm in the B-band region exhibited similar structure, for example, a strong emission peak at 665 nm with relatively weak shoulder at about 716 nm [21]. Both fluorescence spectra of the TSPP and deprotonated-TSPP (H4TSPP) indicated that when excited in the Soret- or B-band region, initially internal conversion (IC) occurs from the B-band to the Q-bands, followed by a fluorescence from the lowest excited state(s) in Q-band region to the ground state S0 a sequence of:
The calculations also indicate that there might be an ISC (intersystem crossing) process between the S3 (at 403 nm)/S4 (at 396 nm) and the T7 (at 404 nm)/T8 (at 395 nm), and between the S1 (at 573 nm) and T4 (at 604nm) for TSPP (where the energy difference between S1 and T4 states is about 0.11 eV or 896 cm−1). For the H4TSPP (or dianionic-TSPP), the ISC process may occur between the S4/5 (at 528 nm) and T3,4,5,6 (at 530 and 529 nm), and S10 (at 518 nm) and T15,16,17,18 (at 518 nm), and between the S1(669 nm) and T3(618 nm) (where the energy difference between the S1(669 nm) and T3(618 nm) states is 0.15 eV or 1233 cm−1) (Figure 6). The results of the calculations suggest that, depending on competition between the IC and ISC processes, there can be ISC through vibrational coupling or potential energy surface (PES) touching between singlet and triplet states.

Figure 6.
Calculated and measured absorption spectra of TSPP and protonated-TSPP (H4TSPP).
Likewise, for the H8TSPP (dicationic-TSPP molecule), the IC process may happen from the B-bands (S3/4 at 424 nm) to the Q-bands (S1/2 at 620/619 nm). Furthermore, the energy difference between the S1/2 (at 424/424 nm) and T3/4 (at 639/637 nm) is about 0.056 eV or 455 cm−1, which may lead to a strong vibrational coupling in their excited vibroelectronic states. Owing to this small energy distance between singlet and triplet states of the H8TSPP, there would likely occur ISC that may originate from B-bands (S1/2) to triplet states (T3/4).
TSPP | H4TSPP | H8TSPP | |||
---|---|---|---|---|---|
H | π(Cα─Cm─Cα/Cβ─Cβ) + n(N and Cϕ/C(S)) | H | π(Cα─Cm─Cα/Cβ─Cβ/C─C in phenyl) + n(N) |
H | π(Cα─Cm─Cα/Cβ─Cβ) + n(N and Cϕ/C(S) in phenyl) |
H − 1 | π(Cα─Cα) | H − m m = 1–4 |
n(O) | H − 1 | π(Cα─Cβ) |
H − m m = 2–5 |
n(O) | H − 5 | π(Cα─Cβ) | H − m m = 2–5 |
π(C─C─C in phenyl) + n(Cα and Cβ, minor) |
H − 6 | π(Cβ─Cβ and Cα ─N─Cα) |
H − m m = 6–8 |
π(C─C─C in phenyl) + minor n(O and Cβ) |
H − 6/H − 7 |
π(C(S)─C/C─Cϕ in phenyl and Cm─Cα─Cβ) |
L/L + 1 | π*(Cβ─Cα/Cβ─Cβ and Cα─Cm) + n*(N(H)) | L/L + 1 | π*(Cα─Cm/Cβ─Cβ) + n*(N) | L/L + 1 | π*(Cα─Cm/Cβ─Cβ) + n*(N); |
L + 2 | π*(Cϕ─Cm/Cα─Cm) + n*(C(S), minor) | L + 2 | π*(Cϕ─Cm/Cβ ─Cα) and minor n*(C in phenyl) |
L + 2 | π*(Cϕ─Cm/Cβ─Cβ and C─ C/C─S in phenyl) |
L + m m = 3–6 |
π*(C─C in phenyl) | L + m m = 3–5 |
π*(C─C in phenyl) + n* (C(S) and Cϕ in phenyl) |
L + m m = 3–6 |
π*(C─C/C─S in phenyl) and n*(Cϕ)/n*(Cα and N, minor) |
Table 7.
Bond type of the highest occupied molecular orbitals (H − m) and the lowest unoccupied molecular orbitals (L + m), m = 0, 1, 2, …
Consequently, the results of calculated absorption spectra for the porphyrin molecules studied here (Tables 2–7) reveal several important points: (1) protonation of the N atoms at the porphyrin core and the
3.5. Relaxed potential energy surface (RPES) scan of TSPP molecule
The relaxed potential energy surface (RPES) scan was performed to calculate the ground state PES of the TSPP molecule in water by rotating one of four dihedral angles

Figure 7.
The calculated spectra of the TSPP as a function of the dihedral angle (Cα─Cm─Cϕ─C(ph)) rotation varying from 40 to 130° with 10° increment: (A) plot of dipole-allowed singlet electronic transitions S0 → Sn, with n = 1–24; (B) the relaxed potential energy surfaces of the ground state (S0) and upper singlet (Sn) and triplet (Tn) states, n = 1–24; (C)–(E) illustrate the RPES curves for the ground state S0, Q-bands and Soret bands at a low scale for a better view. It is noteworthy that only one of the four
It is to be noted that the PES curves of the upper singlet (Sn) and triplet (Tn) energy states were calculated by the following Eqs. (1) and (2), respectively:
(1) In the aforementioned equations,
4. Calculation section
The calculations were carried out in water used as solvent at the B3LYP level of the density functional theory (DFT) [34, 35] with the 6-311G(d,p) basis set [36]. The solvent effects were considered by using the self-consistent reaction field (SCRF) calculations [37] with the conductor-like polarizable continuum model [38–40] and a dielectric constant of 78.39 for water; SCRF = (CPCM, solvent = water) as implemented within the Gaussian 09 software package [41]. All compounds studied here were optimized to minima on their ground state relaxed potential energy surfaces (RPESs) that were verified by revealing the absence of imaginary frequencies in calculated vibrational spectra. Time-dependent DFT (TD-DFT) was performed to calculate the first 24 singlet-singlet (S0 → Sn) and singlet-triplet (S0 → Tn; n = 1 to 24) vertical electronic transitions in water. Finally, to investigate the dependence of the potential energy of the ground state (S0) and excited states (Sn and Tn) on the rotation of the Cm─C
We would like to point out that the electron densities in HOMO and LUMO molecular orbitals, and electronic spectra of the molecules studied here were plotted using GaussSum software [42].
Acknowledgments
We would like to thank the following: the U.S. National Science Foundation (NSF) for support of research efforts under grant no. HRD-08-33180 and Ömer Andaç (of the Chemistry Department of Ondokuz Mayıs University) for kindly making available computing facilities and software setup. We also thank TUBITAK ULAKBIM, High Performance and Grid Computing Center (TR-Grid e-Infrastructure) for performing the calculations reported in this work.
References
- 1.
Dolphin, D. The Porphyrins ; Dolphin, D., Ed.; Academic Press: New York, 1978; Volume 1–3. - 2.
Shelnutt, J. A,; Alston, K.; Findsen, E. W.; Ondrias, M. R.; Rifkind, J. M. In Polphyrins: Excited States and Dynamics ; Gouterman, M., Rentzepis, P. M., Straub, K. D., Eds.; ACS Symposium Series 321; American Chemical Society: Washington, DC, 1986; Chapter 16. - 3.
Brunner, H.; Nishiyama, H.; Itoh, K. in Catalytic Asymmetric Synthesis ; Ojima, I. (Ed.), VCH, New York, 1993, pp. 303–322. - 4.
Lu, J., Li, H., Liu, S., Chang, Y-C., Wu, H-P., Cheng, Y., Diau, E. W-G., Wang, M. Novel porphyrin-preparation, characterization, and applications in solar energy conversion. Phys.Chem.Chem.Phys . 2016, 18, 6885. - 5.
Bian, Y.; Zhang, Y.; Ou, Z.; Jiang, J. Chemistry of Sandwich Tetrapyrrole Rare Earth Complexes. Handbook of Porphyrin Science, World Scientific Publishing, Singapore, 2011, 14, 249–460. - 6.
Holten, D.; Bocian, D.F. Probing electronic communication in covalently linked multiporphyrin arrays. A guide to the rational design of molecular photonic devices. J.S. Acc. Chem. Res .2002 ,35 , 57–69. - 7.
Hopfield, J.J.; Onuchic, J.N.; Beratan, D.N. A molecular shift register based on electron transfer. Science 1988 ,241 , 817–820. - 8.
Bonnett, R. Photosensitizers of the porphyrin and phthalocyanine series for photodynamic theory. Chem. Soc. Rev .1995 ,24 , 19–33. - 9.
Andrade, S.M.; Costa, S.M.B. Spectroscopic studies on the interaction of a water soluble porphyrin and two drug carrier proteins. Biophys. J. 2002 ,82 , 1607–1619. - 10.
Ben-Hur, E.; Horowitz, B. Advances in photochemical approaches for blood sterilization. Photochem. Photobiol. 1995 ,62 , 383–388. - 11.
Uhehara, K.; Hioki, Y.; Mimuro, M. The chlorophyll a aggregate absorbing near 685 nm is selectively formed in aqueous tetrahydrofuran.Photochem. Photobiol. 1993 ,58 , 127–132. - 12.
Tominaga, T.T.; Yushmanov, V.E.; Borissevitch, I.E.; Imasato, H.; Tabak, M. Aggregation phenomena in the complexes of iron tetraphenylporphine sulfonate with bovine serum albumin. J. Inorg. Biochem .1997 , 65, 235–244. - 13.
Togashi, D.M.; Costa, S.M.B. Absorption, fluorescence and transient triplet-triplet absorption spectra of zinc tetramethylpyridylporphyrin in reverse micelles and microemulsions of aerosol OT-(AOT). Phys. Chem. Chem. Phys .2000 ,2 , 5437–5444. - 14.
Ferreira, G. C.; Kadish, K. M.; Smith, K. M.; Guilard, R., Eds. In The Handbook of Porphyrin Science ; World Scientific Publishers: Singapore, 2013, Vol. 27. - 15.
Gillam, M. E.; Hunter, G. A.; Ferreira, G. C. The ultimate step of heme biosynthesis: orchestration between iron trafficking and porphyrin synthesis. In Heme Biochemistry ; Ferreira, G.C., Ed., ofThe Handbook of Porphyrin Science series, Kadish, K. M.; Smith, K. M.; Guilard, R. series Eds., World Scientific Publishing Co., New Jersey, USA, 2013, Vol. 26, pp. 129–189. - 16.
Borissevitch, I.E.; Tominaga, T.T.; Imasato, H.; Tabak M. Fluorescence and optical absorption study of interaction of two water soluble porphyrins with bovine serum albumin: the role of albumin and porphyrin aggregation. J. Luminesc. 1996 ,69 , 65–76. - 17.
Huang, C.Z.; Li, Y.F.; Li, K.A.; Tong, S.Y. Spectral characteristics of the aggregation of α, β, γ, δ-tetrakis( p -sulfophenyl)porphyrin in the presence of proteins.Bull. Chem. Soc. Jpn. 1998 ,71 , 1791–1797. - 18.
Maiti, N.C.; Ravikanth, M.; Mazumdar, S.; Periasamy N. Fluorescence dynamics of noncovalent linked porphyrin dimers and aggregates. J. Phys. Chem. 1995 ,99 , 17192–17197. - 19.
Ohno, O.; Kaizu, Y.; Kobayashi, H. J-aggregate formation of a water-soluble porphyrin in acidic aqueous media. J. Chem. Phys. 1993 ,99 , 4128–4139. - 20.
Akins, D.L.; Zhu, H.-R.; Guo, C. Absorption and Raman scattering by aggregated meso-tetrakis(p-sulfonatophenyl)porphine. J. Phys. Chem. 1994 ,98 , 3612–3618. - 21.
Akins, D.L.; Özçelik, S.; Zhu, H.R.; Guo, C. Fluorescence decay kinetics and structure of aggregated tetrakis( p -sulfonatophenyl)porphyrin.J. Phys. Chem. 1996 ,100 , 14390–14396. - 22.
Labanowski, J.K.; Andzelm, J.W. Density Functional Theory Methods in Chemistry ; Springer-Verlag: New York, NY, USA, 1991. - 23.
Aydin, M. DFT and Raman spectroscopy of porphyrin derivatives: tetraphenylporphine (TPP). Vib. Spectrosc. 2013 ,68 , 141–152. - 24.
Słota, R.; Broda, M.A.; Dyrda, G.; Ejsmont, K.; Mele, G. Structural and molecular characterization of meso-substituted zinc porphyrins: a DFT supported study. Molecules 2011 ,16 , 9957–9971. - 25.
Wang, C.; Yang, G.; Li, J.; Mele, G.; Słota, R.; Broda, M.A.; Duan, M.; Vasapollo, G.; Zhang, X.; Zhang, F.X. Novel meso-substituted porphyrins: synthesis, characterization and photocatalytic activity of their TiO2-based composites. Dyes Pigments 2009 ,80 , 321–328. - 26.
Zhang, Y.H.; Zhao, W.; Jiang, P.; Zhang, L.J.; Zhang, T.; Wang, J. Structural parameters and vibrational spectra of a series of zinc meso-phenylporphyrins: a DFT and experimental study. Spectrochim. Acta A 2010 ,75 , 880–890. - 27.
Yao, P.; Han, S.; Zhang, Y.; Zhang, X.; Jiang, J. Structures and spectroscopic properties of mesotetrasubstituted porphyrin complexes: Meso-Substitutional and central metallic effect study based on density functional theory calculations. Vib. Spectrosc .2009 ,50 , 169–177. - 28.
De Oliveira, V.E.; Cimini Correa, C.; Pinheiro, C.B.; Diniz, R.; de Oliveira, L.F.C. Structural and spectroscopy studies of the zinc complex of p-hydroxyphenylporphyrin. J. Mol. Struct .2011 ,995 , 125–129. - 29.
Aydin, M. Comparative study of the structural and vibroelectronic properties of porphyrin and its derivatives. Molecules 2014 ,19 , 20988–21021. - 30.
Edwards, L.; Dolphin, D. H.; Gouterman, M.; Adler, A. D. Porphyrins XVII. Vapor absorption spectra and redox reactions: Tetraphenylporphins and porphin. J Mol Spectrosc 1971, 38, 16. - 31.
Murphy, M. J.; Siegel, L. M.; Kamin, H.; Rosenthal, D. Reduced nicotinamide adenine dinucleotide phosphate-sulfite reductase of enterobacteria. II. Identification of a new class of heme prosthetic group: an iron-tetrahydroporphyrin (isobacteriochlorin type) with eight carboxylic acid groups. J Biol Chem. 1973, 248(8), 2801–2814. - 32.
Lan, M.; Zhao, H.; Yuan, H.; Jiang, C.; Zuo, S.; Jiang, Y. Absorption and EPR spectra of some porphyrins and metalloporphyrins. Dyes Pigments 2007 ,74 , 357–362. - 33.
Zhang, Y.H.; Chen, D.M.; He, T.; Liu, F.C. Raman and infrared spectral study of meso-sulfonatophenyl substituted porphyrins (TPPSn, n_/1, 2A, 2O, 3, 4). Spectrochim. Acta Part A 2003 ,59 , 87–101. - 34.
Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys .1993 ,98 , 5648–5652. - 35.
Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988 ,37 , 785–789. - 36.
Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions.. J. Chem. Phys. 1980 ,72 , 650–654. - 37.
Tomasi, J.; Mennucci, B.; Cammi, R. Quantum mechanical continuum solvation Models. Chem. Rev .2005 ,105 , 2999–3093. - 38.
Barone, V.; Cossi, M. Quantum calculation of molecular energies and energy gradients in solution by a conductor solvent model . J. Phys. Chem. A 1998 ,102 , 1995–2001. - 39.
Miertus, S.; Scrocco, E.; Tomasi, J. Electrostatic interaction of a solute with a continuum. A direct utilization of ab initio molecular potentials for the prevision of solvent effects. Chem. Phys .1981 ,55 , 117–129. - 40.
Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem .2003 ,24 , 669–681. - 41.
Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al .Gaussian 09, Revision A. 02. Gaussian. Inc.: Wallingford, CT, 2009, 200. - 42.
O’Boyle, N. M.; Tenderholt, A. L.; Langner, K. M. A library for package-independent computational chemistry algorithms. J. Comp. Chem .2008 ,29 , 839–845.