Abstract
Aromatically π-extended porphyrins possess exceptionally intense one-photon (1P) and sometimes two-photon (2P) absorption bands, presenting interest for construction of optical imaging probes and photodynamic agents. Here we investigated how breaking the molecular symmetry affects linear and 2PA properties of π-extended porphyrins. First, we developed the synthesis of porphyrins fused with two phthalimide fragments, termed syn-diarylphthalimidoporphyrins (DAPIP). Secondly, the photophysical properties of H2, Zn, Pd and Pt DAPIP were measured and compared to those of fully-symmetric tetraarylphthalimidoporphyrins (TAPIP). The data were interpreted using DFT/TDDFT calculations and Sum-Over-States (SOS) formalism. Overall, the picture of 2PA in DAPIP was found to resemble that in centrosymmetric porphyrins, indicating that symmetry breaking, even as significant as by syn-phthalimido-fusion, induces a relatively small perturbation to the porphyrin electronic structure. Collectively, the compact size, versatile synthesis, high 1PA and 2PA cross-sections and bright luminescence makes DAPIP valuable chromophores for construction of imaging probes and other bio-applications.
Keywords: Porphyrins, π-extended porphyrins, phosphorescence, fluorescence, two-photon absorption
Graphical Abstract
Introduction
Tetrapyrroles capable of efficient two-photon absorption (2PA) present interest for several areas of technology and medicine, including optical power limiting,1,2 up-conversion lasing,3 three-dimensional optical data storage,4 biological imaging5–10 and two-photon photodynamic therapy (PDT).11–15 Over the past years porphyrin-based systems have been developed with 2PA cross-sections reaching into thousands of Göppert-Mayer (GM) units, including covalently linked porphyrin arrays, porphyrins with expanded electronic system as well as self-assembled non-covalent porphyrin oligomers.16–23 However, large size, inherently poor solubility, lack of synthetic methods for targeted functionalization and, in many cases, the presence of intramolecular quenching pathways render many of these otherwise spectacular systems of low practical utility, especially in view of those applications where precise control over excited state dynamics is required. In contrast, compact monomeric porphyrins can be constructed with tailored functional groups and high solubility, while rational synthetic strategies exist for manipulation of their excited states in order to reach optimal performance in applications. In view of these advantages, compact porphyrinoids capable of efficient 2PA present an attractive synthetic target, and their development remains an important area of porphyrin chemistry.
Porphyrins in which the pyrrolic fragments are fused via β,β’-positions with external aromatic rings (e.g. benzene, naphthalene, anthracene) are known as aromatically π-extended porphyrins.24–27 π-Extended porphyrins possess intense one-photon (1P) and in some cases strong two-photon (2P) bands in the near infrared spectral region and exhibit high photo- and chemical stability. In recent years synthetic chemistry of π-extended porphyrins has been significantly advanced,28–38 making many of these once exotic systems available for practical applications. However, most attention has been devoted to the synthesis of porphyrins with symmetric π-extension, while asymmetric molecules, such as syn-dibenzo-, syn-dinaphtho- etc porphyrins39–48 have been reported and studied on much fewer occasions. Asymmetric π-extension induces different perturbation to the macrocycle electronic structure compared to the symmetric fusion,47,48 resulting in different photophysical properties, especially in relation to 2PA.
Pt and Pd complexes porphyrins are used for the construction of phosphorescent sensors for oxygen,49,50 including probes with enhanced 2PA for two-photon phosphorescence lifetime microscopy (2PLM)6,10,51,52 - an imaging technique that has gained attention in neuroscience,53–56 stem cell biology57,58 and some other areas.59,60 Enhancement of 2PA in phosphorescent porphyrins while preserving their high emissivity presents an interesting problem. Metalloporphyrin molecule belongs to the D4h symmetry class and thus possesses center of inversion symmetry. In closed-shell systems with centrosymmetric (gerade) ground state wavefunctions the parity selection rules governing electric dipole transitions are mutually exclusive, rendering 1P-allowed ungerade (u) B (Soret) and Q states inaccessible by 2PA. At the same time, 2P-active gerade (g) states in regular (non-extended) porphyrins lie at such high energies that their 2PA excitation is overshadowed by linear (1P) absorption to low-lying vibronic sublevels of the Q-state or, in some cases, directly to the triplet state.61 While conjugation of porphyrins into arrays and other assemblies has been shown to greatly increase 2PA,16,21,62–67 the triplet emissivity of the resulting constructs is usually diminished. Recently, we found that g-states in π-extended porphyrins can be stabilized upon addition of electron-withdrawing groups to the peripheral extension rings.68,69 For example, in symmetric tetraarylphthalimidoporphyrins (TAPIP)69,70 the g-states may occur even below the B state level, while the rigidity of the molecular skeleton causes meso-unsubstituted TAPIP to be highly emissive.69 In view of these findings it was of interest to explore asymmetric counterparts of TAPIP, i.e. porphyrins in which phthalimido-fragments are annealed only on one side of the macrocycle, decreasing its symmetry to C2v. Our previous studies48,68 have shown that syn-π-extension results in porphyrins with absorption bands that are sharp, strong and only minimally overlapping with those of the symmetrically extended counterparts. From the practical viewpoint, these features are advantageous for the construction of multichromophoric sensor systems.
Here we present the synthesis, structural characterization and photophysical properties of asymmetric syn-diarylphthalimidoporphyrins (DAPIPs), including their 2PA spectra. Our results delineate the effects of asymmetric π-extension on the tetrapyrrolic electronic system, showing that spectroscopic properties of porphyrins are highly resistant to symmetry breaking. From the practical perspective DAPIP were found to be valuable chromophores for the construction of imaging probes and other biological applications.
Results and Discussion
Synthesis.
The asymmetric syn-diarylphthalimidioporphyrins (DAPIP) synthesized in this study are shown in Chart 1.
Chart 1.
syn-Diarylphthalimidoporphyrins (DAPIP).
In the past in collaboration with the group of A. Cheprakov25,34 we developed a synthetic approach to π-extended porphyrins, in which the key step is oxidative aromatization of precursor porphyrins annealed with nonaromatic rings.32,33 This method allowed us to generate a variety of syn-dibenzo- and syn-dinaphthoporphyrins with solubilizing substituents.48 In such asymmetric systems the lowest energy absorption bands (Q bands) are positioned between those of non-extended (regular) porphyrins and porphyrins with the corresponding symmetric π-extension. Here we set out to explore the synthesis of meso-unsubstituted and meso-substituted DAPIP employing the same general strategy.
The developed synthesis (Schemes 1 and 2) consists of the assembly of ditetrahydrophthalimido-dipyrromethanes 6, 7 (Scheme 1), [2+2] condensation of 6 and 7 with unsubstituted dipyrromethanes to form porphyrininc precursors with fused non-aromatic rings (Scheme 2) and subsequent oxidative aromatization of the latter (Scheme 2). The synthesis of tetrahydrophthalimido-fused dipyrromethanes (6, 7) begins with the preparation of pyrrole-ester 3 via Diels-Alder cycloaddition of sulfoleno-pyrrole benzyl ester 1 to maleimide 2.28 1 was synthesized using an earlier published method,71 but with a slight modification. Slow addition of reagents (during 4-5 h) and keeping the temperature low (−70°C) allowed us to improve the yield by minimizing competing degradation of intermediates. The target pyrrole-ester 3 was isolated in 78% yield.
Scheme 1.
Synthesis of Dipyrromethanes.
Reagents and conditions: (a) 1,2,4-trichlorobenzene, reflux, 4h (78%); (b) CH2(OMe)2 or PhCHO, p-TsOH, AcOH, rt, 96h, R=H (74%), R=Ph (79%); (c) H2, Pd/C, THF, rt, 24h; (d) ethylene glycol, 150°C.
Scheme 2.
Synthesis of Diarylphthalimidoporphyrins.‡
Reagents and conditions: (a) (i) Zn(OAc)2 2H2O, benzene, Ar, reflux, 2-3h; (ii) air, reflux, 12-14h; (b) DDQ, ethyl benzoate, 140°C, 6h: Zn-12 (65%); (c) 50% HCl: H2-12 (86%), H2-10 (17%)†, H2-11 (14%)† (d) M=Pt: Pt(acac)2, PhCN, reflux, 8h: Pt-12 (65%)‡, Pt-13 (63%)‡; M=Pd: Pd(OAc)2, PhCN, reflux, 1h: Pd-12 (63%)‡, Pd-13 (72%).‡ † over steps a and c. ‡ over steps d and b. ‡The identity of compounds M-10 and M-11 (M=Zn, Pd, Pt) was confirmed by MALDI-TOF spectroscopy, after which they were introduced into the subsequent transformations without further characterization.
Condensation of 3 with benzaldehyde or dimethoxymethane was carried out in acetic acid, following the literature procedures,34,48,72 affording tetrahydrophthalimido-fused dipyrromethanes-esters 4 and 5 in 74% and 79 % yields, respectively. The condensation required rather long time reaction times (72-96 h) to come to completion, likely as a consequence of the large steric bulk of pyrrole 3. The resulting dipyrromethanes-esters 4 and 5 were stable for months when we stored at ~0°C.
Hydrogenolysis of pyrrole-esters 4 or 5 (H2, Pd/C) gave the corresponding carboxylic acids. However, subsequent decarboxylation upon treatment with TFA led to significant scrambling, and the target α,α’-unsubstituted dipyrromethanes 6 and 7 could be isolated only in low yields (~20%). A more successful procedure entails heating of the esters to 150°C in ethylene glycol,32 giving the target pyrroles in 40-50% yields. Note that 6 and 7 are unstable at ambient temperatures. Therefore, the decarboxylation should be carried out immediately prior to the porphyrin synthesis.
The assembly of the non-aromatized porphyrin-precursors via [2+2] condensation followed the method developed by Lindsey et al.73 Propylimine substituents in α,α’-positions of dipyrromethanes 8, 9 are synthetically equivalent to formyl groups and have been shown to minimize scrambling in the porphyrin condensation.73,74 The α,α’-propyliminodipyrromethanes were obtained by reacting pyrrole with suitable aldehydes (or their derivatives) in the presence of InCl3, followed by formylation of the α,α’-positions and subsequent condensation with Propylamine 73,74
The [2+2] condensation was carried out under the conditions developed previously for similar syntheses,48 i.e. as a two-step process, which helped to minimize scrambling. First, 6 or 7 were condensed with 8 or 9 in the presence of Zn(OAc)2 in a refluxing non-polar solvent (benzene or toluene) under inert (argon) atmosphere. The formation of the corresponding dihydroporphyrinogens was monitored by MALDI-TOF spectroscopy, and after their concentrations peaked (usually after ~3 h), the air was let in, and the mixtures were refluxed for 12-15 h. Our previous results indicate that the above conditions are optimal for efficient oxidation of dihydroporphyrinogens into porphyrins, while avoiding oxidation of dipyrromethanes to dipyrrins and other side reactions.48 In the present case the method was successfully used to synthesize both meso-unsubstituted and 5,15-diphenyldicylcohexenoporphyrins, which were obtained as Zn complexes in 17-20% yields.
Aromatization of dicylohexenoporphyrins into the target DAPIP was performed using DDQ as an oxidant. meso-Unsusbtituted Zn-12 was obtained in 82% yield upon refluxing of Zn dicyclohexenoporphyrin (Zn-10) with DDQ in ethyl benzoate for 6 h. Use of ethyl benzoate was found to greatly facilitate aromatization of tetrahydrophlalimido-fused porphyrins with pendant ester groups, as both starting materials and products are high soluble in this solvent unlike in more commonly used toluene.48 The product formation was observed by change in the color from red to deep green. The free-base DAPIP (H2-12) was obtained upon facile demetalation of Zn-12 with HCl.
The synthesis of Pt and Pd complexes of DAPIP followed the same order: first the corresponding metal complexes of the precursor dicylohexenoporphyrins were obtained and then oxidized using DDQ. Zn-10 and Zn-11 were demetalated into free-bases H2-10 and H2-11 upon treatment with HCl. Insertion of Pt occurred upon refluxing of H2-10 and H2-11 with 3-4-fold molar excess of Pt(acac)2 in benzonitrile for 6-8 h.75 In addition, two other methods of Pt insertion were tested, using benzoic acid melt as a solvent76 and microwave-assisted synthesis,77 however, the traditional benzonitrile method in the present case gave superior results. Similarly, Pd complexes were prepared by treating the free-bases with Pd(OAc)2 (3-4 eq) in benzonitrile for 1 h. It is noteworthy that insertion of both Pt and Pd led to partial oxidation of the peripheral cyclohexeno rings.69 Consequently, the metal complexes M-10 and M-11 (M=Pd, Pt) were isolated as crude mixtures and subjected to the final aromatization without purification.
Aromatization of M-10 and M-11 was carried out in ethyl benzoate, since more traditional solvents, e.g. toluene or dioxane,32 gave poor results due to the solubility issues: aromatization could not be completed even after 18-20 h of reflux. In contrast, oxidation in ethyl benzoate upon heating to 140°C for 6-7 h afforded Pt and Pd complexes of DAPIP, M-12 and M-13 (M=Pd, Pt) in 63-65% yields.
Structural features.
All porphyrins were characterized by the standard analytical methods. In addition the molecular structure of the Pt complex of 5,15-diphenyl-DAPIP (Pt-13) was determined by X-ray analysis (Fig. 1; SI 3).
Figure 1.
X-ray crystallographic structure of Pt-13. The structure was analyzed by the Normal Mode Structural Decomposition Analysis (NSD)78 revealing rather small out-of-plane distortion (Doop=0.83Å) dominated by saddling (B2u).
The meso-aryl rings in the 5,15-positions of Pt-13 are tilted by ~70° relative to the mean-square plane of the macrocycle, which is standard for regular nominally planar porphyrins. The aryl groups attached to the phthalimide nitrogens are forced by the 2,6-ester substituents into almost orthogonal positions relative to the porphyrin plane. Such orthogonal orientation is critical for maintaining solubility of DAPIP and similar porphyrins.69 In the absence of orthogonal aryl groups, DAPIP, especially those without meso-aryls (such as M-12), would aggregate and become difficult to characterize and/or use in subsequent reactions.
To relieve the steric strain between the meso-aryl ring and the proximal hydrogen atoms in the benzo groups, the macrocycle in Pt-13 undergoes slight in-plane distortion, leaving more space between the two phthalimide substituents. Consequently, the distance between the two pyrrolic rings fused with phthalimides (a=5.07Å) is slightly larger than the analogous distances on the adjacent sides of the molecule (b=4.97Å). The in-plane distortion is a result of a small inward rotation of the phthalimide groups, which affects only minimally the rest of the macrocycle. A similar although larger effect was observed in 5,15-diaryltetrabenzoporphyrins.79
In order to quantify the distortion of the macrocycle in Pt-13, its structure was subjected to the Normal-mode Structural Decomposition (NSD) analysis, as described by Shelnutt and co-workers.78 In this method the macrocycle’s deviation from the pure D4h symmetric configuration is expanded in the basis of in-plane and out-of plane distortion modes, corresponding to the vibrational modes of distinct symmetry types, such as saddling (B2u), ruffling (B1u), doming (A2u) etc. The square roots of the mean squared deviations of all in-plane and all out-of-plane distortions are designated as Dip (in-plane) and Doop (out-of-plane), respectively. The macrocycle in Pt-13 exhibits moderate out-of-plane (Doop=0.83Å) and a small in-plane (Dip=0.15Å) distortion. The out-of-plane distortion is dominated by the saddling mode with small contribution of waving (see SI 4 for NSD details). Out-of-plane distortions in porphyrins are known to be associated with enhancement of non-radiative relaxation pathways,80–82 however in π-extended porphyrins this effect is less pronounced.79 The fact that Pt-13 is only moderately saddled (for comparison, in tetraaryltetrabenzoporphyrins Doop values can be as high as 3.0-3.2Å) suggests that its emission quantum yield should decrease only slightly compared to that of the presumably completely planar counterpart Pt-12.
The structures of Pt-12 and other DAPIP complexes were not determined experimentally, but computed using DFT (see Experimental for details). The calculations, which were performed initially with B3LYP functional and LanL2DZ basis set, as implemented in Gaussian 16,83 revealed completely planar geometries for all DAPIP complexes (Pd-12,13 and Pt-12,13). Nearly identical planar structures were obtained with two other functionals, CAM-B3LYP and wB97XD, showing that DFT fails to reproduce the non-planarity of the macrocycle seen in the experimental structure. On the other hand, experimental X-ray structures of many other non-planar porphyrins, including π-extended porphyrins,79,84 have been well-reproduced by DFT. Our present result suggests that the non-planarity in the crystal structure of Pt-13 may be in fact a result of crystal packing forces, while in solution the average structure might be almost planar, as predicted by DFT.
Photophysical properties.
In this study we were particularly interested in the photophysics of the phosphorescent Pt and Pd complexes of DAPIP, for which we measured their linear (1P) absorption spectra, phosphorescence spectra along with phosphorescence decay times and quantum yields as well as the 2PA spectra (see below). The photophysical properties of meso-unsubstituted free-base DAPIP (H2-12) and its Zn complex (Zn-12) were also examined. All of the synthesized porphyrins are well-soluble in organic solvents (e.g. CH2Cl2, THF, DMF), showing no sign of aggregation at the concentrations required for optical measurements. The photophysical data for the studied compounds are summarized in Table 1, and the absorption and phosphorescence spectra of Pt and Pd DAPIP are shown in Fig. 2.
Table 1.
Selected photophysical data for the studied porphyrins.a.
Compd. | λmax B(nm) | λmax Q(nm) | B ε (M−1cm−1) σ(2) (GM) | Q ε (M−1cm−1) | σ(2)max (GM)b | λmax(em)c (nm) | ɸd | τe(s) |
---|---|---|---|---|---|---|---|---|
Pd-12 | 430 | 578 | 166540 33.4 |
83520 | 118.4 | 696 (p) 583 (df) |
0.14 0.007 |
3.8×10−4 |
Pt-12 | 415 | 566 | 127950 74.9 |
110490 | 234.7 | 675 (p) | 0.46 | 7×10−5 |
Zn-12 | 457 | 611 | 178280 | 47270 | - | 616 (f) | 0.11 | 2.3×10−9 |
H2-12 | 439 | 647 | - | - | - | 647 (f) | 0.16 | 1.1×10−8 |
Pd-13 | 441 | 586 | 232350 24.9 |
63850 | 138.3 | 703 (p) 592 (df) |
0.12 0.008 |
3.6×10−4 |
Pt-13 | 429 | 573 | 164940 16.8 |
84600 | 94.8 | 683 (p) | 0.43 | 6.1×10−5 |
All measurements were performed in DMF at 22° C. Phosphorescence quantum yields and decay times were measured in deoxygenated solutions.
Maximal 2PA cross-section measured at the blue edge of the spectrum (see Fig. 4 for details).
em - emission type: f - prompt fluorescence, p - phosphorescence, df- thermally activated (E-type) delayed fluorescence.
ɸ - emission quantum yield
τ - emission decay time.
Figure 2.
Absorption spectra of Pd (a) and Pt (c) complexes of porphyrins 12 and 13 in DMF. Emission spectra of Pd (b) and Pt (d) complexes of porphyrins 12 and 13 in deoxygenated DMF: p - phosphorescence; df - delayed fluorescence. The emission spectra are scaled by the respective phosphorescence quantum yields (Table 1).
Generally, the spectroscopic features of DAPIP resemble those of syn-dibenzo and syn-dinaphthoporphyrins48,68 as well as of fully symmetric tetrabenzo- (TBP), tetranaphtho- (TNP) and tetraarylphthalimidoporphyrins (TAPIP).69,70,79,84 The absorption spectra of DAPIP in the UV-visible range consist of two major bands, Q and B (Soret) bands, which are formed by the configuration interaction of single-electron excitations between two HOMO’S and two nearly degenerate LUMO’s.85 The Q and B bands comprise pairs of orthogonally polarized transitions, Qx, Qy and Bx, By, which in fully symmetric D4h porphyrins are pair-wise degenerate. In DAPIP, which have lower symmetry, the degeneracy is expected to be lifted. Indeed, TDDFT calculations predict existence of non-degenerate transitions, however the energies of the individual lines are too close to be resolvable in solution spectra at room temperature.
The TDDFT spectra and the frontier orbitals of Pd-12 are shown in Fig. 3 as an example (Fig. 3b). The summary of the TDDFT results is presented in Table 2. The data for Pd-13, Pt-12 and Pt-13 can be found in the SI (SI. 5). Here and in all calculations the pendant O(CH2)3CO2Et residues were substituted by OMe groups.
Figure 3.
(a) TDDFT (B3LYP/6-31+G(d,p) 1PA spectra of porphyrins Pd-12 and Pd-13. Transitions (vertical bars) are broadened by Lorentzian shapes (Γ=0.07 eV) to facilitate visual comparison with the experimental spectra (Fig. 2a). (b) The shapes and energies (eV) of the frontier Kohn-Sham orbitals in Pd-12. The symbols in the parentheses (a1u, a2u, eg) correspond to the symmetry types of the analogous MO’s in D4h-symmetric porphyrins. The blue arrows depict the single-electron excitations, whose linear combinations dominate the lowest energy Q and B transitions (see Table 2). The polarization axes for these transitions are shown on the image of LUMO1.
Table 2.
TDDFT energies, oscillator strengths and contributions of single-electron excitations to the lowest excited states in porphyrin Pd-12.a
Stateb | Energy (eV) | fc | Excitationd | Coefficientd |
---|---|---|---|---|
Qx | 2.3197 | 0.1261 | 220→223 | 0.3945 |
221→222 | 0.5809 | |||
Qy | 2.3309 | 0.0842 | 220→222 | −0.4069 |
221→223 | 0.5707 | |||
Bx | 2.9381 | 0.6906 | 215→222 | 0.1551 |
220→223 | 0.4676 | |||
221→222 | −0.3395 | |||
221→224 | −0.3302 | |||
By | 3.0038 | 0.6407 | 220→222 | 0.53321 |
220→224 | 0.19836 | |||
221→223 | 0.37291 | |||
221→225 | 0.14758 |
Calculations were performed using TDDFT at the B3LYP//6-31+G(d,p) level.
Transitions are designated by symbols x and y according to the polarizations axes shown in Fig. 3b.
Oscillator strength.
Single-electron excitations and the corresponding coefficients that give rise to the Q and B transitions via configuration interaction. Note that in TDDFT the normalization condition for the case of restricted reference reads: ‹A+B|A-B›=1/2, where A and B are the vectors of the coefficients for the excitation and de-excitation determinants, respectively.
The main effect of the addition of the phthalimide residues to the two adjacent pyrrolic units is the destabilization of the HOMO whose density is localized on the β-pyrrolic carbons (MO 221), relative to HOMO-1 (MO 220), which resides primarily on the meso-carbons and nitrogen atoms. In regular non-extended porphyrins these two orbitals belong to the a1u and a2u symmetry types and are quasi-degenerate, while the eg LUMOs are degenerate.85–87 As in other π-extended porphyrins,84,88 π-extension in DAPIP leads to a substantial red shift of the absorption bands and an increase in the oscillator strength of the Q band vs B band via intensity borrowing.85 The molar extinction coefficients at the peaks of the Q-bands of Pt and Pd DAPIP are nearly as high as those in the analogous TBPs,68 while the transition energies fall between those of the TBP and regular non-extended porphyrins. From the applications perspective, having sharp and strong transitions non-overlapping with transitions of TBPs may be advantageous for construction of sensor systems with orthogonal excitation of a pair of chromophores.
While the Qx and Qy transitions (designated according to their polarization along the axes shown in Fig. 3b) are dominated by the excitations between the frontier orbitals, Bx and By bands have non-negligible contributions from both lower and higher MOs. Similar situation was seen in symmetric TAPIP69 as well as in other large π-extended porphyrins (e.g. TNP84), revealing that the four-orbital approximation breaks down as the π-extended system increases in size.
Addition of two aryl groups to the 5,15-meso-positions in DAPIP destabilizes HOMO-1 (Table S2), which has its density on the meso-carbons (Fig. 3b), leading to further red shifts in the absorption of 13 vs 12 (Figs. 2,3a). This effect of meso-substitution is known.89 The same MO is engaged in the interaction with the central ion, which makes the spectra sensitive to the nature of the metal.85 Thus, the spectra of Pd complexes are red-shifted relative to those of their Pt counterparts, and the energy separation between the Bx and By transitions in PtDAPIP is larger, making the apparent B bands look broader.
Due to the presence of the heavy metal ions, the S1→T1 intersystem crossing in both Pt and Pd DAPIP is extremely efficient. None of these complexes exhibit prompt fluorescence. In contrast, free-base and ZnDAPIP (H2-12 and Zn-12) are fluorescent (SI.6), although their emission quantum yields are only moderate. Moderate fluorescence suggests that the intersystem crossing in both H2-12 and Zn-12 still efficiently competes with the radiative decay of the singlet state, but not nearly as successfully as in Pt and Pd porphyrins.
Pd and Pt complexes of both 12 and 13 emit remarkably bright phosphorescence at ambient temperatures (Fig. 2b,d). The phosphorescence quantum yields are record high (e.g. ɸp=0.46 for Pt-12), and the emissivity is practically unaffected by presence of the meso-aryl group (ɸp=0.43 for Pt-13), as expected from the very small degree of distortion seen in Pt-13 (see above). The ability to include a meso-aryl substituent without jeopardizing emissivity is very useful from the practical viewpoint, as it makes DAPIP much more amendable to synthetic modifications, e.g. attachment of linkers, inclusion into multichromophoric assemblies etc. Similar to other Pd and Pt porphyrins,75 the phosphorescence quantum yields and emission rates are considerably higher for Pt complexes of DAPIP. Pt(II), being a heavier ion, induces stronger spin-orbit coupling and promotes faster T1→S0 decay with favorable for phosphorescence distribution between the radiative and non-radiative decay channels.
An interesting feature of the Pd complexes (Pd-12, 13) is the presence of a relatively strong thermally activated delayed fluorescence, which is emitted in parallel with phosphorescence (Fig. 2b) and has the same decay time. Such E-type delayed fluorescence has been seen previously in PdTAPIP as well as in some Pd tetrabenzoporphyrins.69,90 Thermal exchange between the triplet and singlet manifolds is consistent with the narrow singlet-triplet energy gap (2J) due to the increase of the π-system. Thermal sensitivity of delayed fluorescence may be useful in temperature sensing applications.90
Two-photon absorption (2PA).
syn-Fusion of the macrocycle with phthalimide groups leads not only to the destabilization of one of the HOMOs (MO 221 in Fig. 3b) relative to the other (MO 220), but also decreases its symmetry. In symmetric D4h porphyrins, such as regular porphyrins, TBP, TAPIP etc, the corresponding HOMO is of a1u symmetry type, while in DAPIP HOMOs and LUMOs are neither symmetric (g) nor antisymmetric (u) with respect to the center of inversion of symmetry. Since the Q and B states comprise linear combinations of excitations between the frontier MOs, they cannot be classified as u or g states either. Consequently, Q and B transitions, which are 2PA-forbidden in symmetric porphyrins, should become allowed in DAPIP.
The experimental 2PA spectra of the studied Pd and Pt DAPIP are shown in Fig. 4. The spectra were measured by the method of 2P-excited time-resolved phosphorescence in time domain,68 which allows complete suppression of the scattered excitation and other sources of noise. At every wavelength the power dependence, expressed as a log/log plot of the phosphorescence intensity vs excitation flux, was confirmed to be quadratic (slope 2.0±0.05). The blue edge (750 nm for all porphyrins) was reached when the slope became less than 1.95. Below that limit (λex<750 nm) 2PA was contaminated by 1PA into the lower vibronic components of the Q state and/or by direct spin-forbidden S0→T1 excitation, which becomes partially allowed in porphyrins with strong spin-orbit coupling.61 The red edge of the 2PA spectrum (1080 nm) was set by the tunability range of the excitation laser.
Figure 4.
2PA spectra (blue lines) of Pd and Pt DAPIP in DMF at 22°C measured by the method of 2P phosphorescence excitation. The corresponding 1PA spectra (black lines) are arbitrarily scaled, but the relationship between the spectral intensities (molar extinction coefficients) between the compounds is kept unchanged to facilitate visual comparison. The 2PA spectra are truncated at the blue edge (750 nm; shown on the spectrum of Pd-12) where the slope of the log/log plot, representing the dependence on the signal on the excitation flux, becomes <1.95. The red edge (1080 nm) was determined by the tunability range of the laser.
To interpret the experimental findings we carried out calculations of 2PA using the TDDFT/Sum-Over-States (SOS) methodology, as described previously68,69,91 (see Experimental and SI. 7 for details). The calculated 2PA spectra of porphyrins Pd-12 and Pd-13 are shown in Fig. 5. The spectra of the Pt complexes can be found in SI. 8.
Figure 5.
(a) 2PA spectra (blue bars) of porphyrins Pd-12 and Pd-13 computed using TDDFT/Sum-Over-States (SOS) methodology. The spectra are broadened by Lorentzian shapes (blue line, Γ=0.1 eV). Proportionally scaled 1PA spectra are shown for comparison (black line), (b) Leading single-electron excitations comprising 72% and 85% of the transition amplitudes for the lowest 2P-active states in PdDAPIP (marked with asterisk in the spectra) and PdTAPIP, respectively, along with the respective orbitals.
First we note that the calculations (Fig. 5) reproduce the experimental results (Fig. 4) quite well. The computed values of the 2PA cross-sections are 2-3 times higher than those determined experimentally. However, this disagreement can be viewed as minor, considering that experimental cross-sections routinely vary by orders of magnitude depending on the measurement method. More importantly, qualitatively all the experimental features are reproduced by the calculations correctly.
It can be seen (Fig. 4) that in all cases 2PA for the B state is indeed non-negligible, as expected from the symmetry breaking by the syn-diphthalimido-extension. The 2PA cross-sections for the B states in the studied DAPIP are higher than in the similarly syn-π-extended porphyrins with ester groups.68 However, it is also clear that much stronger 2P-active states in DAPIP are positioned above the B state level, as evidenced by the steep rise of 2PA at the blue edge.
2PA for a transition from the ground state (0) to a state designated as final (f) comprises coherent action of multiple excitation channels, whereby each channel contributes to the total value of the 2P transition strength δ(2) as specified by the SOS expression:
(1) |
where ħω is the photon energy, i runs over all of the energy eigenstates of the system, μnm, are transition dipole moments connecting states n and m, and Ei, are the states’ energies. 2PA cross-section σ(2) is proportional to the rotationally averaged δ(2).92–95 Expression (1) is written for degenerate 2PA, where the two photons have the same energy and the same linear polarization, and a simplified case when the molecule is fixed and all transitions are polarized in the same direction. Under these conditions µij in (1) are simply the projections of the transition dipole moments on the photon polarization axis. For an ensemble of randomly oriented molecules and photons with different energies and polarizations the expression for δ(2) becomes more bulky93 (see SI.7), however the qualitative relationship between the cross-section value and the electronic structure of the molecule remains the same.
Each term in (1) represents an excitation channel, 0→i→f where i denotes an intermediate state on the pathway between the ground and the final state. The contribution of each channel is weighted by the energy difference, Ei-ħω, known as the detuning factor. There are two unique channels, 0→0→f and 0→f→f for which the corresponding terms in (1) contain products μ00μ0f and μ0fμff. The values μ00 and μff, or simply μ0 and μf, are the diagonal elements of the electric dipole operator, i.e. the static (or permanent) dipole moments in the ground and final states. The combined contribution of these two channels, Σ0+Σf, is expressed in (1) as , i.e. it is proportional to the difference between the static dipole moments μf and μ0. In non-centrosymmetric molecules the change in the static polarization upon excitation is generally non-zero, and it is common to approximate 2PA to the lowest excited state (f) in such dyes using only the states 0 and f - an approximation known as the two-state model.95 However, as we show below, for DAPIP this approach is not valid.
Using (1) it is straightforward to define the contribution of all channels encompassing a certain intermediate estate i, or simply the contribution Σi, of state i, to the total transition strength δ(2) (see SI.7).91 The contributions of states Q and B and of the lowest 2P-active state Δ (marked with asterisk in Fig. 5) for PdDAPIP are presented in Table 3 as an example. The contributions of the analogous states in D4h-symmetric PdTAPIP are shown for comparison. In the latter case, the only state with non-zero 2PA strength is the g-state Δ, which in PdTAPIP lies below the B state level.69 Excitation of the Δ state is distributed between four channels involving u-states Qx, Qy and Bx, By, while these states themselves are parity-forbidden for 2PA.
Table 3.
Calculated 2PA cross-sections (σ(2)) and fractional contributions Σi, of the excitation channels 0→i→f encompassing states 0 (ground state), Q, B and Δ (the lowest 2P-active state) as intermediate states (i) to the total 2P transition strength δ(2) in PdDAPIP.a Contributions of the analogous states in a D4h-symmetric PdTAPIP are shown for comparison.
Final state (f) | σ(2) (GM) |
Σf+Σ0b | ΣQx | ΣQy | ΣBx | ΣBy | ΣΔ |
---|---|---|---|---|---|---|---|
Qx | 11.5 | 0.22 | - | 0.13 | 0.36 | 0.46 | −0.26 |
Qy | 41.1 | 0.06 | 0.15 | - | 0.88 | −0.04 | −0.19 |
Bx | 144.0 | 0.39 | 0.15 | 0.32 | - | 0.00 | −0.07 |
By | 112.5 | 0.65 | 0.21 | −0.08 | 0.00 | - | −0.14 |
Δc | 578.7 | −0.04 | 0.30 | 0.20 | 0.17 | 0.33 | - |
ΔpdTAPIPd | 1944.0 | 0.00 | 0.33 | 0.33 | 0.17 | 0.17 | 0.00 |
The electronic structure calculations were performed using TDDFT (B3LYP//6-31G(d,p)) as implemented in Firefly 8.2.096 on the DFT-optimized structures.
The sums of the contributions of channels 0→0→f and 0→f→f representing the term in the SOS. μ0 and μf are the static dipole moments in states 0 and f respectively.
Δ designates the lowest 2P-active state (marked with asterisk in Fig. 5a).
Lowest 2P-active gerade state Δ in D4h-symmetric PdTAPIP. The Q and B states in PdTAPIP (ungerade states) possess zero 2PA activity.
Deviation from D4h in DAPIP enables all of its states to gain some 2PA strength. In each case, the distribution of the transition amplitude over the channels is governed by the orientation of the moments μ0i, and μif and the respective detuning factors. It can be seen (Table 3) that for all states an appreciable fraction of 2PA is due to the static dipoles terms (Σf+Σ0), reflecting the fact that DAPIP molecule is polar, having its static moments oriented along the y-axis. Nevertheless, in spite of the absence of the center of inversion symmetry, the overall picture of 2PA in PdDAPIP still resembles that in TAPIP and other centrosymmetric porphyrins, indicating that the symmetry breaking, even as significant as by syn-phthalimido-fusion, induces only a relatively minor perturbation of the overall highly symmetric electronic structure of the porphyrin. For example, the maximal 2PA in PdDAPIP is predicted for state Δ, whose orbital origin is essentially the same as that in fully-symmetric PdTAPIP (Fig. 5b). The excitation of this state is also distributed between the Q and B channels, although not as evenly as in PdTAPIP. In both DAPIP and TAPIP the 2P-active states Δ are mixtures of multiple orbital configurations, but the leading terms are due to the respective HOMO’S and LUMO+2 (Fig. 5b), where the electron density shifts from the center of the macrocycle out to the peripheral phthalimide groups. It appears that such a delocalization more evenly over a larger molecular framework leads to narrowing of the HOMO-LUMO+2 gap, stabilizing the Δ-state and increasing 2PA.
The B and Q states in DAPIP are much less 2P-active than Δ, however the static dipole terms in each case are supplemented by large contributions of the complementary Q and B channels. The computed transition strengths are different for Bx and By, which is consistent with the experimentally measured spectral shapes (Fig. 4): the 2PA peaks in the B state region are slightly red-shifted relative to the respective 1PA maxima. Also noteworthy is the fact that for all states the channels involving state Δ interfere destructively with other excitation pathways, which is reflected by the negative signs of the respective contributions. Negative interference shows that state Δ effectively acts to diminish the 2PA strength for the Q and B states, while gaining its own intensity through the constructive interference of the channels enabled by these states. This phenomenon resembles the B-to-Q intensity borrowing effect, described by Gouterman,85 where lifting of the degeneracy between the two porphyrin HOMOs leads to redistribution of intensity between the Q and B bands. A similar effect of Δ state was observed recently in the case of three-photon absorption in π-extended porphyrins.91
Conclusions
In this contribution we developed an efficient synthesis of syn-diphthalimidoporphyrins (DAPIP) and performed their structural and photophysical characterization, including measurements of their 2PA spectra. Using quantum-chemical analysis based on the TDDFT/SOS approach we were able to quantify the contributions of individual excitation channels to 2PA in DAPIP and trace the effects of symmetry breaking to the orbital origin of these channels. From the practical viewpoint, the combination of symmetry breaking with the stabilization of the lowest 2P-active state (Δ) significantly enhances 2PA of DAPIP in the B-state region. The exceptionally bright phosphorescence of the Pd and Pt complexes of DAPIP makes them valuable as luminescent markers for biological applications. The ability to concurrently excite DAPIP and their fully-symmetric analogs TAPIP, while differentially sampling their emission may be utilized in the construction of orthogonal optical sensing systems for advanced imaging applications.
Supplementary Material
Acknowledgements
Support of the grants R21EB027397 and U24EB028941 from the National Institutes of Health, USA is gratefully acknowledged. The authors are thankful to Dr. Michael R. Gau for X-ray structure determination. Computations were performed using the computational resource at the Department of Biochemistry of Biophysics (Penn), funded by the grant S10-OD023592 from the NIH USA and by the Johnson Research Foundation.
Footnotes
Supporting Information for Publication.
Description of the synthesis, experimental details, optical spectroscopic data, results of calculations.
References
- (1).Senge MO; Fazekas M; Notaras EGA; Blau WJ; Zawadzka M; Locos OB; Ni Mhuircheartaigh EM, Nonlinear optical properties of porphyrins. Adv. Mater 2007,19 (19), 2737–2774. [Google Scholar]
- (2).Calvete M; Yang GY; Hanack M, Porphyrins and phthalocyanines as materials for optical limiting. Synthetic Metals 2004,141 (3), 231–243. [Google Scholar]
- (3).Ye C; Zhou L; Wang X; Liang Z, Photon upconversion: from two-photon absorption (TPA) to triplet–triplet annihilation (TTA). Phys. Chem. Chem. Phys 2016, 18(16), 10818–10835. [DOI] [PubMed] [Google Scholar]
- (4).Ogawa K; Kobuke Y, Design of two-photon absorbing materials for molecular optical memory and photodynamic therapy. Org. & Biomol. Chem 2009, 7 (11), 2241–2246. [DOI] [PubMed] [Google Scholar]
- (5).Mathai S; Bird DK; Stylli SS; Smith TA; Ghiggino KP, Two-photon absorption cross-sections and time-resolved fluorescence imaging using porphyrin photosensitisers. Photochem. & Photobiol. Sci 2007, 6 (9), 1019–1026. [DOI] [PubMed] [Google Scholar]
- (6).Finikova OS; Lebedev AY; Aprelev A; Troxler T; Gao F; Garnacho C; Muro S; Hochstrasser RM; Vinogradov SA, Oxygen microscopy by two-photon-excited phosphorescence. ChemPhysChem 2008, 9 (12), 1673–1679. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Kuimova MK; Collins HA; Balaz M; Dahlstedt E; Levitt JA; Sergent N; Suhling K; Drobizhev M; Makarov NS; Rebane A; Anderson HL; Phillips D, Photophysical properties and intracellular imaging of water-soluble porphyrin dimers for two-photon excited photodynamic therapy. Org. & Biomol. Chem 2009, 7 (5), 889–896. [DOI] [PubMed] [Google Scholar]
- (8).Chen Y; Guan R; Zhang C; Huang J; Ji L; Chao H, Two-photon luminescent metal complexes for bioimaging and cancer phototherapy. Coord. Chem. Rev 2016, 310, 16–40. [Google Scholar]
- (9).Khadria A; Fleischhauer J; Boczarow T; Wilkinson JD; Kohl MM; Anderson HL, Porphyrin dyes for nonlinear optical imaging of live cells. iScience 2018, 4, 153–163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (10).Esipova TV; Barrett MJP; Erlebach E; Masunov AE; Weber B; Vinogradov SA, Oxyphor 2P: A high-performance probe for deep-tissue longitudinal oxygen imaging. Cell Metabolism 2019, 29 (3), 736–744. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (11).Karotki A; Kruk M; Drobizhev M; Rebane A; Nickel E; Spangler CW, Efficient singlet oxygen generation upon two-photon excitation of new porphyrin with enhanced nonlinear absorption. IEEE J. Sel. Top. Quantum Electron 2001, 7 (6), 971–975. [Google Scholar]
- (12).Schmitt J; Heitz V; Sour A; Bolze F; Ftouni H; Nicoud J-F; Flamigni L; Ventura B, Diketopyrrolopyrrole-porphyrin conjugates with high two-photon absorption and singlet oxygen generation for two-photon photodynamic therapy. Angew. Chem. Int. Ed 2015, 54 (1), 169–173. [DOI] [PubMed] [Google Scholar]
- (13).Collins HA; Khurana M; Moriyama EH; Mariampillai A; Dahlstedt E; Balaz M; Kuimova MK; Drobizhev M; Yang VXD; Phillips D; Rebane A; Wilson BC; Anderson HL, Blood-vessel closure using photosensitizers engineered for two-photon excitation. Nature Photonics 2008, 2 (7), 420–424. [Google Scholar]
- (14).Jenni S; Sour A; Bolze F; Ventura B; Heitz V, Tumour-targeting photosensitisers for one- and two-photon activated photodynamic therapy. Org. & Biomol. Chem 2019, 17 (27), 6585–6594. [DOI] [PubMed] [Google Scholar]
- (15).Shi L; Nguyen C; Daurat M; Dhieb AC; Smirani W; Blanchard-Desce M; Gary-Bobo M; Mongin O; Paul-Roth C; Paul F, Biocompatible conjugated fluorenylporphyrins for two-photon photodynamic therapy and fluorescence imaging. Chem. Commun 2019, 55 (81), 12231–12234. [DOI] [PubMed] [Google Scholar]
- (16).Kim KS; Lim JM; Osuka A; Kim D, Various strategies for highly-efficient two-photon absorption in porphyrin arrays. J. Photochem. Photobiol. C: Photochem. Rev 2008, 9 (1), 13–28. [Google Scholar]
- (17).Pawlicki M; Collins HA; Denning RG; Anderson HL, Two-photon absorption and the design of two-photon dyes. Angew. Chem. Int. Ed 2009, 48 (18), 3244–3266. [DOI] [PubMed] [Google Scholar]
- (18).Myung Kim H; Rae Cho B, Two-photon materials with large two-photon cross sections. structure–property relationship. Chem. Commun 2009, (2), 153–164. [DOI] [PubMed] [Google Scholar]
- (19).Tanaka T; Osuka A, Conjugated porphyrin arrays: synthesis, properties and applications for functional materials. Chem. Soc. Rev 2015, 44 (4), 943–969. [DOI] [PubMed] [Google Scholar]
- (20).Alam MM; Bolze F; Daniel C; Flamigni L; Gourlaouen C; Heitz V; Jenni S; Schmitt J; Sour A; Ventura B, π-Extended diketopyrrolopyrrole–porphyrin arrays: one- and two-photon photophysical investigations and theoretical studies. Phys. Chem. Chem. Phys 2016, 18 (31), 21954–21965. [DOI] [PubMed] [Google Scholar]
- (21).Mikhaylov A; Kondratuk DV; Cnossen A; Anderson HL; Drobizhev M; Rebane A, Cooperative enhancement of two-photon absorption in self-assembled Zinc-porphyrin nanostructures. J. Phys. Chem. C 2016,120 (21), 11663–11670. [Google Scholar]
- (22).Lewtak JP; Gryko DT, Synthesis of π-extended porphyrins via intramolecular oxidative coupling. Chem. Commun 2012, 48 (81), 10069–10086. [DOI] [PubMed] [Google Scholar]
- (23).Ogawa K; Ohashi A; Kobuke Y; Kamada K; Ohta K, Strong two-photon absorption of self-assembled butadiyne-linked bisporphyrin. J. Am. Chem. Soc 2003,125 (12), 13356–13357. [DOI] [PubMed] [Google Scholar]
- (24).Lash TD Synthesis of novel porphyrinoid chromophores. In The Porphyrin Handbook; Kadish KMS, K. M; Guilard R, Ed.; Academic Press: New York, 2000; pp 125–199. [Google Scholar]
- (25).Cheprakov AV The synthesis of π-extended porphyrins. In Handbook of porphyrin science; Kadish KM, Smith KM, Guilard R, Eds.; World Scientific Publishing Co.: Singapoure, 2011; Vol. 13; pp 1–150. [Google Scholar]
- (26).Galanin NE; Shaposhnikov GP; Koifman OI, Methods for synthesis of meso-substituted tetrabenzoporphyrins. Russ. Chem. Rev 2013, 82 (5), 412–428. [Google Scholar]
- (27).Carvalho CMB; Brocksom TJ; de Oliveira KT, Tetrabenzoporphyrins: synthetic developments and applications. Chem. Soc. Rev 2013, 42 (8), 3302–3317. [DOI] [PubMed] [Google Scholar]
- (28).Vicente MGH; Tome AC; Walter A; Cavaleiro JAS, Synthesis and cycloaddition reactions of pyrrole-fused 3-sulfolenes: A new versatile route to tetrabenzoporphyrins. Tetrahedron Lett. 1997, 38 (20), 3639–3642. [Google Scholar]
- (29).Ito S; Murashima T; Uno Η; Ono N, A new synthesis of benzoporphyrins using 4,7-dihydro-4,7- ethano-2H-isoindole as a synthon of isoindole. Chem. Commun 1998, (16), 1661–1662. [Google Scholar]
- (30).Ito S; Ochi N; Uno H; Murashima T; Ono N, A new synthesis of [2,3]naphthoporphyrins. Chem. Commun 2000, (11), 893–894. [Google Scholar]
- (31).Lash TD, Modification of the porphyrin chromophore by ring fusion: identifying trends due to annelation of the porphyrin nucleus. J. Porph. Phthalocyan 2001, 5 (3), 267–288. [Google Scholar]
- (32).Finikova OS; Cheprakov AV; Beletskaya IP; Carroll PJ; Vinogradov SA, Novel versatile synthesis of substituted tetrabenzoporphyrins. J. Org. Chem 2004, 69 (2), 522–535. [DOI] [PubMed] [Google Scholar]
- (33).Finikova OS; Aleshchenkov SE; Brinas RP; Cheprakov AV; Carroll PJ; Vinogradov SA, Synthesis of symmetrical tetraaryltetranaphtho[2,3]porphyrins. J. Org. Chem 2005, 70 (12), 4617–4628. [DOI] [PubMed] [Google Scholar]
- (34).Cheprakov AV; Filatov MA, The dihydroisoindole approach to linearly annelated π-extended porphyrins. J. Porph. Phthalocyan 2009,13 (3), 291–303. [Google Scholar]
- (35).Filatov MA; Cheprakov AV, The synthesis of new tetrabenzo- and tetranaphthoporphyrins via the addition reactions of 4,7-dihydroisoindole. Tetrahedron 2011, 67 (19), 3559–3566. [Google Scholar]
- (36).Jiang L; Zaenglein RA; Engle JT; Mittal C; Hartley CS; Ziegler CJ; Wang H, Water-soluble ionic benzoporphyrins. Chem. Commun 2012, 48 (55), 6927–6929. [DOI] [PubMed] [Google Scholar]
- (37).Kumar S; Jiang X; Shan W; Jinadasa RGW; Kadish KM; Wang H, β-Functionalized trans-A2B2 push–pull tetrabenzoporphyrins. Chem. Commun 2018, 54 (42), 5303–5306. [DOI] [PubMed] [Google Scholar]
- (38).Moss A; Zhou Z; Jiang L; Vicente MGH; Wang H, Synthesis of highly water soluble tetrabenzoporphyrins and their application toward photodynamic therapy. J. Porph. Phthalocyan 2020, 24 (01n03), 456–464. [Google Scholar]
- (39).Bonnett R; McManus KA, Opp-dibenzoporphyrins from benzopyrromethene derivatives. J. Chem. Soc., Chem. Commun 1994, (9), 1129–1130. [Google Scholar]
- (40).Lash TD; Roper TJ, Synthesis of dinaphthoporphyrins from dihydronaphtho[1,2-c]pyrroles. Tetrahedron Lett. 1994, 35 (42), 7715–7718. [Google Scholar]
- (41).Lash TD; Novak BH, New highly conjugated porphyrin chromophores: Synthesis of mono- and diphenanthroporphyrins. Tetrahedron Lett. 1995, 36 (25), 4381–4384. [Google Scholar]
- (42).Lee SH; Smith KM, Sulfolenoporphyrins: synthons for refunctionalization of porphyrins. Tetrahedron Lett. 2005, 46 (12), 2009–2013. [Google Scholar]
- (43).Manley JM; Roper TJ; Lash TD, Synthesis of isomeric angularly annealed dinaphthoporphyrin systems: examination of the relative positioning and orientation of ring fusion as factors influencing the porphyrin chromophore. J. Org. Chem 2005, 70 (3), 874–891. [DOI] [PubMed] [Google Scholar]
- (44).Silva AMG; de Oliveira KT; Faustino MAF; Neves MGPMS; Tomé AC; Silva AMS; Cavaleiro JAS; Brandão P; Felix V, Chemical transformations of mono- and bis(buta-l,3-dien-1-yl)porphyrins: a new synthetic approach to mono- and dibenzoporphyrins. Eur. J. Org. Chem 2008, 2008 (4), 704–712. [Google Scholar]
- (45).Gandhi V; Thompson ML; Lash TD, Porphyrins with exocyclic rings. Part 24. Synthesis and spectroscopic properties of pyrenoporphyrins, potential building blocks for porphyrin molecular wires. Tetrahedron 2010, 66 (10), 1787–1799. [Google Scholar]
- (46).Uoyama H; Takiue T; Tominaga K; Ono N; Uno H, Synthesis, structures, and properties of BCOD-fused porphyrins and benzoporphyrins. J. Porph. Phthalocyan 2009,13 (01), 122–135. [Google Scholar]
- (47).Kim P; Sung J; Uoyama H; Okujima T; Uno H; Kim D, Ultrafast Intramolecular energy relaxation dynamics of benzoporphyrins: influence of fused benzo rings on singlet excited states. J. Phys. Chem. B 2011, 115 (14), 3784–3792. [DOI] [PubMed] [Google Scholar]
- (48).Esipova TV; Vinogradov SA, Synthesis of phosphorescent asymmetrically π-extended porphyrins for two-photon applications. J. Org. Chem 2014, 79 (18), 8812–8825. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (49).Lebedev AY; Cheprakov AV; Sakadzic S; Boas DA; Wilson DF; Vinogradov SA, Dendritic phosphorescent probes for oxygen imaging in biological systems. ACS Appl. Materials & Interfaces 2009, 1 (6), 1292–1304. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Vinogradov SA; Wilson DF Porphyrin-dendrimers as biological oxygen sensors. In Designing Dendrimers; Capagna S, Ceroni P, Eds.; Wiley: New York, 2012. [Google Scholar]
- (51).Brinas RP; Troxler T; Hochstrasser RM; Vinogradov SA, Phosphorescent oxygen sensor with dendritic protection and two-photon absorbing antenna. J. Am. Chem. Soc 2005, 127 (33), 11851–11862. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (52).Roussakis E; Spencer JA; Lin C; Vinogradov SA, Two-photon antenna-core oxygen probe with enhanced performance. Anal. Chem 2014, 86, 5937–5945. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (53).Sakadzic S; Roussakis E; Yaseen MA; Mandeville ET; Srinivasan VJ; Arai K; Ruvinskaya S; Devor A; Lo EH; Vinogradov SA; Boas DA, Two-photon high-resolution measurement of partial pressure of oxygen in cerebral vasculature and tissue. Nature Methods 2010, 7 (9), 755–U125. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (54).Lecoq J; Parpaleix A; Roussakis E; Ducros M; Houssen YG; Vinogradov SA; Charpak S, Simultaneous two-photon imaging of oxygen and blood flow in deep cerebral vessels. Nature Medicine 2011, 17 (7), 893–U262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (55).Sencan I; Esipova T; Kilic K; Li BQ; Desjardins M; Yaseen MA; Wang H; Porter JE; Kura S; Fu BY; Secomb TW; Boas DA; Vinogradov SA; Devor A; Sakadzic S, Optical measurement of microvascular oxygenation and blood flow responses in awake mouse cortex during functional activation. JCBFM 2020, Article Number: 0271678X20928011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (56).Parpaleix A; Houssen YG; Charpak S, Imaging local neuronal activity by monitoring pO2 transients in capillaries. Nature Medicine 2013, 19 (2), 241–246. [DOI] [PubMed] [Google Scholar]
- (57).Spencer JA; Ferraro F; Roussakis E; Klein A; Wu JW; Runnels JM; Zaher W; Mortensen LJ; Alt C; Turcotte R; Yusuf R; Cote D; Vinogradov SA; Scadden DT; Lin CP, Direct measurement of local oxygen concentration in the bone marrow of live animals. Nature 2014, 508 (7495), 269–273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (58).Christodoulou C; Spencer JA; Yeh SCA; Turcotte R; Kokkaliaris KD; Panero R; Ramos A; Guo GJ; Seyedhassantehrani N; Esipova TV; Vinogradov SA; Rudzinskas S; Zhang Y; Perkins AS; Orkin SH; Calogero RA; Schroeder T; Lin CP; Camargo FD, Live-animal imaging of native haematopoietic stem and progenitor cells. Nature 2020, 578 (7794), 278–283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (59).Schilling K; El Khatib M; Plunkett S; Xue JJ; Xia YN; Vinogradov SA; Brown E; Zhang XP, Electrospun fiber mesh for high-resolution measurements of oxygen tension in cranial bone defect repair. ACS Appl. Mater. Interfaces 2019, 11 (37), 33548–33558. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (60).Rytelewski M; Haryutyunan K; Nwajei F; Shanmugasundaram M; Wspanialy P; Zal MA; Chen CH; El Khatib M; Plunkett S; Vinogradov SA; Konopleva M; Zal T, Merger of dynamic two-photon and phosphorescence lifetime microscopy reveals dependence of lymphocyte motility on oxygen in solid and hematological tumors. J. Immunother. Canc 2019, 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (61).Finikova OS; Troxler T; Senes A; DeGrado WF; Hochstrasser RM; Vinogradov SA, Energy and electron transfer in enhanced two-photon-absorbing systems with triplet cores. J. Phys. Chem. A 2007, 111 (30), 6977–6990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (62).Kim DY; Ahn TK; Kwon JH; Kim D; Ikeue T; Aratani N; Osuka A; Shigeiwa M; Maeda S, Large two-photon absorption (TPA) cross-section of directly linked fused diporphyrins. J. Phys. Chem. A 2005, 109 (13), 2996–2999. [DOI] [PubMed] [Google Scholar]
- (63).Drobizhev M; Stepanenko Y; Dzenis Y; Karotki A; Rebane A; Taylor PN; Anderson HL, Extremely Strong near-IR two-photon absorption in conjugated porphyrin dimers: quantitative description with three-erssential-states model. J. Phys. Chem. B 2005, 109 (15), 7223–7236. [DOI] [PubMed] [Google Scholar]
- (64).Tanihara J; Ogawa K; Kobuke Y, Two-photon absorption properties of conjugated supramolecular porphyrins with electron donor and acceptor. J. Photochem. Photobiol. A: Chemistry 2006, 178 (2), 140–149. [Google Scholar]
- (65).Mongin O; Sankar M; Chariot M; Mir Y; Blanchard-Desce M, Strong enhancement of two-photon absorption properties in synergic ‘semi-disconnected’ multiporphyrin assemblies designed for combined imaging and photodynamic therapy. Tetrahedron Lett. 2013, 54 (48), 6474–6478. [Google Scholar]
- (66).Raymond JE; Bhaskar A; Goodson T; Makiuchi N; Ogawa K; Kobuke Y, Synthesis and two-photon absorption enhancement of porphyrin macrocycles. J. Am. Chem. Soc 2008, 130 (51), 17212–17213. [DOI] [PubMed] [Google Scholar]
- (67).Aratani N; Kim D; Osuka A, π-Conjugation enlargement toward the creation of multiporphyrinic systems with large two-photon absorption properties. Chem. Asian J 2009, 4 (8), 1172–1182. [DOI] [PubMed] [Google Scholar]
- (68).Esipova TV; Rivera-Jacquez HJ; Weber B; Masunov AE; Vinogradov SA, Two-photon absorbing phosphorescent metalloporphyrins: effects of π-extension and peripheral substitution. J. Am. Chem. Soc 2016, 138 (48), 15648–15662. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (69).Esipova TV; Rivera-Jacquez HJ; Weber B; Masunov AE; Vinogradov SA, Stabilizing g-states in centrosymmetric tetrapyrroles: two-photon-absorbing porphyrins with bright phosphorescence. J. Phys. Chem. A 2017, 121 (33), 6243–6255. [DOI] [PubMed] [Google Scholar]
- (70).Zems Y; Moiseev AG; Perepichka DF, Convenient synthesis of a highly soluble and stable phosphorescent platinum porphyrin dye. Organic Lett. 2013, 15 (20), 5330–5333. [DOI] [PubMed] [Google Scholar]
- (71).Gunter MJ; Tang H; Warrener RN, Establishing a library of porphyrin building blocks for superstructured assemblies: porphyrin dienes and dienophiles for cycloaddition reactions. J. Porph. Phthalocyan 2002, 06 (11), 673–684. [Google Scholar]
- (72).Filatov MA; Cheprakov AV; Beletskaya IP, A facile and reliable method for the synthesis of tetrabenzoporphyrin from 4,7-dihydroisoindole. Eur. J. Org. Chem 2007, 2007 (21), 3468–3475. [Google Scholar]
- (73).Taniguchi M; Balakumar A; Fan D; McDowell BE; Lindsey JS, Imine-substituted dipyrromethanes in the synthesis of porphyrins bearing one or two meso substituents. J. Porph. Phthalocyan 2005, 09 (08), 554–574. [Google Scholar]
- (74).Borbas KE; Kee HL; Holten D; Lindsey JS, A compact water-soluble porphyrin bearing an iodoacetamido bioconjugatable site. Org. & Biomol. Chem 2008, 6 (1), 187–194. [DOI] [PubMed] [Google Scholar]
- (75).Eastwood D; Gouterman M, Porphyrins. XVIII Luminescence of (Co), (Ni), Pd, Pt complexes. J. Mol. Spectros 1970, 35, 359–375. [Google Scholar]
- (76).Montes VA; Pérez-Bolívar C; Agarwal N; Shinar J; Anzenbacher P, Molecular-wire behavior of OLED materials: exciton dynamics in multichromophoric Alq3-oligofluorene-Pt(II)porphyrin triads. J. Am. Chem. Soc 2006, 128 (38), 12436–12438. [DOI] [PubMed] [Google Scholar]
- (77).Dean ML; Schmink JR; Leadbeater NE; Brückner C, Microwave-promoted insertion of Group 10 metals into free base porphyrins and chlorins: scope and limitations. Dalton Trans. 2008, (10), 1341–1345. [DOI] [PubMed] [Google Scholar]
- (78).Jentzen W; Song X-Z; Shelnutt JA, Structural characterization of synthetic and protein-bound porphyrins in terms of the lowest-frequency normal coordinates of the macrocycle. J. Phys. Chem. B 1997, 101 (9), 1684–1699. [Google Scholar]
- (79).Lebedev AY; Filatov MA; Cheprakov AV; Vinogradov SA, Effects of structural deformations on optical properties of tetrabenzoporphyrins: free-bases and Pd complexes. J. Phys. Chem. A 2008, 112 (33), 7723–7733. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (80).Gentemann S; Medforth CJ; Forsyth TP; Nurco DJ; Smith KM; Fajer J; Holten D, Photophysical properties of conformationally distorted metal-free porphyrins. Investigation into the deactivation mechanisms of the lowest excited singlet state. J. Am. Chem. Soc 1994, 116 (16), 7363–7368. [Google Scholar]
- (81).Chirvony VS; van Hoek A; Galievsky VA; Sazanovich IV; Schaafsma TJ; Holten D, Comparative study of the photophysical properties of nonplanar tetraphenylporphyrin and octaethylporphyrin diacids. J. Phys. Chem. B 2000, 104 (42), 9909–9917. [Google Scholar]
- (82).Sazanovich IV; Galievsky VA; van Hoek A; Schaafsma TJ; Malinovskii VL; Holten D; Chirvony VS, Photophysical and structural properties of saddle-shaped free base porphyrins: Evidence for an “orthogonal” dipole moment. J. Phys. Chem. B 2001, 105 (32), 7818–7829. [Google Scholar]
- (83).Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Petersson GA; Nakatsuji H; Li X; Caricato M; Marenich AV; Bloino J; Janesko BG; Gomperts R; Mennucci B; Hratchian HP; Ortiz JV; Izmaylov AF; Sonnenberg JL; Williams; Ding F; Lipparini F; Egidi F; Goings J; Peng B; Petrone A; Henderson T; Ranasinghe D; Zakrzewski VG; Gao J; Rega N; Zheng G; Liang W; Hada M; Ehara M; Toyota K; Fukuda R; Hasegawa J; Ishida M; Nakajima T; Honda Y; Kitao O; Nakai H; Vreven T; Throssell K; Montgomery JA Jr.; Peralta JE; Ogliaro F; Bearpark MJ; Heyd JJ; Brothers EN; Kudin KN; Staroverov VN; Keith TA; Kobayashi R; Normand J; Raghavachari K; Rendell AP; Burant JC; Iyengar SS; Tomasi J; Cossi M; Millam JM; Klene M; Adamo C; Cammi R; Ochterski JW; Martin RL; Morokuma K; Farkas O; Foresman JB; Fox DJ, Gaussian 16 Rev. C.01 2016. [Google Scholar]
- (84).Rogers JE; Nguyen KA; Hufnagle DC; McLean DG; Su WJ; Gossett KM; Burke AR; Vinogradov SA; Pachter R; Fleitz PA, Observation and interpretation of annulated porphyrins: studies on the photophysical properties of meso-tetraphenylmetalloporphyrins. J. Phys. Chem. A 2003, 107 (51), 11331–11339. [Google Scholar]
- (85).Gouterman M, Spectra of porphyrins. J. Mol. Spctros 1961, 6, 138–163. [Google Scholar]
- (86).Mandal AK; Taniguchi M; Diers JR; Niedzwiedzki DM; Kirmaier C; Lindsey JS; Bocian DF; Holten D, Photophysical properties and electronic structure of porphyrins bearing zero to four meso-phenyl substituents: new insights into seemingly well understood tetrapyrroles. J. Phys. Chem. A 2016, 120 (49), 9719–9731. [DOI] [PubMed] [Google Scholar]
- (87).Stillman M; Mack J; Kobayashi N, Theoretical aspects of the spectroscopy of porphyrins and phthalocyanines. J. Porph. Phthalocyan 2002, 6 (4), 296–300. [Google Scholar]
- (88).Rosa A; Ricciardi G; Baerends EJ; van Gisbergen SJA, The optical spectra of NiP, NiPz, NiTBP, and NiPc: Electronic effects of meso-tetraaza substitution and tetrabenzo annulation. J. Phys. Chem. A 2001, 105 (13), 3311–3327. [Google Scholar]
- (89).Rosa A; Ricciardi G; Baerends EJ, Synergism of porphyrin-core saddling and twisting of meso-aryl substituents. J. Phys. Chem. A 2006, 110 (15), 5180–5190. [DOI] [PubMed] [Google Scholar]
- (90).Zach PW; Freunberger SA; Klimant I; Borisov SM, Electron-deficient near-infrared Pt(II) and Pd(II) benzoporphyrins with dual phosphorescence and unusually efficient thermally activated delayed fluorescence: first demonstration of simultaneous oxygen and temperature sensing with a single emitter. ACS Applied Materials & Interfaces 2017, 9 (43), 38008–38023. [DOI] [PubMed] [Google Scholar]
- (91).Ravotto L; Meloni SL; Esipova TV; Masunov AE; Anna JM; Vinogradov SA, Three-photon spectroscopy of porphyrins. J. Phys. Chem. A 2020, 124 (52), 11038–11050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (92).Monson PR; McClain WM, Polarization dependence of two-photon absorption of tumbling molecules with application to liquid 1-chloronapthalene and benzene. J. Chem. Phys 1970, 53 (1), 29–37. [Google Scholar]
- (93).Monson PR; McClain WM, Complete polarization study of two-photon absorption of liquid 1-chloronapthalene. J. Chem. Phys 1972, 56 (10), 4817–4825. [Google Scholar]
- (94).Friese DH; Beerepoot MT; Ringholm M; Ruud K, Open-ended recursive approach for the calculation of multiphoton absorption matrix elements. J. Chem. Theory. Comput 2015, 11 (3), 1129–1144. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (95).Cronstrand P; Luo Y; Ågren H Multiphoton absorption of molecules. In Response Theory and Molecular Properties (A Tribute to Jan Linderberg and Poul Jørgensen), 2005; pp 1–21. [Google Scholar]
- (96).Granovsky AA, FireFly 8.2.0, http://classic.chem.msu.su/gran/firefly/index.html. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.