Next Article in Journal
Synthesis and Thermochromic Luminescence of Ag(I) Complexes Based on 4,6-Bis(diphenylphosphino)-Pyrimidine
Previous Article in Journal
Tungstoenzymes: Occurrence, Catalytic Diversity and Cofactor Synthesis
Previous Article in Special Issue
Cyclometalated Ir(III) Complexes with Curcuminoid Ligands as Active Second-Order NLO Chromophores and Building Blocks for SHG Polymeric Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Second Order Nonlinear Optical Properties of 4-Styrylpyridines Axially Coordinated to A4 ZnII Porphyrins: A Comparative Experimental and Theoretical Investigation †

1
Department of Chemistry, University of Milan, INSTM Research Unit, Via C. Golgi 19, 20133 Milan, Italy
2
CNR-SCITEC, Istituto di Scienze e Tecnologie Chimiche “G. Natta”, c/o University of Milan, Via C. Golgi 19, 20133 Milan, Italy
3
Fusion and Nuclear Security Department, Photonics Micro and Nanostructures Laboratory, ENEA, Via E. Fermi 45, 00044 Frascati, Rome, Italy
4
Department of Chemical Science and Technologies, University of Rome Tor Vergata, Via della Ricerca Scientifica 1, 00133 Rome, Italy
5
Department of Chemistry, University of Pavia, Via Taramelli 12, 27100 Pavia, Italy
*
Author to whom correspondence should be addressed.
This paper is dedicated to Prof. Maddalena Pizzotti for her 70th birthday, and for her scientific contribution in the field of porphyrin chemistry.
Submission received: 13 July 2020 / Revised: 8 August 2020 / Accepted: 12 August 2020 / Published: 14 August 2020

Abstract

:
In this research, two 4-styrylpyridines carrying an acceptor –NO2 (L1) or a donor –NMe2 group (L2) were axially coordinated to A4 ZnII porphyrins displaying in 5,10,15,20 meso position aryl moieties with remarkable electron withdrawing properties (pentafluorophenyl (TFP)), and with moderate to strong electron donor properties (phenyl (TPP) < 3,5-di-tert-butylphenyl (TBP) < bis(4-tert-butylphenyl)aniline) (TNP)). The second order nonlinear optical (NLO) properties of the resulting complexes were measured in CHCl3 solution by the Electric-Field-Induced Second Harmonic generation technique, and the quadratic hyperpolarizabilities βλ were compared to the Density Functional Theory (DFT)-calculated scalar quantities β||. Our combined experimental and theoretical approach shows that different interactions are involved in the NLO response of L1- and L2-substituted A4 ZnII porphyrins, suggesting a role of backdonation-type mechanisms in the determination of the negative sign of Electric-Field-Induced Second Harmonic generation (EFISH) βλ, and a not negligible third order contribution for L1-carrying complexes.

Graphical Abstract

1. Introduction

Since the beginning of the nineties, organometallic and coordination compounds have been extensively studied as chromophores for second order nonlinear optical (NLO) applications. Indeed, the introduction of a metal fragment in an organic environment allows a fine-tuning of its electronic properties [1]. By changing the electronic configuration, the oxidation state and the coordination sphere of the metal, new charge transfer (CT) transitions at low energy between the metal and the ligand (such as ligand-to-metal, LMCT, or metal-to-ligand, MLCT) can occur, enhancing the NLO response [2,3]. A topic of considerable investigation has been the effect of the coordination to a metal center on the second order NLO properties of π-delocalized nitrogen-donor push–pull monodentate and chelating polydentate ligands (such as pyridines [4], bipyridines [5], phenanthrolines [6], and terpyridines [7,8]). In particular, when the metal center is ZnII, a remarkable enhancement of the second order NLO response is recorded due to its inductive acceptor strength and its Lewis acid properties [5,6,7,8].
4-styrylpyridines have been coordinated to a variety of metal fragments, including carbonyl moieties [4,9] and metal carbonyl clusters [10]. Through coordination, the modulus of their quadratic hyperpolarizability β (which is the figure of merit of the second order NLO response) increases significantly, with a sign that depends on the nature of the substituent in para position. Electron donor groups lead to positive values, since the second order NLO response is dominated by an intraligand CT transition (ILCT), whereas electron acceptors result in a negative sign, due to the predominance of a MLCT transition [4,9].
Indeed, according to the “two-level” model developed by Oudar [11,12], the dipolar contribution to the quadratic hyperpolarizability β of a molecule depends on the electronic CT transitions of mobile polarizable electrons. Assuming that only one major CT dominates the second order NLO response, the component of the tensor β along the CT direction (βCT) is (Equation (1)):
β C T = 3 2 h 2 c 2 ν e g 2 r e g 2 Δ μ e g ( ν e g 2 υ L 2 ) ( v e g 2 4 υ L 2 ) ,
where υeg is the frequency of the CT transition, reg the transition dipole moment, Δμeg the difference between the excited and the ground state dipole moments, and υL the frequency of the incident radiation. Since all the terms in Equation (1) are positive, the sign of βCT is related to that of Δμeg, which depends on the β-dictating CT transition [9]. A useful way to evaluate βCT is solvatochromism [13].
The NLO response of porphyrin-based chromophores has been extensively investigated surveying traditional substitution patterns as for meso [14,15,16,17,18] and β-pyrrolic [19,20,21] push–pull arranged systems. The substituents position, the effect of the central metal atom, the solvent acidity, and the influence of aggregates formation have been thoroughly studied in such systems. On the contrary, the NLO properties of axially coordinated metal-porphyrins with both electron-donating and electron-accepting ligands have been only roughly investigated. Although some studies on electronic properties of axially coordinated A4 ZnII porphyrins were reported [22,23,24], only the effect of the coordination of 4-styrylpyridines carrying a –NMe2 donor or a –CF3 acceptor group to the axial position of ZnII, RuII, and OsII 5,10,15,20-Tetraphenylporphyrin (TPP) complexes was investigated for NLO purpose [25], with the latter being SHG-inactive due to the presence of a center of symmetry, which is lost by axial coordination. The lack of the expected increase of the β modulus of the ligand upon coordination was attributed to a noticeable axial π-backdonation from the dπ orbitals of the metal to the π* antibonding orbitals of the pyridine ligand, which is opposite to the σ-donation from the pyridine nitrogen atom to the metal [26]. Furthermore, for coordination of both 4-styrylpyridines to ZnTPP, the solvatochromic investigation provided a negative value of Δμeg, and therefore of βCT, in contrast with the positive β1907 experimentally measured by the Electric-Field-Induced Second Harmonic generation (EFISH) technique [27,28]. However, the paper did not fully address this discrepancy and did not provide any theoretical support to the experimental data.
The EFISH technique allows the determination of the quadratic hyperpolarizability through Equation (2):
γEFISH = μ0βλ(−2ω; ω, ω)/5kT + γ(−2ω; ω, ω, 0).
μ0βλ(−2ω; ω, ω)/5kT is a quadratic dipolar orientational contribution in which μ0 is the ground state molecular dipole moment and βλ is the projection along the dipole moment direction of the vectorial component βvec of the quadratic hyperpolarizability tensor working at an incident wavelength λ. γ(−2ω; ω, ω, 0) is a purely electronic cubic contribution that is a third order term at frequency ω of the incident light, which is usually negligible for asymmetric dipolar chromophores. However, for some A4 ZnII porphyrins with a substituent in β-pyrrolic position, it was recently shown that this approximation cannot be made [21].
βCT and βλ can be safely compared when the direction of the CT and of the dipole moment are almost coincident, as for 4-styrylpyridines axially coordinated to ZnTPP.
The purpose of the present research is to deepen the previous investigation [25] by considering from both an experimental and a theoretical point of view the impact that A4 ZnII porphyrin cores with different electron density might have on the second order NLO properties of 4-styrylpyridines bound in axial position.
As ligands, we have chosen an acceptor –NO2 (L1) and a donor –NMe2 group (L2) (Figure 1), equipped with a strong electron withdrawing –NO2 and a strong electron donor –NMe2 group, respectively. The arrows emphasize the direction conventionally assumed for the ground state dipole moment (from the negative to the positive pole of the molecule, Figure S1), which is opposite to the CT direction.
In order to have a comprehensive understanding of the role of the metal complex, four A4 ZnII porphyrin cores with increasing electron density were selected (Figure 1). More in detail, the ZnII porphyrins display in 5,10,15,20 meso position aryl moieties with remarkable electron acceptor properties (pentafluorophenyl (TFP)), and with moderate to strong electron donor properties (phenyl (TPP) < 3,5-di-tert-butylphenyl (TBP) < bis(4-tert-butylphenyl)anilines) (TNP)).
L1 and L2 have been coordinated to all the four porphyrin cores, and their EFISH quadratic hyperpolarizabilities βλ have been compared to the Density Functional Theory (DFT)-calculated scalar quantities β||, which derive from the full β tensor and correspond to 3/5 times βλ|| = (3/5) Ʃiiβi)/μ, where βi = (1/5) Ʃjijj + βjij + βjji)) [29]. Our combined experimental and theoretical approach sheds light on the different interactions involved in the second order response of L1 and L2 axially coordinated to A4 ZnII porphyrins, suggesting a role of backdonation-type interactions in the determination of the negative sign of EFISH βλ, and a not negligible third order contribution to the second order NLO response for L1-substituted complexes.

2. Results and Discussion

2.1. Synthesis of L1 and L2 Axially Substituted A4 ZnII Porphyrins

The fabrication of push–pull porphyrin systems showing a meso-substitution pattern on 5,15-positions is commonly known far from being trivial. Indeed, the asymmetric 10,20-diaryl substituted porphyrin core requires two reaction steps to be achieved and the further insertion of electron-donating and accepting pendants in 5,15-meso positions involves a tedious multistep pathway [30]. Instead, the β-pyrrolic substituted porphyrins, consisting of a more symmetric tetraaryl-substituted porphyrin core, are promptly accessible through a one-pot cyclo-condensation step among pyrrole and selected aldehydes [31,32]. However, the functionalization of β-pyrrolic positions and the subsequent introduction of proper substituents complicates the synthetic route [33,34].
Since the synthetic strategy of axially coordinated porphyrins is of a great value, the poor interest devoted on such class of porphyrins is quite surprising. Their fabrication relies on a less demanding synthetic strategy than that of β- and meso-substituted porphyrins by involving three effective and straightforward steps (Scheme 1): (a) cyclization of the core; (b) metalation; (c) axial coordination of metal center with proper ligands. As for β-pyrrolic substituted derivatives, the symmetric tetraaryl-substituted porphyrin core is easily attainable. Further, the coordination metals are typically inserted in the tetrapyrrolic core to quantitatively yield the desired metal complex. Finally, proper ligands can be successfully connected to the metal porphyrins by a simple axial-coordination step.
Except TPP, which was purchased from chemical vendors, the investigated free-base porphyrins TBP [30] and TFP [35] were synthesized as reported elsewhere. The acid-catalyzed Lindsey method [36] to prepare a porphyrin ring from aldehyde and pyrrole was adapted to the desired macrocycles. The following metalation step quantitatively determined the ZnII-complexes by refluxing the corresponding free-base porphyrin with Zn(OAc)2 in CHCl3. TBP was successfully obtained (43% yield) from pyrrole and 3,5-di-tert-butyl-benzaldehyde in CH2Cl2 with trifluoroacetic acid (TFA) as catalyst and 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) as oxidant before undergoing a metalation step [30]. TFP synthesis instead required a BF3·OEt2-catalyst in anhydrous CH2Cl2 to enable the condensation between pyrrole and pentafluorbenzaldehyde, and refluxing the solution with 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) to oxidize the macrocyclic ring (61% yield) before the subsequent zinc complexation [35,37]. Unlike TBP and TFP porphyrins, TNP nucleus was only obtained as Zn-complex from the following preparation strategy. A first cyclization step from p-bromobenzaldehyde and pyrrole yielded the p-bromophenyl substituted porphyrin core which was subsequently coordinated with a ZnII metal ion. A multiple Buchwald–Hartwig amination of the resulting metal-porphyrin successfully provided the ZnTNP metal complex [38].
While the electron-donating ligand L2 was purchased, the electron-accepting ligand L1 was designed and synthesized with the purpose of combining their electronic effect with the above-mentioned electron-deficient and electron-rich porphyrins. L1 was obtained by adapting an elsewhere reported procedure for similar stilbazole-based ligands [10] (see Supporting Information). Finally, L1 and L2 were coupled with the four ZnIIporphyrins TPP, TBP, TNP, and TFP by refluxing CH2Cl2 solutions of the proper components from 3 h to 72 h depending on the specific combinations between porphyrin and ligands.
Axial coordination enabled to address the main issues related to the poor overall reaction yields of fabrication of porphyrins with more complicated architectures. The synthetic efforts indeed were largely reduced by overcoming the more sensitive steps, namely, the functionalization and the subsequent covalent bonds formation at the periphery of porphyrin rings.
Thus, the trivial structural engineering demands of axially coordinated porphyrins allowed us to provide a prompt access to a large array of chromophores, offering a reasonable basis for a comprehensive screening of their NLO properties.

2.2. 1H-NMR and UV–Vis Spectroscopy

The 1H-NMR spectra of the axially substituted A4 ZnII porphyrins show a remarkable shielding effect for the protons in α-position of the pyridine nitrogen atom of the ligand, due to the local magnetic cone produced by the anisotropic interaction of the π-conjugated system of the porphyrin core with the magnetic field. For example, for free L1 in CDCl3, they are at 8.59 ppm (Figure S2), and they downshift to 3.74 for ZnTPP-L1 (Figure S3), 4.39 ppm for ZnTBP-L1 (Figure S4), and 2.44 ppm for ZnTFP-L1 (Figure S5). A comparable shift was reported for the axial coordination to ZnTPP of a ligand analogous to L1, but with a –CF3 instead of a –NO2 group (4.43 ppm) [25].
Moreover, the 1H-NMR spectra show that the axial coordination of the 4-styrylpyridine to the porphyrin occurs with retention of the (E)-configuration of the double bond of the ligand. Indeed, the coupling constant between the olefinic hydrogens is 16.0 Hz, as expected for a trans arrangement.
Finally, some 1H-NMR signals appear broadened, as typically observed after complexation of free-base porphyrins with metals and with the axial coordination of pyridine ligands [39].
We investigated L1, L2, and the axially coordinated ZnII porphyrins by UV–Vis absorption spectroscopy. The spectra in CHCl3 solution are reported in the Supporting Information (Figures S6–S10), while the corresponding experimental data are summarized in Table 1.
L1 and L2 show one electronic absorption band due to a π→π* internal transition for the former, whereas a n→π* ILCT transition from the –NMe2 donor end to the pyridine ring for the latter [10,40,41].
The UV–Vis spectra of the free porphyrins and of the corresponding ZnII complexes display the typical pattern expected on the basis of the Gouterman’s “four orbital model” [42]: an intense (ε ~ 105 M−1 cm−1) Soret or B band at about 412–440 nm, due to the S0→S2 transition (from the ground to the second excited state), and four (for free bases) or two (for the ZnII complexes) weaker (ε ~ 104 M−1 cm−1) Q bands in the range 500–660 nm, due to S0→S1 transitions (from the ground to the first excited state). The complexation to the metal ion induces a slight bathochromic shift of the B band (3–7 nm; 168–406 cm−1). A similar red shift occurs for the QIII and QII bands in ZnTPP and for the QII band in ZnTBP. Conversely, for the electron-poor TFP core, the QIII and QII bands experience a remarkable hypsochromic shift (30–50 nm; 996–1425 cm−1) after complexation to ZnII.
Despite the reported red shift of the B and particularly of the Q bands in a ZnTPP series with donor moieties in axial position [43,44], the UV–Vis spectra of the axially substituted ZnII porphyrins investigated here match those of the unsubstituted complexes. ZnTNP-L1 and ZnTNP-L2 are exceptions since they show a slight blue shift of the B (4–5 nm; 210–264 cm−1) and of the QII (2 nm; 55 cm−1) band in comparison to ZnTNP.
When a 4-styrylpyridine is coordinated in axial position of a metal porphyrin, three possible interactions can occur (Figure 2): (i) an axial σ-donation from the pyridine nitrogen atom to the metal center, when R is a donor group [26]; (ii) an axial π-backdonation from the dπ orbitals of the metal to the π* antibonding orbitals of the 4-styrylpyridine ligand, when R is an acceptor group [26]; and (iii) a peripheral π-backdonation from the dπ orbitals of the metal to the π* orbitals of the porphyrin ring [43,44].
When R is a donor group, σ-donation prevails with an accumulation of electron density on the metal center, which can be dissipated from the metal to the porphyrin core through peripheral d-π* backdonation. As a result, a red shift of the absorption spectrum of the axially substituted metal porphyrin is expected in comparison to the unsubstituted one [43,44]. The lack of such a bathochromic shift in the spectra of our axial complexes suggests a significant role of the axial π-backdonation from ZnII to the ligand, thanks to the energetically available π* antibonding orbitals of the 4-styrylpyridine. Therefore, the competition between peripheral and axial backdonation flattens out any spectroscopic effect that an axial coordination may produce.

2.3. Experimental and Theoretical Investigation of the Second Order NLO Properties

We performed EFISH measurements of L1, L2, and of the axially substituted A4 ZnII porphyrins on a 10−3 M solution in CHCl3 with a 1907 nm incident wavelength. Ground state dipole moments (μ0) and β|| (that is equal to 3/5 β1907) were computed by Density Functional Theory (DFT) and Coupled Perturbed DFT (CP-DFT) M06-2X/6-311G (d) calculations. The details on both methodologies are in the Materials and Methods Section, and Table 2 collects the data.
It should be noted that μ0 of L1 is significantly greater than that computed, at the same level of the theory, for the previously investigated 4-styrylpyridine carrying a –CF3 instead of a –NO2 group [10] equal to 0.44 D. Moreover, the dipole of L2 is comparable to the one previously obtained by HF/6-311++G** calculations (6.06 D) [4].
Upon coordination to the axial position of the ZnII porphyrins, an increase of the μ0 value of L2 occurs, with enhancement factors (μ0 EF = μ0,complex0,ligand) in the range of 1.41–1.61. The computed μ0 are substantially the same regardless of the porphyrin core, suggesting that, when the σ-donation by the donor-substituted 4-styrylpyridine is remarkable, the core acts as a buffer of the metal ion electron density through the peripheral d-π* backdonation mechanism (Figure 2). Accordingly, a closer look at the axial complexes with L2 shows that the μ0 and μ0 EF values follow the trend based on the different electron properties imparted to the porphyrin macrocycles by the substituents in 5,10,15,20 meso positions. As the core becomes more electron-poor (ZnTNP < ZnTBP < ZnTPP < ZnTFP), its ability to dissipate the metal ion electron density by peripheral backdonation increases slightly, pushing up μ0 and μ0 EF.
On the other hand, when L1 is coordinated to the axial position of the ZnII porphyrins, the μ0 values experience a huge decrease and essentially vanish for all the compounds. This is quite surprising since for the ligand similar to L1, carrying a –CF3 instead of a –NO2 acceptor group, coordination to ZnTPP led to a remarkable increase of μ0 [25]. Apparently, in the present case, the higher electron acceptor properties of the –NO2 moiety (Hammett σpara = 0.78 vs. 0.54 for –CF3) [45] result into the quasi-cancellation of the ground state dipole moment of the axially coordinated Zn porphyrins. This could be ascribed to the axial π-backdonation from the metal towards the pyridinic nitrogen atom (see Figure 1), counteracting the polarity (from –NO2 to Npy) of the free ligand in the same way as described for 4-nitropyridine-1-oxide in comparison to 4-nitropyridine [46]. The μ0 of ZnTFP-L1, which is the lowest among the series, supports this hypothesis, since the pentafluorophenyl rings in meso position impart a significant electron depletion to the core, thus enhancing the Lewis acid properties of the ZnII ion.
The axial coordination of L2 to ZnII porphyrins maintains and emphasizes the ground state charge distribution in agreement with an enhancement of the polarity (from Npy to –NMe2) in comparison to the free ligand. The opposite relative orientation of the ligands’ dipole moment with respect to the metal-porphyrins reflects also into significantly different Zn–Npy distances, which are longer by about 0.02 Å for the L1-complexes with respect to the L2 ones (Table S1).
In agreement with the above considerations, the calculated β|| values for L2-complexes are positive, because the second order NLO response of L2 is dominated by a n→π* ILCT transition along the dipole moment axis, which is enhanced by σ donation to the ZnII porphyrin core. As expected [6], the β|| value of L2 increases upon coordination, with enhancement factors (β|| EF = β||,complex||,ligand) that are in the range of 1.62–2.0 and follow the trend of the μ0 EF.
On the other hand, in accordance with the quasi-null μ0, the values of β|| computed for the ZnII porphyrins with axially coordinated L1 are very low and with no enhancement in comparison to the free ligand, as expected for a significant axial π-backdonation. Furthermore, when L1 is in combination with ZnTFP, CP-DFT calculations provide a β|| with a negative sign, associated with an inversion of the dipole moment direction thanks to the very electron-poor porphyrin core.
The experimental EFISH β1907 of L1 and L2 are positive, as expected, and the value recorded for L2 is in nice agreement with the computed β|| and with the experimental value already reported in the literature [10].
Conversely, all the axial porphyrins display a negative EFISH second order NLO response in accordance to the former research that reported a negative βCT (provided by solvatochromism) for the coordination to ZnTPP of L2 and of the ligand similar to L1, but with a –CF3 instead of a –NO2 group [25].
Since a negative value of βCT (and of β1907) arises from a negative Δμeg (Equation (1)), the CT transition mainly responsible for the second order NLO response of our ZnII porphyrins with L1 and L2 in axial position leads to a decrease of the excited state dipole moment in comparison to the ground state one.
Thus, for L2-substituted complexes, we ought to assume that the key role in the determination of the sign of the EFISH response is played by the peripheral d-π* backdonation (Figure 2), which favors a charge dissipation on the π-delocalized system of the porphyrin macrocycle, with a lowering of the dipole moment in the excited state. Moreover, our data confirm that no enhancement of the β1907 of L2 occurs for coordination to ZnTPP (Table 2 and Reference [25]), whereas its second order NLO response increases when it is in axial position of the other ZnII porphyrins, reaching the highest absolute value in combination to the most electron-poor core (ZnTFP). This is in agreement with an axial interaction mainly dominated by σ-donation, followed by peripheral backdonation, as discussed also for the dipole moments.
The interpretation of the high and negative μβ1907 data for the axial complexes with L1 is more intriguing. A low μ0 is not inconsistent with a high quadratic hyperpolarizability, since the latter depends on Δμeg (Equation (1)). Therefore, even when μ0 is small, a high value of the excited state dipole moment is enough to reach a significant second order NLO response. However, when the μ0 of the solute is negligible, the EFISH technique is hardly feasible.
Very recently, some of us reported a not negligible contribution to γEFISH of the third order term γ(−2ω;ω,ω,0) in Equation (2) for some A4 ZnII porphyrins with a substituent in β-pyrrolic position [21]. In particular, CP-DFT calculations provided large and negative γ|| values (γ|| = γ(−2ω;ω,ω,0)) for two chromophores having μ0 = 0.6 D, which is of the same order of magnitude as the ones of the L1-substituted axial porphyrins investigated here. Hence, we suggest that the EFISH second order NLO response of the latter might be affected by a significant and negative contribution of the electronic third order cubic term, which exceeds the value of the dipolar orientational contribution μ0βλ/5kT. Therefore, apparently, coordination of a 4-stryrylpyridine with a strong electron withdrawing group in the axial position of an A4 ZnII porphyrin, being dominated by axial π-backdonation, leads to a quasi-null dipole moment, and therefore, to a chromophore with important third order properties. The very high value recorded for ZnTFP-L1, albeit the lowest dipole moment within the series, is unexpected and prompts us to deepen the present investigation with further DFT calculations and Third Harmonic Generation (THG) measurements, which will be the topic of another paper.

3. Materials and Methods

3.1. General

All reagents and solvents were purchased from Sigma-Aldrich (Merck Life Science S.r.l., Milan, Italy) and used as received, except for NEt3 (freshly distilled over KOH) and CH2Cl2 anhydrous for the synthesis of TFP (freshly distilled over CaH2). Milli-Q water was collected from the Millipore apparatus, equipped with 0.22 μm filters. Glassware was flame-dried under vacuum before use when necessary. Microwave assisted reactions were performed using a Milestone Micro-SYNTH instrument (Milestone Srl, Sorisole, Italy). Silica gel for gravimetric chromatography (Geduran Si 60, 63–200 μm) and for flash chromatography (Kieselgel 60, 0.040–0.063 mm) were purchased from Merck (Merck KGaA, Darmstadt, Germany). 1H-NMR spectra were recorded on a Bruker AMX 300 and on a Bruker Avance DRX-400 (Bruker Italia S.r.l., Milan, Italy) in CDCl3 or in THF-d8 to enhance resolution (Cambridge Isotope Laboratories Inc., Tewksbury, MA, USA). Elemental analyses were carried out with a Perkin-Elmer CHN 2400 instrument in the Analytical Laboratories of the Department of Chemistry at the University of Milan. Electronic absorption spectra were recorded in CHCl3 solution at room temperature on a Shimadzu UV 3600 spectrophotometer (Shimadzu Corporation, Kyoto, Japan). The starting A4 ZnII porphyrin complexes were prepared as reported in literature [30]. The experimental details on the synthesis and characterization of the investigated axial compounds are reported in the Supplementary Materials.

3.2. EFISH and THG Measurements

The second order NLO responses of the axially coordinated A4 ZnII porphyrins with L1 and L2 ligands were measured by the EFISH technique [27,28] in the Department of Chemistry of the University of Milano (Milano, Italy) through a prototype apparatus made by SOPRA (Paris, France). For each chromophore, measurements were performed on freshly prepared solutions in CHCl3 at 10−3 M concentration. The 1907 nm laser incident wavelength was chosen because its second harmonic (at 953 nm) is far enough from the absorption bands of the chromophores to avoid possible enhancement of the second order NLO response due to resonance effects. The incident beam was obtained by Raman shifting of the 1064 nm emission of a Q-switched Nd:YAG laser in a high-pressure hydrogen cell (60 bar). A liquid cell with thick windows in the wedge configuration was used to obtain the Maker fringe pattern originated by the harmonic intensity variation as a function of the liquid cell translation. In the EFISH experiments, this incident beam was synchronized with a direct current field applied to the solution, with 60 and 20 ns pulse duration, respectively, in order to break its centrosymmetry. The comparison of the harmonic signal of the chromophore solution with that of the pure solvent allowed the determination of its second order NLO response (assumed to be real because the imaginary part was neglected). The μ0β1907 values reported in Table 2 are the mean values of 12 successive measurements performed on the same sample and are defined according to the “phenomenological” convention [47]. The experimental error on the EFISH measurements is 10–15%.

3.3. Computational Calculations

Geometry optimizations were performed with the 6-311G(d) basis set using the M06 functional [48] due to its specific parametrization on organometallic complexes. Using the same basis set, SHG first hyperpolarizabilities, i.e., the β(–2ω; ω, ω) tensors, were computed within the Coupled Perturbed Kohn–Sham (CPKS) approach at the same frequency (1907 nm) used in the EFISH experiments. The M06-2X functional [48], which has been recently recommended for hyperpolarizability calculations of mid-size chromophores [49], was adopted for β calculation. The same functional was used for determining the dipole moments μ0. A pruned (99,590) grid was selected for computation and use of two-electron integrals and their derivatives. To get a meaningful comparison with the experimental data, the scalar quantity β|| was derived from the full tensors β; β|| corresponds to 3/5 times βλ, the projection along the dipole moment direction of the vectorial component of the β tensor, that is β|| = (3/5) Ʃi(μiβi)/μ, where βi = (1/5)Ʃjijj + βjij + βjji) [29]. All Density Functional Theory (DFT) calculations were performed using the Gaussian16 suite of programs [50].

4. Conclusions

Two 4-styrylpyridines carrying an electron acceptor –NO2 (L1) or an electron donor –NMe2 group (L2), respectively, were axially coordinated to A4 ZnII porphyrins with aryl moieties of increasing electron density in 5,10,15,20 meso position: pentafluorophenyl (TFP) < phenyl (TPP) < 3,5-di-tert-butylphenyl (TBP) < bis(4-tert-butylphenyl)aniline (TNP). The EFISH quadratic hyperpolarizabilities βλ were measured and compared to the DFT-calculated scalar quantities β||. Our combined experimental and theoretical approach sheds light on the different interactions involved in the second order response of L1 and L2 axially coordinated to A4 ZnII porphyrins, suggesting a role of backdonation-type interactions in the determination of the negative sign of EFISH βλ, and a not negligible third order contribution to the second order NLO response for L1-substituted complexes.
An increase of the ground state dipole moment of L2 occurs upon coordination, and the computed μ0 values are the same regardless of the macrocycle, suggesting a remarkable axial σ-donation from the pyridine nitrogen atom to the metal center, with an accumulation of electron density on this latter, which can be dissipated through a peripheral metal→porphyrin dπ-π* backdonation.
On the other hand, when L1 is coordinated to the axial position of the ZnII porphyrins, the μ0 values essentially vanish, as expected for a significant axial π-backdonation from the metal towards the pyridine nitrogen atom, counteracting the polarity (from –NO2 to Npy) of the free ligand.
In accordance, the computed β|| for L2 and the corresponding axial complexes are positive, with an enhancement upon coordination, while the quasi-null μ0 of the ZnII porphyrins with L1 produces very low β|| values and with no enhancement in comparison to the free ligand. Moreover, when L1 is in combination with the most electron-poor core of the series (ZnTFP), CP-DFT calculations provide a β|| with a negative sign.
The EFISH measurements confirm a positive β1907 for L1 and L2, whereas all the axial porphyrins display a negative second order NLO response in accordance to the negative βCT reported in a previous research for the coordination to ZnTPP of L2 and of the ligand similar to L1, but with a –CF3 instead of a –NO2 group [26]. Hence, we suggest that for L2-substituted complexes, the key role in the determination of the sign of the EFISH response is played by the peripheral d-π* backdonation, which leads to a decrease of the excited state dipole moment in comparison to the ground state one.
For L1-substituted complexes, the striking contrast between the almost vanishing μ0 and the high and negative μβ1907 data prompts us to suggest that the EFISH second order NLO response might be affected by a significant and negative contribution of the electronic third order cubic term γ(−2ω; ω, ω, 0) to γEFISH, overwhelming the dipolar orientational contribution μ0βλ/5kT. Therefore, apparently, the coordination of a 4-stryrylpyridine with a strong electron withdrawing group in the axial position of an A4 ZnII porphyrin, being dominated by axial π-backdonation, originates a chromophore with important third order properties.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2304-6740/8/8/45/s1: the synthesis and characterization of L1 and of the axially substituted ZnII porphyrins; the 1H-NMR spectra; the UV–Vis spectra in CHCl3 solution; the optimized geometry of L1 and L2; the computed distances between Zn and the pyridine nitrogen atoms.

Author Contributions

Conceptualization: F.T., G.D.C. and A.O.B.; Formal analysis: A.F.; Funding acquisition: G.D.C.; Investigation: F.T. and S.R.; Writing-original draft: F.T. and G.D.C.; Writing-review and editing: F.L. and A.O.B. All the authors have given approval to the final version of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Milan (Piano Sostegno alla Ricerca 2018 LINEA 2 Azione A—Giovani Ricercatori).

Acknowledgments

We gratefully acknowledge Regione Lombardia and Fondazione Cariplo for the use of instrumentation purchased through the SmartMatLab Centre project (Fondazione Cariplo Grant 2013-1766).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Di Bella, S. Second-order nonlinear optical properties of transition metal complexes. Chem. Soc. Rev. 2001, 30, 355–366. [Google Scholar] [CrossRef]
  2. Cariati, E.; Pizzotti, M.; Roberto, D.; Tessore, F.; Ugo, R. Coordination and organometallic compounds and inorganic-organic hybrid crystalline materials for second-order non-linear optics. Coord. Chem. Rev. 2006, 250. [Google Scholar] [CrossRef]
  3. Di Bella, S.; Dragonetti, C.; Pizzotti, M.; Roberto, D.; Tessore, F.; Ugo, R. Molecular Organometallic Materials for Optics; Springer: Berlin, Germany, 2010; Volume 28. [Google Scholar]
  4. Roberto, D.; Ugo, R.; Bruni, S.; Cariati, E.; Cariati, F.; Fantucci, P.; Invernizzi, I.; Quici, S.; Ledoux, I.; Zyss, J. Quadratic hyperpolarizability enhancement of para-substituted pyridines upon coordination to organometallic moieties:  The ambivalent donor or acceptor role of the metal. Organometallics 2000, 19, 1775–1788. [Google Scholar] [CrossRef]
  5. Baccouche, A.; Peigné, B.; Ibersiene, F.; Hammoutène, D.; Boutarfäa, A.; Boucekkine, A.; Feuvrie, C.; Maury, O.; Ledoux, I.; Le Bozec, H. Effects of the metal center and substituting groups on the linear and nonlinear optical properties of substituted styryl-bipyridine metal(II) dichloride complexes: DFT and TDDFT computational investigations and harmonic light scattering measurements. J. Phys. Chem. A 2010, 114, 5429–5438. [Google Scholar] [CrossRef] [PubMed]
  6. Roberto, D.; Ugo, R.; Tessore, F.; Lucenti, E.; Quici, S.; Vezza, S.; Fantucci, P.; Invernizzi, I.; Bruni, S.; Ledoux-Rak, I.; et al. Effect of the Coordination to M(II) Metal Centers (M = Zn, Cd, Pt) on the Quadratic Hyperpolarizability of Various Substituted 5-X-1,10-phenanthrolines (X) Donor Group) and of trans-4-(Dimethylamino)-4′-stilbazole. Organometallics 2002, 21. [Google Scholar] [CrossRef]
  7. Roberto, D.; Tessore, F.; Ugo, R.; Bruni, S.; Manfredi, A.; Quici, S. Terpyridine Zn(II), Ru(III) and Ir(III) complexes as new asymmetric chromophores for nonlinear optics: First evidence for a shift from positive to negative value of the quadratic hyperpolarizability of a ligand carrying an electron donor substituent upon. Chem. Commun. 2002, 2, 846–847. [Google Scholar] [CrossRef]
  8. Tessore, F.; Roberto, D.; Ugo, R.; Pizzotti, M.; Quici, S.; Cavazzini, M.; Bruni, S.; De Angelis, F. Terpyridine Zn(II), Ru(III), and Ir(III) complexes: The relevant role of the nature of the metal ion and of the ancillary ligands on the second-order nonlinear response of terpyridines carrying electron donor or electron acceptor groups. Inorg. Chem. 2005, 44. [Google Scholar] [CrossRef]
  9. Kanis, D.R.; Lacroix, P.G.; Ratner, M.A.; Marks, T.J. Electronic Structure and Quadratic Hyperpolarizabilities in Organotransition-Metal Chromophores Having Weakly Coupled π-Networks. Unusual Mechanisms for Second-Order Response. J. Am. Chem. Soc. 1994, 116, 10089–10102. [Google Scholar] [CrossRef]
  10. Lucenti, E.; Cariati, E.; Dragonetti, C.; Manassero, L.; Tessore, F. Effect of the coordination to the “Os3(CO)11” cluster core on the quadratic hyperpolarizability of trans-4-(4′-X-styryl)pyridines (X = NMe2, t-Bu, CF3) and trans,trans-4-(4′-NMe2-phenyl-1,3-butadienyl. Organometallics 2004, 23. [Google Scholar] [CrossRef]
  11. Oudar, J.L. Optical nonlinearities of conjugated molecules. Stilbene derivatives and highly polar aromatic compounds. J. Chem. Phys. 1977, 67, 446–457. [Google Scholar] [CrossRef]
  12. Oudar, J.L.; Chemla, D.S. Hyperpolarizabilities of the nitroanilines and their relations to the excited state dipole moment. J. Chem. Phys. 1977, 66, 2664–2668. [Google Scholar] [CrossRef]
  13. Bruni, S.; Cariati, E.; Cariati, F.; Porta, F.A.; Quici, S.; Roberto, D. Determination of the quadratic hyperpolarizability of trans-4-[4-(dimethylamino)styryl]pyridine and 5-dimethylamino-1,10-phenanthroline from solvatochromism of absorption and fluorescence spectra: A comparison with the electric-field-induced second-harmon. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2001, 57, 1417–1426. [Google Scholar] [CrossRef]
  14. Pizzotti, M.; Annoni, E.; Ugo, R.; Bruni, S.; Quici, S.; Fantucci, P.; Bruschi, M.; Zerbi, G.; Zoppo, M. Del A multitechnique investigation of the second order NLO response of a 10,20-diphenylporphyrinato nickel(II) complex carrying a phenylethynyl based push-pull system in the 5- and 15-positions. J. Porphyr. Phthalocyanines 2004, 8, 1311–1324. [Google Scholar] [CrossRef]
  15. Morotti, T.; Pizzotti, M.; Ugo, R.; Quici, S.; Bruschi, M.; Mussini, P.; Righetto, S. Electronic characterisation and significant second-order NLO response of 10,20-diphenylporphyrins and their ZnII complexes substituted in the meso position with π-delocalised linkers carrying push or pull groups. Eur. J. Inorg. Chem. 2006, 1743–1757. [Google Scholar] [CrossRef]
  16. De Angelis, F.; Fantacci, S.; Sgamellotti, A.; Pizzotti, M.; Tessore, F.; Orbelli Biroli, A. Time-dependent and coupled-perturbed DFT and HF investigations on the absorption spectrum and non-linear optical properties of push-pull M(II)-porphyrin complexes (M = Zn, Cu, Ni). Chem. Phys. Lett. 2007, 447, 10–15. [Google Scholar] [CrossRef]
  17. Pizzotti, M.; Tessore, F.; Orbelli Biroli, A.; Ugo, R.; De Angelis, F.; Fantacci, S.; Sgamellotti, A.; Zuccaccia, D.; Macchioni, A. An EFISH, theoretical, and PGSE NMR investigation on the relevant role of aggregation on the second order response in CHCl3 of the push-pull chromophores [5-[[4′-(dimethylamino)phenyl]ethynyl]-15-[(4′′-nitrophenyl)ethynyl]-10,20diphenylporphyri. J. Phys. Chem. C 2009, 113. [Google Scholar] [CrossRef]
  18. Orbelli Biroli, A.; Tessore, F.; Righetto, S.; Forni, A.; Macchioni, A.; Rocchigiani, L.; Pizzotti, M.; Di Carlo, G. Intriguing Influence of −COOH-Driven Intermolecular Aggregation and Acid-Base Interactions with N,N-Dimethylformamide on the Second-Order Nonlinear-Optical Response of 5,15 Push-Pull Diarylzinc(II) Porphyrinates. Inorg. Chem. 2017, 56. [Google Scholar] [CrossRef] [Green Version]
  19. Annoni, E.; Pizzotti, M.; Ugo, R.; Quici, S.; Morotti, T.; Bruschi, M.; Mussini, P. Synthesis, electronic characterisation and significant second-order non-linear optical responses of meso-tetraphenylporphyrins and their ZnII complexes carrying a push or pull group in the β pyrrolic position. Eur. J. Inorg. Chem. 2005, 3857–3874. [Google Scholar] [CrossRef]
  20. Tessore, F.; Orbelli Biroli, A.; Di Carlo, G.; Pizzotti, M. Porphyrins for Second Order Nonlinear Optics (NLO): An Intriguing History. Inorganics 2018, 6, 81. [Google Scholar] [CrossRef] [Green Version]
  21. Di Carlo, G.; Pizzotti, M.; Righetto, S.; Forni, A.; Tessore, F. Electric-Field-Induced Second Harmonic Generation Nonlinear Optic Response of A4 β-Pyrrolic-Substituted ZnII Porphyrins: When Cubic Contributions Cannot Be Neglected. Inorg. Chem. 2020, 59, 7561–7570. [Google Scholar] [CrossRef]
  22. Bajju, G.D.; Kundan, S.; Kapahi, A.; Gupta, D. Synthesis and Spectroscopic Studies of Axially Ligated Zn(II)5,10,15,20-meso-tetra(p-chlorophenyl)porphyrin with Oxygen and Nitrogen Donors. J. Chem. 2013, 2013, 135815. [Google Scholar] [CrossRef]
  23. Charisiadis, A.; Stangel, C.; Nikolaou, V.; Roy, M.S.; Sharma, G.D.; Coutsolelos, A.G. A supramolecular assembling of zinc porphyrin with a π-conjugated oligo(phenylenevinylene) (oPPV) molecular wire for dye sensitized solar cell. Rsc Adv. 2015, 5, 88508–88519. [Google Scholar] [CrossRef]
  24. Xie, M.; Bai, F.-Q.; Zhang, H.-X.; Zheng, Y.-Q. The influence of an inner electric field on the performance of three types of Zn-porphyrin sensitizers in dye sensitized solar cells: A theoretical study. J. Mater. Chem. C 2016, 4, 10130–10145. [Google Scholar] [CrossRef]
  25. Annoni, E.; Pizzotti, M.; Ugo, R.; Quici, S.; Morotti, T.; Casati, N.; Macchi, P. The effect on E-stilbazoles second order NLO response by axial interaction with M(II) 5,10,15,20-tetraphenyl porphyrinates (M = Zn, Ru, Os); a new crystalline packing with very large holes. Inorg. Chim. Acta 2006, 359, 3029–3041. [Google Scholar] [CrossRef]
  26. Cole, S.J.; Curthoys, G.C.; Magnusson, E.A. Ligand Binding by Metalloporphyrins. I. Thermodynamic Functions of Porphyriniron(II)-Pyridine Complexes. J. Am. Chem. Soc. 1970, 92, 2991–2996. [Google Scholar] [CrossRef]
  27. Levine, B.F.; Bethea, C.G. Molecular hyperpolarizabilities determined from conjugated and nonconjugated organic liquids. Appl. Phys. Lett. 1974, 24, 445–447. [Google Scholar] [CrossRef]
  28. Singer, K.D.; Garito, A.F. Measurements of molecular second order optical susceptibilities using dc induced second harmonic generation. J. Chem. Phys. 1981, 75, 3572–3580. [Google Scholar] [CrossRef]
  29. Kurtz, H.A.; Dudis, D.S. Quantum mechanical methods for predicting nonlinear optical properties. In Reviews in Computational Chemistry; 2007; Volume 12, pp. 241–279. [Google Scholar] [CrossRef]
  30. Di Carlo, G.; Orbelli Biroli, A.; Pizzotti, M.; Tessore, F.; Trifiletti, V.; Ruffo, R.; Abbotto, A.; Amat, A.; De Angelis, F.; Mussini, P.R. Tetraaryl ZnII porphyrinates substituted at β-pyrrolic positions as sensitizers in dye-sensitized solar cells: A comparison with meso-disubstituted push-pull ZnII porphyrinates. Chem. A Eur. J. 2013, 19, 10723–10740. [Google Scholar] [CrossRef]
  31. Di Carlo, G.; Orbelli Biroli, A.; Tessore, F.; Caramori, S.; Pizzotti, M. β-Substituted ZnII porphyrins as dyes for DSSC: A possible approach to photovoltaic windows. Coord. Chem. Rev. 2018, 358, 153–177. [Google Scholar] [CrossRef]
  32. Di Carlo, G.; Orbelli Biroli, A.; Pizzotti, M.; Tessore, F. Efficient sunlight harvesting by A4 β-Pyrrolic Substituted ZnII Porphyrins: A Mini-Review. Front. Chem. 2019, 7. [Google Scholar] [CrossRef]
  33. Di Carlo, G.; Caramori, S.; Casarin, L.; Orbelli Biroli, A.; Tessore, F.; Argazzi, R.; Oriana, A.; Cerullo, G.; Bignozzi, C.A.; Pizzotti, M. Charge Transfer Dynamics in β and Meso Substituted Dithienylethylene Porphyrins. J. Phys. Chem. C 2017, 121, 18385–18400. [Google Scholar] [CrossRef]
  34. Di Carlo, G.; Orbelli Biroli, A.; Tessore, F.; Rizzato, S.; Forni, A.; Magnano, G.; Pizzotti, M. Light-Induced Regiospecific Bromination of meso-Tetra(3,5-di-tert-butylphenyl)Porphyrin on 2,12 β-Pyrrolic Positions. J. Org. Chem. 2015, 80, 4973–4980. [Google Scholar] [CrossRef] [PubMed]
  35. Orbelli Biroli, A.; Tessore, F.; Di Carlo, G.; Pizzotti, M.; Benazzi, E.; Gentile, F.; Berardi, S.; Bignozzi, C.A.; Argazzi, R.; Natali, M.; et al. Fluorinated ZnII Porphyrins for Dye-Sensitized Aqueous Photoelectrosynthetic Cells. Acs Appl. Mater. Interfaces 2019, 11, 32895–32908. [Google Scholar] [CrossRef] [PubMed]
  36. Lindsey, J.S.; Schreiman, I.C.; Hsu, H.C.; Kearney, P.C.; Marguerettaz, A.M. Rothemund and Adler-Longo reactions revisited: Synthesis of tetraphenylporphyrins under equilibrium conditions. J. Org. Chem. 1987, 52, 827–836. [Google Scholar] [CrossRef]
  37. Berardi, S.; Caramori, S.; Benazzi, E.; Zabini, N.; Niorettini, A.; Orbelli Biroli, A.; Pizzotti, M.; Tessore, F.; Di Carlo, G. Electronic Properties of Electron-Deficient Zn(II) Porphyrins for HBr Splitting. Appl. Sci. 2019, 9, 2739. [Google Scholar] [CrossRef] [Green Version]
  38. Covezzi, A.; Orbelli Biroli, A.; Tessore, F.; Forni, A.; Marinotto, D.; Biagini, P.; Di Carlo, G.; Pizzotti, M. 4D-π-1A type β-substituted ZnII-porphyrins: Ideal green sensitizers for building-integrated photovoltaics. Chem. Commun. 2016, 52. [Google Scholar] [CrossRef] [Green Version]
  39. Shirazi, A.; Goff, H.M. Carbon-13 and Proton NMR Spectroscopy of Four- and Five-Coordinate Cobalt(II) Porphyrins: Analysis of NMR Isotropic Shifts. Inorg. Chem. 1982, 3420–3425. [Google Scholar] [CrossRef]
  40. Tessore, F.; Roberto, D.; Ugo, R.; Mussini, P.; Quici, S.; Ledoux-Rak, I.; Zyss, J. Large, concentration-dependent enhancement of the quadratic hyperpolarizability of [Zn(CH3CO2)2(L)2] in CHCl3 on substitution of acetate by triflate. Angew. Chem. Int. Ed. 2003, 42. [Google Scholar] [CrossRef]
  41. Tessore, F.; Locatelli, D.; Righetto, S.; Roberto, D.; Ugo, R.; Mussini, P. An Investigation on the Role of the Nature of Sulfonate Ancillary Ligands on the Strength and Concentration Dependence of the Second-Order NLO Responses in CHCl3 of Zn(II) Complexes with 4,4′-trans-NC5H4CH=CHC6H4NMe2 and 4,4′-trans,trans-NC5H4(CH=CH)2C6H4NMe2. Inorg. Chem. 2005, 44, 2437–2442. [Google Scholar]
  42. Gouterman, M. Spectra of porphyrins. J. Mol. Spectrosc. 1961, 6, 138–163. [Google Scholar] [CrossRef]
  43. Nappa, M.; Valentine, J.S. The Influence of Axial Ligands on Metalloporphyrin Visible Absorption Spectra. Complexes of Tetraphenylporphinatozinc. J. Am. Chem. Soc. 1978, 100, 5075–5080. [Google Scholar] [CrossRef]
  44. Szintay, G.; Horváth, A. Temperature dependence study of five-coordinate complex formation of zinc(II) octaethyl and tetraphenylporphyrin. Inorg. Chim. Acta 2000, 310, 175–182. [Google Scholar] [CrossRef]
  45. Hansch, C.; Leo, A.; Taft, R.W. A Survey of Hammett Substituent Constants and Resonance and Field Parameters. Chem. Rev. 1991, 91, 165–195. [Google Scholar] [CrossRef]
  46. Zyss, J.; Chemla, D.S.; Nicoud, J.F. Demonstration of efficient nonlinear optical crystals with vanishing molecular dipole moment: Second-harmonic generation in 3-methyl-4-nitropyridine-1-oxide. J. Chem. Phys. 1981, 74, 4800–4811. [Google Scholar] [CrossRef]
  47. Willetts, A.; Rice, J.E.; Burland, D.M.; Shelton, D.P. Problems in the comparison of theoretical and experimental hyperpolarizabilities. J. Chem. Phys. 1992, 97, 7590–7599. [Google Scholar] [CrossRef]
  48. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other function. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef] [Green Version]
  49. Johnson, L.E.; Dalton, L.R.; Robinson, B.H. Optimizing calculations of electronic excitations and relative hyperpolarizabilities of electrooptic chromophores. Acc. Chem. Res. 2014, 47, 3258–3265. [Google Scholar] [CrossRef]
  50. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16; revision, A.03; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
Figure 1. Molecular structures of ZnII porphyrins and ligands.
Figure 1. Molecular structures of ZnII porphyrins and ligands.
Inorganics 08 00045 g001
Scheme 1. Schematic synthetic pathway for the preparation of A4 ZnII porphyrins axially coordinated with 4-styrylpyridine ligands.
Scheme 1. Schematic synthetic pathway for the preparation of A4 ZnII porphyrins axially coordinated with 4-styrylpyridine ligands.
Inorganics 08 00045 sch001
Figure 2. Possible interactions in axially substituted metal porphyrins.
Figure 2. Possible interactions in axially substituted metal porphyrins.
Inorganics 08 00045 g002
Table 1. Electronic absorption data in CHCl3 solution of ligands acceptor –NO2 (L1) and donor –NMe2 (L2) and of the axially coordinated A4 ZnII porphyrins.
Table 1. Electronic absorption data in CHCl3 solution of ligands acceptor –NO2 (L1) and donor –NMe2 (L2) and of the axially coordinated A4 ZnII porphyrins.
CompoundLigand Band
λmax (nm)
Soret or B Band
λmax (nm)
QIV Band
λmax (nm)
QIII Band
λmax (nm)
QII Band
λmax (nm)
QI Band
λmax (nm)
L1332
L2376
TFP 412506583637657
ZnTFP 419 551584
ZnTFP-L1321419 551585627
ZnTFP-L2 419480550585617
TPP 418515550589646
ZnTPP 422 553594
ZnTPP-L1319421 553594
ZnTPP-L2 422 553594
TBP 421518554592648
ZnTBP 424 552595
ZnTBP-L1329424 552595
ZnTBP-L2372424 552595
ZnTNP306438 559605
ZnTNP-L1310433 559603
ZnTNP-L2305434 559603
Table 2. Experimental Electric-Field-Induced Second Harmonic generation (EFISH) μ0β1907 values (10−3 M solution in CHCl3), and theoretical μ0 and β|| values of ligands L1 and L2 and of the axially coordinated A4 ZnII porphyrins.
Table 2. Experimental Electric-Field-Induced Second Harmonic generation (EFISH) μ0β1907 values (10−3 M solution in CHCl3), and theoretical μ0 and β|| values of ligands L1 and L2 and of the axially coordinated A4 ZnII porphyrins.
Compoundμ0β1907
(×10−48 esu)
μ0 calc (μ0 EF a)
(D)
β1907 (β1907 EF b)
(×10−30 esu)
β|| (β|| EF d)
(×10−48 esu)
L1+3102.67+11611
ZnTFP-L1−16800.29nd c−3
ZnTPP-L1−5400.36nd7
ZnTBP-L1−8800.68nd7
ZnTNP-L1−7330.70nd12
L2+2506.59+38 d37
ZnTFP-L2−98010.63 (1.61)−92 (2.42)75 (2.0)
ZnTPP-L2−2809.84 (1.49)−28 (0.74)65 (1.75)
ZnTBP-L2−8609.64 (1.46)−89 (2.34)64 (1.73)
ZnTNP-L2−6069.31 (1.41)−65 (1.71)60 (1.62)
a μ0 EF = μ0,complex0,L2. b β1907 EF = β1907,complex1907,L2. c nd = not determined. d β1907 = +35 × 10−30 esu from Reference [10]. d β|| EF = β||,complex||,L2.

Share and Cite

MDPI and ACS Style

Tessore, F.; Di Carlo, G.; Forni, A.; Righetto, S.; Limosani, F.; Orbelli Biroli, A. Second Order Nonlinear Optical Properties of 4-Styrylpyridines Axially Coordinated to A4 ZnII Porphyrins: A Comparative Experimental and Theoretical Investigation. Inorganics 2020, 8, 45. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8080045

AMA Style

Tessore F, Di Carlo G, Forni A, Righetto S, Limosani F, Orbelli Biroli A. Second Order Nonlinear Optical Properties of 4-Styrylpyridines Axially Coordinated to A4 ZnII Porphyrins: A Comparative Experimental and Theoretical Investigation. Inorganics. 2020; 8(8):45. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8080045

Chicago/Turabian Style

Tessore, Francesca, Gabriele Di Carlo, Alessandra Forni, Stefania Righetto, Francesca Limosani, and Alessio Orbelli Biroli. 2020. "Second Order Nonlinear Optical Properties of 4-Styrylpyridines Axially Coordinated to A4 ZnII Porphyrins: A Comparative Experimental and Theoretical Investigation" Inorganics 8, no. 8: 45. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8080045

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop