Next Article in Journal
CO2 Absorption Mechanism by the Nonaqueous Solvent Consisting of Hindered Amine 2-[(1,1-dimethylethyl)amino]ethanol and Ethylene Glycol
Next Article in Special Issue
Self-Assembly of Chiral Cyclohexanohemicucurbit[n]urils with Bis(Zn Porphyrin): Size, Shape, and Time-Dependent Binding
Previous Article in Journal
Effect of DODAB Nano-Sized Cationic Bilayer Fragments against Leishmania amazonensis
Previous Article in Special Issue
Supramolecular Sensing of a Chemical Warfare Agents Simulant by Functionalized Carbon Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Kinetic Investigation on Tetrakis(4-Sulfonatophenyl)Porphyrin J-Aggregates Formation Catalyzed by Cationic Metallo-Porphyrins

by
Ilaria Giuseppina Occhiuto
1,
Roberto Zagami
2,
Mariachiara Trapani
2,
Maria Angela Castriciano
2,
Andrea Romeo
1,2 and
Luigi Monsù Scolaro
1,2,*
1
Dipartimento di Scienze Chimiche, Biologiche, Farmaceutiche ed Ambientali, University of Messina and C.I.R.C.M.S.B. V.le F. Stagno D’Alcontres, 31-98166 Messina, Italy
2
CNR-ISMN Istituto per lo Studio dei Materiali Nanostrutturati c/o Dipartimento di Scienze Chimiche, Biologiche, Farmaceutiche ed Ambientali, University of Messina, V.le F. Stagno D’Alcontres, 31-98166 Messina, Italy
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(23), 5742; https://doi.org/10.3390/molecules25235742
Submission received: 18 November 2020 / Revised: 30 November 2020 / Accepted: 3 December 2020 / Published: 5 December 2020
(This article belongs to the Special Issue Self-Aggregation in Supramolecular Systems)

Abstract

:
Under mild acidic conditions, various metal derivatives of tetrakis(4-N-methylpyridinium)porphyrin (gold(III), AuT4; cobalt(III), CoT4; manganese(III), MnT4 and zinc(II), ZnT4) catalytically promote the supramolecular assembling process of the diacid 5,10,15,20-tetrakis(4-sulfonatophenyl)porphyrin (H2TPPS4) into J-aggregates. The aggregation kinetics have been treated according to a well-established model that involves the initial formation of a critical nucleus containing m porphyrin units, followed by autocatalytic growth, in which the rate evolves as a power of time. An analysis of the extinction time traces allows to obtain the rate constants for the auto-catalyzed pathway, kc, and the number of porphyrins involved in the initial seeding. The aggregation kinetics have been investigated at fixed H2TPPS4 concentration as a function of the added metal derivatives MT4. The derived rate constants, kc, obey a rate law that is first order in [MT4] and depend on the specific nature of the catalyst in the order AuT4 > CoT4 > MnT4 > ZnT4. Both resonance light scattering (RLS) intensity and extinction in the aggregated samples increase on increasing [MT4]. With the exception of AuT4, the final aggregated samples obtained at the highest catalyst concentration exhibit a negative Cotton effect in the J-band region, evidencing the occurrence of spontaneous symmetry breaking. The role of the nature of the metal derivative in terms of overall charge and presence of axial groups will be discussed.

1. Introduction

Spontaneous self-assembling of molecular units into ordered and larger stable arrangements through non-covalent interactions is a topic that in the last few years has received significant attention from the research community due to its large impact on many aspects of chemistry [1,2]. Porphyrins certainly represent a class of compounds that are extremely interesting in this regard because, depending on both their electronic and steric characteristics, they are susceptible to aggregation forming dimers and/or larger molecular oligomers [3,4,5,6]. In this respect, water-soluble porphyrins are excellent candidates to explore the aggregation process, which can be suitably modulated by varying the medium properties, such as pH and ionic strength, or by taking advantage of the coordination chemistry properties of the central macrocycle. The latter may be indeed readily metallated with a wide variety of metal ions, thereby giving access to peculiar photophysical and photochemical properties [7,8,9,10]. However, despite many reports on the formation of porphyrin homo-aggregates, there are relatively few literature investigations on hetero-aggregated systems. Self-assembly between cationic and anionic meso-substituted porphyrins, mainly driven by electrostatic attraction of peripheral substituent groups, hydrophobic interactions and axial coordination, represents an easy way to design hetero-aggregated supramolecular species [11,12,13,14,15,16,17], even if the formation of discrete hetero-dimers has been reported only in water/organic solvents mixtures [18,19]. In this context, the targeted use of anionic porphyrins such as 5,10,15,20-tetrakis(4-sulfonatophenyl) porphyrin TPPS4, whose diacid form is able to give both J- (edge-to-edge) and H-type (face-to-face) self-aggregates, is particularly attractive since these structures exhibit peculiar emergent photophysical properties tunable through aggregate topology and molecular arrangement [20,21,22,23,24,25,26,27,28,29,30]. In the literature, some examples concerning the use of TPPS4 in the formation of hetero-aggregated systems have been reported [31,32,33,34,35,36,37] and, among them, some focus on the possibilities of using TPPS4 in applicative fields. Shelnutt et al. [35] described the formation of nanotubular J-aggregates whose structure retains own individual porphyrin properties as photocatalytic reduction of metals, while Oztek et al. reported on the potential use of hetero-aggregates as hydrogen storage materials [36].
Other relevant examples in the field involve the additional use of suitable chiral templating units aimed to obtain aggregates exhibiting a well-defined stoichiometric ratio that are able to memorize and amplify the chiral information [33,34,38,39,40,41]. Despite the many reports on thermodynamic and spectroscopic characterization of these supramolecular species, kinetic studies are rather scarce due to the difficulty in controlling the various parameters that influence the aggregation pathways and eventually the final structures at the nano- or mesoscale [42,43]. Notwithstanding, the knowledge of the kinetic behavior is a prerequisite to understanding the factors controlling the rate determining step and the early stages in these complex phenomena. On the basis of our previous spectroscopic investigations on the formation of hetero-oligomers and the subsequent J-aggregates induction [31], here, we describe a detailed kinetic study on the role of cationic metallo-porphyrins MT4 in the supramolecular assembling process leading to J-aggregates of TPPS4 (Scheme 1). Both MT4 and TPPS4 porphyrins do not show any tendency towards self-aggregation in the adopted experimental conditions and we anticipate that: (i) the addition of MT4 to solutions of TPPS4 easily promotes J-aggregation, even under mild acidic conditions; (ii) the kinetic rates and the extent of aggregation depend linearly on the cationic porphyrin concentration and on the overall characteristics of these metal derivatives (charge and axial substituent groups), following the order AuT4 > CoT4 > MnT4 > ZnT4; (iii) in most cases, the number of porphyrins in the critical nucleus is slightly less than 3-4, as previously reported for the uncatalyzed process; and iv) except for the most effective metal derivative (AuT4), small CD spectra are always detectable, thus adding to the growing number of examples of spontaneous symmetry breaking in J-aggregates [44,45,46,47,48].

2. Results and Discussion

At neutral or alkaline pH, aqueous solutions of the parent free-base porphyrin at micro-molar concentration stabilize toward self-aggregation. The nitrogen atoms at the core of the tetra-anionic porphyrin TPPS4 are protonated (pKa = 4.9 [49]) and this species is fully converted to its diacid dianionic form H2TPPS4 at pH 2.8 (B-band at 434 nm, see black line in Figure 1). In the presence of only HCl at this pH, the formation of J-aggregates of H2TPPS4 is kinetically very slow and a tiny concentration can be detected after at least 24 h. In order to investigate the catalytic role of the various metal derivatives of the cationic tetrakis(4-N-methylpyridinium)porphyrin (MT4), these latter species have been pre-incubated at different molar ratios R with TPPS4 in water (R = [TPPS4]/[MT4], with R spanning in the range 2 ÷ 30) and then aggregation has been triggered by lowering the pH to 2.8 through addition of HCl. All the investigated MT4 derivatives form hetero-aggregates with the anionic TPPS4 and their spectroscopic features and stoichiometry have been reported in literature [31]. Figure 1 displays the typical time evolution of the UV/Vis spectra observed for both CoT4 (Figure 1, left) and AuT4 (Figure 1, right) at the highest ratio R = 2. The gold(III) metal complex interacting with H2TPPS4 exhibits an initial B-band at 405 nm that decreases in intensity and undergoes a further bathochromic shift to 411 nm, matched by the decrease of the B- and Q-bands of the diacid H2TPPS4 at 434 and 644 nm, respectively, and the concomitant increase of the J-aggregate B-band at 491 nm, together with its Q-band at 704 nm. MnT4 display a B-band at 462 nm and it undergoes a bathochromic shift of 6 nm upon aggregation (see Supporting Information, Figure S7). In the case of CoT4, the B-band of this species (434 nm) is under the envelope of the diacid H2TPPS4 and consequently, not detectable. ZnT4 exhibits a B-band at 436 nm and follows the same pattern of CoT4. Upon addition of HCl, a decrease of the B-band of the diacid species also occurs, while the J-band at 490 nm increases in intensity.
The kinetics of growth of J-aggregates can be easily followed by monitoring the extinction of the samples at 490 nm. Figure 2 shows a typical kinetic profile for the aggregation induced by CoT4. As already reported for acid-induced aggregation of H2TPPS4, a sigmoidal behavior is clearly observed with a quite short initial incubation or lag-time followed by an exponential growth. These data have been treated by a well-established autocatalytic model, in which two independent pathways are operative: (i) an uncatalyzed one, controlled by a rate k0 and (ii) a catalyzed one, dominated by a stretched exponential form with kc as the rate constant and n as the time exponent. The rate determining step of the second pathway is the initial formation of a nucleus containing m monomeric porphyrin units [50,51]. All the kinetic parameters have been obtained by a non-linear best-fitting procedure that was applied to the experimental extinction data and collected in Table S1 (see Supporting Information), with the exception of k0, which makes no or only a minor contribution to the overall fit of the data.
As an example, Figure 3 (left) displays the typical linear dependence of the values of the rate constant kc as a function of the increasing concentration of [CoT4]. The related values for the time exponent n range between 3 and 4, while the size of the initial nucleus m ≈ 2–3 (dimer or trimer) is slightly lower with respect to what is reported in the literature for acid-induced aggregation of the parent porphyrin. In these latter cases, m values suggested the involvement of a trimer or a tetramer in the rate determining step [47,52,53,54]. This difference could be ascribed to the different nature of the starting building block for the aggregation process, where a heterodimer stabilized by electrostatic interactions could be responsible for the seeding of the growth. As already reported in the literature [31], the cationic porphyrins investigated here form such kinds of supramolecular species with variable stoichiometry. This observation, together with a certain degree of variation on m values measured for ZnT4 and AuT4 (see Table S1), suggests the potential formation of concentration dependent hetero-oligomers as initiators for the catalytic growth of the H2TPPS4 J-aggregates.
Resonance light scattering from aggregated samples confirms the presence of quite large arrays of electronically coupled porphyrin units (N > 25) [55]. The intensity of a resonance light scattering (RLS) peak depends on the scattering and the absorption cross-sections of the aggregates and therefore on their size and concentration [56]. Figure 4 shows the RLS spectra of aggregated samples for the case of the CoT4 catalyzed process. The intensities of the peaks centered at about 490 nm increase on increasing [CoT4].
Analogously, both the intensity of RLS corrected for the extinction of the solutions and the extinction increase linearly on increasing [CoT4] (Figure 5), in line with an increment in size and concentration of the formed aggregates. Since the enhancement of the scattered light at resonance with the absorption feature reflects the strong electronic coupling among the porphyrins and increases with their numbers in the aggregate [55], this evidence points to a progressive stabilization of the nano-assemblies mediated by the cationic porphyrins.
Interestingly, the intensity of RLS increases on increasing the aggregation rate constant kc (Figure 6). The data have been analyzed through a linear best fitting, obtaining IRLScorr = (−912 ± 905) + 105 × (7.53 ± 0.56) × kc. On one hand, this behavior is in contrast with that observed for the acid-induced aggregation of TPPS4, where an inverse dependence has been reported for the growth of porphyrin nanotubes [47]. On the other hand, when these nanoassemblies are formed in the presence of a high concentration of Zn2+, again a linear dependence between these two parameters has been reported [52]. These findings suggest that the structural arrangement of the chromophores in these specific supramolecular species are quite different with respect to that involved in nanotubes. Further evidence for a difference in the aggregate architecture stems from the slight broadening of the J-band for all the final aggregated samples. These spectral features appear generally larger than those reported for systems where the Frenkel exciton model for the electronic coupling among chromophores is operative [57].
When comparing the values of kc for the various metal derivatives, a monotonic increase is well evident together with a different efficiency in promoting the aggregation of TPPS4. Figure 7 reports the behavior of kc as a function of increasing [MT4] for the investigated cationic porphyrins. All the data have been best-fitted to the equation kc = kMT4 × [MT4] and the values of the relative slope kMT4 follow the trend: AuT4 > CoT4 > MnT4 > ZnT4 (Figure 8, left). The ability of the various porphyrins in accelerating the formation of J-aggregates is paralleled by their attitude in stabilizing larger aggregates or higher concentrations, as shown in Figure 8 (right), where the values of IRLScorr and extinction for fully aggregated samples at the highest investigated concentration are displayed.
An intriguing property of TPPS4 J-aggregates is their propensity to spontaneous symmetry breaking, leading to an unbalance in the mixture of enantiomorphous nanostructures. This occurrence can be detected through circular dichroism spectroscopy and these supramolecular assemblies exhibit exciton coupled CD spectra at the wavelength of their absorption bands. In the case of J-aggregates formed in the presence of inorganic acids, the intensity of the usually observed positive Cotton effect can be quite large and an inverse dependence on the aggregation rate has been reported [46,47]. In the present case, very weak negative exciton split CD spectra are detectable at the J-band for all the samples obtained in the presence of the cationic porphyrins, with the exception of AuT4 (Figure 9). This evidence supports the hypothesis that the presence of MT4 species promotes structurally different aggregates with respect to nanotubes.
On the basis of the collected evidence, a mechanistic pathway to aggregation can be proposed (Scheme 2). In water at neutral pH, the TPPS4 porphyrin is not protonated and therefore it is a tetra-anion able to strongly interact with cationic MT4 metallo-porphyrins, leading to different hetero-aggregates (MT4@TPPS4). These latter species possess an overall charge that depends on the coordinated metal ion (Au(III) +5; Co(III) +5; Mn(III) +5; Zn(II) +4) in the cationic species. Further, the nature of the metal ion imposes different coordination numbers and geometries: (i) Au(III) 4-coordinated and square planar; (ii) Co(III) and Mn(III) 6-coordinated and octahedral; and (iii) Zn(II) 5-coordinated and square pyramidal. The resulting different charge and geometrical structures of the various metal complexes have an impact on the stoichiometry and interaction among porphyrins in the formation of the hetero-aggregated species. Moreover, upon lowering the pH to 2.8, the TPPS4 porphyrin is further protonated at the macrocyclic core and its overall charge decreases to −2 in the diacid H2TPPS4. This event determines a rearrangement of the hetero-aggregates (MT4@H2TPPS4) and their stoichiometry were reported as ranging from 1:3 to 2:3 for the AuT4@H2TPPS4 species and 2:3 for MnT4@H2TPPS4 and ZnT4@H2TPPS4 [31]. The preference for an edge-to-edge interaction (J-type) in these hetero-oligomers was explained (i) in terms of a steric role of the axial substituent groups, as in the penta- (ZnT4) and hexa-coordinated (CoT4 and MnT4) complexes, thus preventing a potential face-to-face contact (H-type) and (ii) the net positive charge at the core of the macrocycle (AuT4), leading to a repulsive interaction with the diprotonated H2TPPS4. All these factors determine the observed trend in the efficiency of the MT4 species to influence both the rates and the eventual extent of aggregation. Indeed, these J-hetero-oligomers can be considered as the real initiators of the assembling process. Alternatively, the further addition of H2TPPS4 monomers determines the formation of the critical nucleus (MT4@(H2TPPS4)m). From this point on, the autocatalytic pathway controlled by the rate constants kc leads to the growth of the final J-aggregates. It is interesting to observe that, when detectable, the final position of the B-band related to the MT4 species in the UV/Vis spectra suggests their involvement into the nano-assemblies (Figure S7). In addition, slight bathochromic shifts can be measured in the B-and Q-bands of the final J-aggregates (AuT4, 492/704 nm; CoT4, 492/709 nm; MnT4 493/709 nm; ZnT4, 496/712 nm), supporting the hypothesis of a substantial incorporation of the cationic metallo-porphyrins in the J-aggregates. This evidence is in line with literature data showing the incorporation of Sn(IV) cationic species into nanotubes of H2TPPS4 [35].

3. Materials and Methods

Materials. 5, 10, 15, 20-tetrakis(4-sulfonatophenyl)porphine (TPPS4) was purchased from Aldrich Chemical Co. Au(III), Mn(III), Co(III), Zn(II) derivatives of tetrakis(4-N-methylpyridyl)porphine (H2T4) were prepared according to literature procedures [7,8,9,10]. These porphyrins were solubilized in high-purity doubly distilled water from Galenica Senese. The range of concentration used in our experiments was determined spectrophotometrically using the molar extinction coefficients at the Soret maxima (TPPS4: 5.33 × 105 M−1cm−1, λ = 414 nm; AuT4: 2.82 × 105 M−1cm−1, λ = 403 nm; MnT4: 1.29 × 105 M−1cm−1, λ = 462 nm; CoT4: 1.68 × 105 M−1cm−1, λ = 434 nm; ZnT4: 1.81 × 105 M−1cm−1, λ = 436 nm).
Methods. UV/Vis absorption spectra and kinetic traces were measured on an Agilent model HP 8453 diode array spectrophotometer. An UV filter (Hoya glass type UV-34, cut-off: 340 nm) was used in the kinetic measurements in order to cut off the UV component of the spectrophotometer lamp, preventing the photodegradation of porphyrins. Resonance light scattering (RLS) experiments were performed on a Jasco mod. FP-750 spectrofluorimeter, adopting a synchronous scan protocol with a right angle geometry [56]. Circular dichroism experiments were carried out on a Jasco mod. J-710 spectropolarimeter. Aggregation occurs by addition of a known volume of a concentrated stock solution of hydrochloric acid (final concentration [HCl] = 1.58 mM, pH = 2.8) to a solution of porphyrin (3 µM) premixed with a solution of metal cationic porphyrins (final concentration [MT4] = 0.1, 0.3, 0.4, 0.5, 1, 1.5 µM). Kinetic experiments were run by acquiring extinction spectra from the mixed solutions placed in the thermostatic sample holder of the spectrophotometer (298 K). The aggregation kinetics exhibit a sigmoidal profile that can be fitted by an autocatalytic model [47,50,51]. The extinction traces collected at 490 nm were analyzed by a non-linear least square fit to the equation:
Extt = Ext + (Ext0 − Ext) (1 + (m – 1){k0t + (n + 1)−1 (kc t)n+1})−1/(m−1)
where Ext0, Ext, k0, kc, m, and n are the optimized parameters.

4. Conclusions

Kinetic investigations offer the opportunity to gain insights on how the growth of supramolecular assemblies is controlled by specific changes of the medium and other external factors. The J-aggregates formed by the water sulfonated porphyrin TPPS4 have been highly investigated for their ability to self-assemble into a variety of nano- and mesoscopic supramolecular structures. The specific arrangement and electronic coupling of the chromophores inside these species control their spectroscopic features and properties, such as their chirality. Understanding the role of other chemical components able to mediate, catalyze, and eventually to be incorporated into the final aggregates is therefore very important. The present study adds another small piece of information on these systems. Here, we demonstrated that MT4 porphyrins are effective in accelerating the J-aggregation of H2TPPS4 and that their efficiency is strongly dependent on their electronic and steric characteristics. Cationic metallo-porphyrins enter as part of the nano-architecture and the presence of transition metal ions could potentially introduce novel properties, such as catalytic activity.

Supplementary Materials

The following are available online. Figure S1: Autocatalytic rate constants kc and the values of m and n for the aggregation of TPPS4 vs. [AuT4], Figure S2: RLS intensity and extinction vs. [AuT4], Figure S3: Autocatalytic rate constants kc and the values of m and n for the aggregation of TPPS4 vs. [MnT4], Figure S4: RLS intensity and extinction vs. [MnT4], Figure S5: Autocatalytic rate constants kc and the values of m and n for the aggregation of TPPS4 as function of [ZnT4], Figure S6: RLS intensity and extinction vs. [ZnT4], Figure S7: UV/Vis extinction spectra of TPPS4 J-aggregates in the presence of AuT4, CoT4, ZnT4 and MnT4, Table S1: Kinetic parameters for TPPS4 aggregation vs. [MT4].

Author Contributions

Conceptualization, L.M.S., M.A.C. and A.R.; experiments, I.G.O., and R.Z.; formal analysis, M.T.; data curation, M.A.C.; writing—original draft preparation, L.M.S.; writing—review and editing, all Authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors thank University of Messina and CNR for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vantomme:, G.; Meijer, E.W. The construction of supramolecular systems. Science 2019, 363, 1396–1397. [Google Scholar] [CrossRef] [PubMed]
  2. Mattia, E.; Otto, S. Supramolecular systems chemistry. Nat. Nanotechnol. 2015, 10, 111–119. [Google Scholar] [CrossRef] [PubMed]
  3. Alessio, E. Non-Covalent Multi-Porphyrin Assemblies Synthesis and Properties; Springer: Berlin/Heidelberg, Germany, 2006. [Google Scholar]
  4. Magna, G.; Monti, D.; Di Natale, C.; Paolesse, R.; Stefanelli, M. The Assembly of Porphyrin Systems in Well-Defined Nanostructures: An Update. Molecules 2019, 24, 4307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Shee, N.K.; Kim, M.K.; Kim, H.-J. Supramolecular Porphyrin Nanostructures Based on Coordination-Driven Self-Assembly and Their Visible Light Catalytic Degradation of Methylene Blue Dye. Nanomaterials 2020, 10, 2314. [Google Scholar] [CrossRef] [PubMed]
  6. Stefanelli, M.; Mandoj, F.; Magna, G.; Lettieri, R.; Venanzi, M.; Paolesse, R.; Monti, D. The Self-Aggregation of Porphyrins with Multiple Chiral Centers in Organic/Aqueous Media: The Case of Sugar- and Steroid-Porphyrin Conjugates. Molecules 2020, 25, 4544. [Google Scholar] [CrossRef] [PubMed]
  7. Abou-Gamra, Z.; Harriman, A.; Neta, P. Redox chemistry of gold(III) porphyrins in water. J. Chem. Soc. Far. Transact. 1986, 82, 2337–2350. [Google Scholar] [CrossRef]
  8. Adler, A.D.; Longo, F.R.; Kampas, F.; Kim, J. On the preparation of metalloporphyrins. J. Inorg. Nucl. Chem. 1970, 32, 2443–2445. [Google Scholar] [CrossRef]
  9. Harriman, A.; Porter, G. Photochemistry of manganese porphyrins. Part 1.—Characterisation of some water soluble complexes. J. Chem. Soc. Far. Transact. 2 1979, 75, 1532–1542. [Google Scholar] [CrossRef]
  10. Lin, M.; Marzilli, L.G. Solution Chemistry of Cobalt(III) Porphyrins in Water and Nonaqueous Solvents. Axial Ligation by Solvent and Counterion. Inorg. Chem. 1994, 33, 5309–5315. [Google Scholar] [CrossRef]
  11. Monti, D.; Venanzi, M.; Stefanelli, M.; Sorrenti, A.; Mancini, G.; Di Natale, C.; Paolesse, R. Chiral Amplification of Chiral Porphyrin Derivatives by Templated Heteroaggregation. J. Am. Chem. Soc. 2007, 129, 6688–6689. [Google Scholar] [CrossRef]
  12. Segawa, H.; Nishino, H.; Kamikawa, T.; Honda, K.; Shimidzu, T. Hetero-aggregation between Gold Porphyrins and Zinc Porphyrins through Charge Transfer Interaction. Chem. Lett. 1989, 18, 1917–1920. [Google Scholar] [CrossRef]
  13. Mazur, U.; Hipps, K.W. A Systematic Approach toward Designing Functional Ionic Porphyrin Crystalline Materials. J. Phys. Chem. C 2018, 122, 22803–22820. [Google Scholar] [CrossRef]
  14. Tian, Y.; Busani, T.; Uyeda, G.H.; Martin, K.E.; van Swol, F.; Medforth, C.J.; Montaño, G.A.; Shelnutt, J.A. Hierarchical cooperative binary ionic porphyrin nanocomposites. Chem. Commun. 2012, 48, 4863–4865. [Google Scholar] [CrossRef] [PubMed]
  15. Medforth, C.J.; Wang, Z.; Martin, K.E.; Song, Y.; Jacobsen, J.L.; Shelnutt, J.A. Self-assembled porphyrin nanostructures. Chem. Commun. 2009, 47, 7261–7277. [Google Scholar] [CrossRef] [PubMed]
  16. Tian, Y.; Martin, K.E.; Shelnutt, J.Y.T.; Evans, L.; Busani, T.; Miller, J.E.; Medforth, C.J.; Shelnutt, J.A. Morphological families of self-assembled porphyrin structures and their photosensitization of hydrogen generation. Chem. Commun. 2011, 47, 6069–6071. [Google Scholar] [CrossRef]
  17. Shimidzu, T.; Iyoda, T. Accordion-type aggregate of water-soluble meso-tetraphenylporphyrin derivatives. Chem. Lett. 1981, 10, 853–856. [Google Scholar] [CrossRef]
  18. Ojadi, E.; Selzer, R.; Linschitz, H. Properties of porphyrin dimers, formed by pairing cationic and anionic porphyrins. J. Am. Chem. Soc. 1985, 107, 7783–7784. [Google Scholar] [CrossRef]
  19. Endisch, C.; Fuhrhop, J.-H.; Buschmann, J.; Luger, P.; Siggel, U. β-Tetraethyl-β′-tetrapyridin-4-yl Porphyrins, Their N-Methylated Tetracations, and Heterodimers with ms-Tetraphenylsulfonato Porphyrins. J. Am. Chem. Soc. 1996, 118, 6671–6680. [Google Scholar] [CrossRef]
  20. Schwab, A.D.; Smith, D.E.; Bond-Watts, B.; Johnston, D.E.; Hone, J.; Johnson, A.T.; de Paula, J.C.; Smith, W.F. Photoconductivity of Self-Assembled Porphyrin Nanorods. Nano Lett. 2004, 4, 1261–1265. [Google Scholar] [CrossRef]
  21. Collini, E.; Ferrante, C.; Bozio, R.; Lodi, A.; Ponterini, G. Large third-order nonlinear optical response of porphyrin J-aggregates oriented in self-assembled thin films. J. Mater. Chem. 2006, 16, 1573–1578. [Google Scholar] [CrossRef]
  22. Collini, E.; Ferrante, C.; Bozio, R. Strong Enhancement of the Two-Photon Absorption of Tetrakis(4-sulfonatophenyl)porphyrin Diacid in Water upon Aggregation. J. Phys. Chem. B 2005, 109, 2–5. [Google Scholar] [CrossRef] [PubMed]
  23. Kobayashi, T. J-Aggregates; World Scientific Publishing Company: Singapore, 1996; Volume 1. [Google Scholar]
  24. Kobayashi, T. J-Aggregates; World Scientific Publishing Company: Singapore, 2012; Volume 2. [Google Scholar]
  25. Micali, N.; Villari, V.; Castriciano, M.A.; Romeo, A.; Scolaro, L.M. From fractal to nanorod porphyrin J-aggregates. Concentration-induced tuning of the aggregate size. J. Phys. Chem. B 2006, 110, 8289–8295. [Google Scholar] [CrossRef] [PubMed]
  26. Castriciano, M.A.; Romeo, A.; Villari, V.; Micali, N.; Scolaro, L.M. Nanosized porphyrin J-aggregates in water/AOT/decane microemulsions. J. Phys. Chem. B 2004, 108, 9054–9059. [Google Scholar] [CrossRef]
  27. Castriciano, M.; Romeo, A.; Villari, V.; Micali, N.; Scolaro, L.M. Structural rearrangements in 5,10,15,20-tetrakis(4-sulfonatophenyl)porphyrin J-aggregates under strongly acidic conditions. J. Phys. Chem. B 2003, 107, 8765–8771. [Google Scholar] [CrossRef]
  28. Castriciano, M.A.; Leone, N.; Cardiano, P.; Manickam, S.; Scolaro, L.M.; Lo Schiavo, S. A new supramolecular polyhedral oligomeric silsesquioxanes (POSS)-porphyrin nanohybrid: Synthesis and spectroscopic characterization. Mater. Chem. C 2013, 1, 4746–4753. [Google Scholar] [CrossRef]
  29. Castriciano, M.A.; Carbone, A.; Sacca, A.; Donato, M.G.; Micali, N.; Romeo, A.; De Luca, G.; Scolaro, L.M. Optical and sensing features of TPPS4 J-aggregates embedded in Nafion membranes: Influence of casting solvents. J. Mater. Chem. 2010, 20, 2882–2886. [Google Scholar] [CrossRef]
  30. Castriciano, M.A.; Donato, M.G.; Villari, V.; Micali, N.; Romeo, A.; Scolaro, L.M. Surfactant-like Behavior of Short-Chain Alcohols in Porphyrin Aggregation. J. Phys. Chem. B 2009, 113, 11173–11178. [Google Scholar] [CrossRef]
  31. Lauceri, R.; Gurrieri, S.; Bellacchio, E.; Contino, A.; Scolaro, L.M.; Romeo, A.; Toscano, A.; Purrello, R. J-type aggregates of the anionic meso-tetrakis(4-sulfonatophenyl)porphine induced by ‘hindered’ cationic porphyrins. Supramol. Chem. 2000, 12, 193–202. [Google Scholar] [CrossRef]
  32. Micali, N.; Romeo, A.; Lauceri, R.; Purrello, R.; Mallamace, F.; Scolaro, L.M. Fractal structures in homo- and heteroaggregated water soluble porphyrins. J. Phys. Chem. B 2000, 104, 9416–9420. [Google Scholar] [CrossRef]
  33. Lauceri, R.; Fasciglione, G.F.; D’Urso, A.; Marini, S.; Purrello, R.; Coletta, M. Kinetic Investigation of Porphyrin Interaction with Chiral Templates Reveals Unexpected Features of the Induction and Self-Propagation Mechanism of Chiral Memory. J. Am. Chem. Soc. 2008, 130, 10476–10477. [Google Scholar] [CrossRef] [Green Version]
  34. Mammana, A.; De Napoli, M.; Lauceri, R.; Purrello, R. Induction and memory of chirality in porphyrin hetero-aggregates: The role of the central metal ion. Biorg. Med. Chem. 2005, 13, 5159–5163. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, Z.; Medforth, C.J.; Shelnutt, J.A. Porphyrin Nanotubes by Ionic Self-Assembly. J. Am. Chem. Soc. 2004, 126, 15954–15955. [Google Scholar] [CrossRef] [PubMed]
  36. Oztek, M.T.; Hampton, M.D.; Slattery, D.K.; Loucks, S. Hydrogen storage with hetero porphyrin aggregates. Int. J. Hydrogen Energy 2011, 36, 6705–6710. [Google Scholar] [CrossRef]
  37. Di Natale, C.; Monti, D.; Paolesse, R. Chemical sensitivity of porphyrin assemblies. Mater. Today 2010, 13, 46–52. [Google Scholar] [CrossRef]
  38. Purrello, R.; Scolaro, L.M.; Bellacchio, E.; Gurrieri, S.; Romeo, A. Chiral H- and J-Type Aggregates of meso-Tetrakis(4-sulfonatophenyl)porphine on a-Helical Polyglutamic Acid Induced by Cationic Porphyrins. Inorg. Chem. 1998, 37, 3647–3648. [Google Scholar] [CrossRef]
  39. Purrello, R.; Raudino, A.; Scolaro, L.M.; Loisi, A.; Bellacchio, E.; Lauceri, R. Ternary porphyrin aggregates and their chiral memory. J. Phys. Chem. B 2000, 104, 10900–10908. [Google Scholar] [CrossRef]
  40. Bellacchio, E.; Lauceri, R.; Gurrieri, S.; Scolaro, L.M.; Romeo, A.; Purrello, R. Template-imprinted chiral porphyrin aggregates. J. Am. Chem. Soc. 1998, 120, 12353–12354. [Google Scholar] [CrossRef]
  41. Lauceri, R.; Raudino, A.; Scolaro, L.M.; Micali, N.; Purrello, R. From Achiral Porphyrins to Template-Imprinted Chiral Aggregates and Further. Self-Replication of Chiral Memory from Scratch. J. Am. Chem. Soc. 2002, 124, 894–895. [Google Scholar] [CrossRef]
  42. Monsu Scolaro, L.; Castriciano, M.; Romeo, A.; Mazzaglia, A.; Mallamace, F.; Micali, N. Nucleation effects in the aggregation of water-soluble porphyrin aqueous solutions. Physica A 2002, 304, 158–169. [Google Scholar] [CrossRef]
  43. Mallamace, F.; Monsu’ Scolaro, L.; Romeo, A.; Micali, N. Crossover in the Kinetic Growth Process of Porphyrin Aggregation. Phys. Rev. Lett. 1999, 82, 3480–3483. [Google Scholar] [CrossRef]
  44. Short, J.M.; Berriman, J.A.; Kubel, C.; El-Hachemi, Z.; Naubron, J.V.; Balaban, T.S. Electron Cryo-Microscopy of TPPS4 center dot 2HCl Tubes Reveals a Helical Organisation Explaining the Origin of their Chirality. Chem. Phys. Chem. 2013, 14, 3209–3214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Romeo, A.; Castriciano, M.A.; Zagami, R.; Pollicino, G.; Monsu Scolaro, L.; Pasternack, R.F. Effect of zinc cations on the kinetics for supramolecular assembling and the chirality of porphyrin J-aggregates. Chem. Sci. 2017, 8, 961–967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Occhiuto, I.G.; Zagami, R.; Trapani, M.; Bolzonello, L.; Romeo, A.; Castriciano, M.A.; Collini, E.; Monsu Scolaro, L. The role of counter-anions in the kinetics and chirality of porphyrin J-aggregates. Chem. Commun. 2016, 52, 11520–11523. [Google Scholar] [CrossRef] [PubMed]
  47. Romeo, A.; Castriciano, M.A.; Occhiuto, I.; Zagami, R.; Pasternack, R.F.; Scolaro, L.M. Kinetic Control of Chirality in Porphyrin J-Aggregates. J.Am. Chem. Soc. 2014, 136, 40–43. [Google Scholar] [CrossRef] [PubMed]
  48. El-Hachemi, Z.; Escudero, C.; Acosta-Reyes, F.; Casas, M.T.; Altoe, V.; Aloni, S.; Oncins, G.; Sorrenti, A.; Crusats, J.; Campos, J.L.; et al. Structure vs. properties—Chirality, optics and shapes—in amphiphilic porphyrin J-aggregates. J. Mater. Chem. C 2013, 1, 3337–3346. [Google Scholar] [CrossRef] [Green Version]
  49. Kalyanasundaram, K. Photochemistry of water-soluble porphyrins: Comparative study of isomeric tetrapyridyl- and tetrakis(N-methylpyridiniumyl)porphyrins. Inorg. Chem. 1984, 32, 2453–2459. [Google Scholar] [CrossRef]
  50. Pasternack, R.F.; Fleming, C.; Herring, S.; Collings, P.J.; de Paula, J.; DeCastro, G.; Gibbs, E.J. Aggregation kinetics of extended porphyrin and cyanine dye assemblies. Biophys. J. 2000, 79, 550–560. [Google Scholar] [CrossRef] [Green Version]
  51. Pasternack, R.F.; Gibbs, E.J.; Collings, P.J.; de Paula, J.C.; Turzo, L.C.; Terracina, A. A nonconventional approach to supramolecular formation dynamics. The kinetics of assembly of DNA-bound porphyrins. J. Am. Chem. Soc. 1998, 120, 5873–5878. [Google Scholar] [CrossRef]
  52. Occhiuto, I.G.; Castriciano, M.A.; Trapani, M.; Zagami, R.; Romeo, A.; Pasternack, R.F.; Monsù Scolaro, L. Controlling J-Aggregates Formation and Chirality Induction through Demetallation of a Zinc(II) Water Soluble Porphyrin. Intern. J. Mol. Sci. 2020, 21, 4001. [Google Scholar] [CrossRef]
  53. Zagami, R.; Romeo, A.; Castriciano, M.A.; Monsù Scolaro, L. Inverse Kinetic and Equilibrium Isotope Effects on Self-Assembly and Supramolecular Chirality of Porphyrin J-Aggregates. Chem. Eur. J. 2017, 23, 70–74. [Google Scholar] [CrossRef]
  54. Castriciano, M.A.; Romeo, A.; Zagami, R.; Micali, N.; Scolaro, L.M. Kinetic effects of tartaric acid on the growth of chiral J-aggregates of tetrakis(4-sulfonatophenyl)porphyrin. Chem. Commun. 2012, 48, 4872–4874. [Google Scholar] [CrossRef] [PubMed]
  55. Parkash, J.; Robblee, J.H.; Agnew, J.; Gibbs, E.; Collings, P.; Pasternack, R.F.; de Paula, J.C. Depolarized resonance light scattering by porphyrin and chlorophyll a aggregates. Biophys. J. 1998, 74, 2089–2099. [Google Scholar] [CrossRef] [Green Version]
  56. Pasternack, R.F.; Collings, P.J. Resonance Light-Scattering—A New Technique for Studying Chromophore Aggregation. Science 1995, 269, 935–939. [Google Scholar] [CrossRef] [PubMed]
  57. Frenkel, J. On the Transformation of light into Heat in Solids. I. Phys. Rev. 1931, 37, 17–44. [Google Scholar] [CrossRef]
Scheme 1. Pictorial sketch of TPPS4 protonation and subsequent J-aggregation in the presence of cationic metallo-porphyrins MT4.
Scheme 1. Pictorial sketch of TPPS4 protonation and subsequent J-aggregation in the presence of cationic metallo-porphyrins MT4.
Molecules 25 05742 sch001
Figure 1. UV/Vis extinction spectral changes for the aggregation of TPPS4 into J-aggregates in the presence of CoT4 (left) and AuT4 (right) after the addition of hydrochloric acid. For CoT4: 0 (black), 40 (red), 80 (blue), 120 (green), 160 (violet), 210 (yellow), 340 s (cyano); For AuT4: 0 (black), 10 (red), 20 (blue), 30 (green), 40 (violet), 50 (yellow), 80 (cyano), 220 s (brown). (Experimental conditions: [TPPS4] = 3 μM; [MT4] = 1.5 μM; [HCl] = 0.00158 M, T = 298 K, total acquisition time 400 s).
Figure 1. UV/Vis extinction spectral changes for the aggregation of TPPS4 into J-aggregates in the presence of CoT4 (left) and AuT4 (right) after the addition of hydrochloric acid. For CoT4: 0 (black), 40 (red), 80 (blue), 120 (green), 160 (violet), 210 (yellow), 340 s (cyano); For AuT4: 0 (black), 10 (red), 20 (blue), 30 (green), 40 (violet), 50 (yellow), 80 (cyano), 220 s (brown). (Experimental conditions: [TPPS4] = 3 μM; [MT4] = 1.5 μM; [HCl] = 0.00158 M, T = 298 K, total acquisition time 400 s).
Molecules 25 05742 g001
Figure 2. Typical kinetic profile of extinction at 491 nm vs. time for the aggregation of TPPS4 catalyzed by CoT4. The black line is the best fit of the experimental data to Equation (1) (Experimental conditions: [TPPS4] = 3 μM; [CoT4] = 1.5 μM; [HCl] = 0.00158 M, T = 298 K).
Figure 2. Typical kinetic profile of extinction at 491 nm vs. time for the aggregation of TPPS4 catalyzed by CoT4. The black line is the best fit of the experimental data to Equation (1) (Experimental conditions: [TPPS4] = 3 μM; [CoT4] = 1.5 μM; [HCl] = 0.00158 M, T = 298 K).
Molecules 25 05742 g002
Figure 3. Plot of the autocatalytic rate constants kc (s−1) (left), the line represents the linear fit of the experimental data to the equation kc = (1.7 × 10−3 ± 5 × 10−4) + (1.88 × 10−2 ± 6 × 10−4) × [CoT4], and the values of m (full squares) and n (empty squares) for the aggregation of TPPS4 as a function of [CoT4] (right). (Experimental conditions: [TPPS4] = 3 μM; [HCl] = 0.00158 M, T = 298 K).
Figure 3. Plot of the autocatalytic rate constants kc (s−1) (left), the line represents the linear fit of the experimental data to the equation kc = (1.7 × 10−3 ± 5 × 10−4) + (1.88 × 10−2 ± 6 × 10−4) × [CoT4], and the values of m (full squares) and n (empty squares) for the aggregation of TPPS4 as a function of [CoT4] (right). (Experimental conditions: [TPPS4] = 3 μM; [HCl] = 0.00158 M, T = 298 K).
Molecules 25 05742 g003
Figure 4. Resonance light scattering (RLS) spectra of the samples at the end of the aggregation process catalyzed by CoT4 at different concentrations of this metal derivative. (Experimental conditions: [TPPS4] = 3 μM; [HCl] = 0.00158 M, [CoT4] = 0.1 μM (black), 0.3 μM (red), 0.4 μM (blue), 0.5 μM (green), 1 μM (violet), 1.5 μM (yellow), T = 298 K).
Figure 4. Resonance light scattering (RLS) spectra of the samples at the end of the aggregation process catalyzed by CoT4 at different concentrations of this metal derivative. (Experimental conditions: [TPPS4] = 3 μM; [HCl] = 0.00158 M, [CoT4] = 0.1 μM (black), 0.3 μM (red), 0.4 μM (blue), 0.5 μM (green), 1 μM (violet), 1.5 μM (yellow), T = 298 K).
Molecules 25 05742 g004
Figure 5. Intensity of RLS spectra corrected for extinction (left) and extinction of the samples at the end of the TPPS4 aggregation process catalyzed by CoT4 (right) as a function of the concentration of this metal derivative. (Experimental conditions: [TPPS4] = 3 μM; [HCl] = 0.00158 M, T = 298 K). The lines represent the linear best fits to the experimental data (IRLScorr = 104 × (0.035 ± 0.074) +104 × (1.419 ± 0.093) × [CoT4]; Ext = (0.075 ± 0.022) + (0.313 ± 0.028) × [CoT4]).
Figure 5. Intensity of RLS spectra corrected for extinction (left) and extinction of the samples at the end of the TPPS4 aggregation process catalyzed by CoT4 (right) as a function of the concentration of this metal derivative. (Experimental conditions: [TPPS4] = 3 μM; [HCl] = 0.00158 M, T = 298 K). The lines represent the linear best fits to the experimental data (IRLScorr = 104 × (0.035 ± 0.074) +104 × (1.419 ± 0.093) × [CoT4]; Ext = (0.075 ± 0.022) + (0.313 ± 0.028) × [CoT4]).
Molecules 25 05742 g005
Figure 6. Plot of the RLS intensity at maxima corrected for the extinction of the samples as a function of the corresponding rate constants kc for the J-aggregation catalyzed by CoT4. (Experimental conditions: [TPPS4] = 3 μM; [CoT4] = 1.5 μM; [HCl] = 0.00158 M, T = 298 K). The line represents the linear best fit to the experimental data (IRLScorr = 105 × (−0.009 ± 0.009) + 105 (7.531 ± 0.556) × kc).
Figure 6. Plot of the RLS intensity at maxima corrected for the extinction of the samples as a function of the corresponding rate constants kc for the J-aggregation catalyzed by CoT4. (Experimental conditions: [TPPS4] = 3 μM; [CoT4] = 1.5 μM; [HCl] = 0.00158 M, T = 298 K). The line represents the linear best fit to the experimental data (IRLScorr = 105 × (−0.009 ± 0.009) + 105 (7.531 ± 0.556) × kc).
Molecules 25 05742 g006
Figure 7. Plot of the autocatalytic rate constants kc (s−1) for the aggregation of TPPS4 as a function of [MT4] for the various metal derivatives. AuT4 (black squares), CoT4 (red circles), MnT4 (blue triangles), ZnT4 (green triangles). (Experimental conditions: [TPPS4] = 3 μM; [HCl] = 0.00158 M, T = 298 K). The straight lines represent the linear fit to the rate data, according to the equation kc = kMT4 × [MT4].
Figure 7. Plot of the autocatalytic rate constants kc (s−1) for the aggregation of TPPS4 as a function of [MT4] for the various metal derivatives. AuT4 (black squares), CoT4 (red circles), MnT4 (blue triangles), ZnT4 (green triangles). (Experimental conditions: [TPPS4] = 3 μM; [HCl] = 0.00158 M, T = 298 K). The straight lines represent the linear fit to the rate data, according to the equation kc = kMT4 × [MT4].
Molecules 25 05742 g007
Figure 8. Bar plot of the rate constants k’MT4 (μM × s−1) obtained as slope from Figure 7 (left) and the RLS intensity at maxima corrected for the extinction (orange bars) and extinction (green bars) of the samples after complete aggregation of TPPS4 for the various metal derivatives MT4 (right). (Experimental conditions: [TPPS4] = 3 μM; [MT4] = 1.5 μM; [HCl] = 0.00158 M, T = 298 K).
Figure 8. Bar plot of the rate constants k’MT4 (μM × s−1) obtained as slope from Figure 7 (left) and the RLS intensity at maxima corrected for the extinction (orange bars) and extinction (green bars) of the samples after complete aggregation of TPPS4 for the various metal derivatives MT4 (right). (Experimental conditions: [TPPS4] = 3 μM; [MT4] = 1.5 μM; [HCl] = 0.00158 M, T = 298 K).
Molecules 25 05742 g008
Figure 9. CD spectra of the samples after complete aggregation of TPPS4 for the various metal derivatives MT4 (AuT4 (black), CoT4 (red), MnT4 (blue), ZnT4 (green). (Experimental conditions: [TPPS4] = 3 μM; [MT4] = 1.5 μM; [HCl] = 0.00158 M, T = 298 K).
Figure 9. CD spectra of the samples after complete aggregation of TPPS4 for the various metal derivatives MT4 (AuT4 (black), CoT4 (red), MnT4 (blue), ZnT4 (green). (Experimental conditions: [TPPS4] = 3 μM; [MT4] = 1.5 μM; [HCl] = 0.00158 M, T = 298 K).
Molecules 25 05742 g009
Scheme 2. Proposed mechanism for J-aggregation of TPPS4 in the presence of cationic metallo-porphyrins MT4 under mild acidic conditions.
Scheme 2. Proposed mechanism for J-aggregation of TPPS4 in the presence of cationic metallo-porphyrins MT4 under mild acidic conditions.
Molecules 25 05742 sch002
Sample Availability: Samples of the investigated J-aggregates are available from the authors.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Occhiuto, I.G.; Zagami, R.; Trapani, M.; Castriciano, M.A.; Romeo, A.; Scolaro, L.M. Kinetic Investigation on Tetrakis(4-Sulfonatophenyl)Porphyrin J-Aggregates Formation Catalyzed by Cationic Metallo-Porphyrins. Molecules 2020, 25, 5742. https://doi.org/10.3390/molecules25235742

AMA Style

Occhiuto IG, Zagami R, Trapani M, Castriciano MA, Romeo A, Scolaro LM. Kinetic Investigation on Tetrakis(4-Sulfonatophenyl)Porphyrin J-Aggregates Formation Catalyzed by Cationic Metallo-Porphyrins. Molecules. 2020; 25(23):5742. https://doi.org/10.3390/molecules25235742

Chicago/Turabian Style

Occhiuto, Ilaria Giuseppina, Roberto Zagami, Mariachiara Trapani, Maria Angela Castriciano, Andrea Romeo, and Luigi Monsù Scolaro. 2020. "Kinetic Investigation on Tetrakis(4-Sulfonatophenyl)Porphyrin J-Aggregates Formation Catalyzed by Cationic Metallo-Porphyrins" Molecules 25, no. 23: 5742. https://doi.org/10.3390/molecules25235742

Article Metrics

Back to TopTop