Next Article in Journal
Two-Dimensional Capillary Electrophoresis with On-Line Sample Preparation and Cyclodextrin Separation Environment for Direct Determination of Serotonin in Human Urine
Next Article in Special Issue
Density Functional Theory Applied to Excited State Intramolecular Proton Transfer in Imidazole-, Oxazole-, and Thiazole-Based Systems
Previous Article in Journal
Synthesis of Chromone-Related Pyrazole Compounds
Previous Article in Special Issue
Nonempirical Simulations of Inhomogeneous Broadening of Electronic Transitions in Solution: Predicting Band Shapes in One- and Two-Photon Absorption Spectra of Chalcones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Theoretical Study of the N to O Linkage Photoisomerization Efficiency in a Series of Ruthenium Mononitrosyl Complexes

1
Laboratoire de Chimie et Physique Quantiques, IRSAMC, CNRS et Université Toulouse 3, 118 route de Narbonne, 31062 Toulouse, France
2
Present address: Institut de Recherche de Chimie Paris, PSL Research University, CNRS, Chimie ParisTech, 11 Rue Pierre et Marie Curie, F-75005 Paris, France
*
Authors to whom correspondence should be addressed.
Molecules 2017, 22(10), 1667; https://doi.org/10.3390/molecules22101667
Submission received: 8 September 2017 / Revised: 29 September 2017 / Accepted: 3 October 2017 / Published: 6 October 2017

Abstract

:
Ruthenium nitrosyl complexes are fascinating versatile photoactive molecules that can either undergo NO linkage photoisomerization or NO photorelease. The photochromic response of three ruthenium mononitrosyl complexes, trans-[RuCl(NO)(py)4]2+, trans-[RuBr(NO)(py)4]2+, and trans-(Cl,Cl)[RuCl2(NO)(tpy)]+, has been investigated using density functional theory and time-dependent density functional theory. The N to O photoisomerization pathways and absorption properties of the various stable and metastable species have been computed, providing a simple rationalization of the photoconversion trend in this series of complexes. The dramatic decrease of the N to O photoisomerization efficiency going from the first to the last complex is mainly attributed to an increase of the photoproduct absorption at the irradiation wavelength, rather than a change in the photoisomerization pathways.

1. Introduction

Photochromic systems are made of photoswitchable molecular building blocks, which often undergo a photoisomerization process [1,2,3]. While largely dominated by photoinduced electrocyclic reactions of organic compounds [4,5,6,7,8,9], photochromic systems have also been designed with metal complexes capable of undergoing linkage photoisomerization of ambidentate ligands (e.g., NO, RR’SO, and SO2) coordinated to a metal center [10,11,12,13,14]. In such inorganic systems, the photochemical and photophysical properties can be tuned by varying the nature of the different ligands. Photochromic molecular systems have entered the new generation of innovative functional materials with high added value. Applications are already widespread in nanosciences, biology, and photonic or optoelectronic devices as light-activated switches [15,16,17,18,19,20,21].
While many organic photochromic systems have been scrutinized from an experimental point of view, sometimes complemented by theoretical studies [22,23,24,25,26,27,28,29,30,31,32,33], studies of photochromic inorganic systems are much scarcer [10,11,12,13,14]. Owing to the difficulty of computing photochemical pathways in metal complexes [34], the number of theoretical studies devoted to the linkage photoisomerization mechanisms in these systems is unsurprisingly low [35,36,37,38,39,40,41,42]. Among these systems, ruthenium nitrosyl complexes have attracted considerable interest due to both their photochromic properties [43,44,45,46,47,48,49,50,51,52,53,54,55] and their capability to release nitric oxide [56,57,58,59,60,61,62,63,64,65,66]. Recently, the N→O linkage photoisomerization mechanism in the trans-[RuCl(NO)(py)4]2+ (where py denotes a pyridine ligand) ruthenium nitrosyl complex was investigated based on density functional theory (DFT) calculations of the lowest singlet and triplet potential energy profiles along the N→O linkage photoisomerization coordinate [41,42]. These studies brought some important insight into the mechanistic picture of the N→O linkage photoisomerization. In contrast to other photoisomerizable ruthenium complexes, for which energetically favorable routes on the lowest triplet-excited state were characterized [36,37,38], the triplet pathway for the N→O linkage photoisomerization reveals an activated and energetically uphill process, which forbids an adiabatic mechanism in this electronic excited state [41,42]. Thus, a complex sequential two-photon photoisomerization mechanism involving nonadiabatic processes was proposed [41] and confirmed experimentally very recently [54]. This mechanistic picture is supported by accurate multiconfigurational wavefunction-based calculations [67].
While a single crystal of trans-[RuCl(NO)(py)4](PF6)2·1/2H2O presents a quasi-complete photoconversion yield between the stable N-bonded nitrosyl (denoted GS) and metastable O-bonded isonitrosyl (denoted MS1) isomers under continuous irradiation in the 473–476 nm region [51,52,53], much lower efficiencies were observed for trans-[RuBr(NO)(py)4](PF6)2 and trans-(Cl,Cl)[RuCl2(NO)(tpy)](PF6) (where tpy denotes a terpyridine ligand) with yields of 46% and 8%, respectively, under comparable blue light irradiation [53]. This drastic decrease of photoconversion yield may be resulting from two main factors: (i) modifications in the potential energy surfaces (PES); and/or (ii) changes in the optical properties of the isomers involved. In the first case, the relative stabilities of the GS and MS1 isomers on the ground-state PES is a critical feature, but, also, the nonadiabatic photoisomerization pathway could be modified in a way that the photoconversion becomes less efficient. Regarding the optical properties, they could be modified in such a way that either the system does not absorb light so efficiently or the photoproduct (MS1) also absorbs in the same spectral region, inducing the backward photoisomerization.
The aim of this theoretical study is to shed light on these different aspects in order to rationalize qualitatively the trend in the photoconversion yield in this series of ruthenium nitrosyl complexes, shown in Figure 1. DFT calculations were thus performed on the isolated trans-[RuBr(NO)(py)4]2+ complex (denoted complex (2) hereafter) to determine and compare the potential energy landscape of the lowest singlet and triplet states with those of the trans-[RuCl(NO)(py)4]2+ (complex (1)) and trans-(Cl,Cl)[RuCl2(NO)(tpy)]+ (complex (3)) complexes previously studied [41,42,67]. Time-dependent DFT (TD-DFT) calculations were performed to determine the optical properties of all the relevant stable and metastable isomers involved in the photoisomerization mechanism. In the following, we present the results of these calculations and show that the main reason for the decrease of photoconversion efficiency in this series of complexes involves an increase of the photoproduct absorption at the irradiation wavelength rather than a change in the photoisomerization pathways.

2. Results

2.1. Overview on the Mechanistic Picture for the trans-[RuCl(NO)(py)4]2+ Complex

Figure 2 summarizes the mechanistic picture that was derived from a previous DFT study of complex (1) [41]. Upon blue light irradiation, the most stable N-bonded ground-state isomer GS, in which the Cl, Ru, N, and O atoms are collinear, is promoted to excited singlet metal-to-ligand charge transfer (1MLCT) states. These states are known to quickly relax by intersystem crossing (ISC) due to significant spin-orbit coupling (SOC) and by internal conversion (IC) to lower triplet MLCT excited states (3MLCT) [67,68]. Assuming that the lowest 3MLCT is mainly populated, the system then relaxes to a N-bonded triplet state (denoted 3GS) in which the Ru–N–O is bent. A large barrier (ca. 1 eV) on the adiabatic triplet PES forbids the system to reach a sideways NO-bonded intermediate (denoted 3MS2). However, a more accessible route is provided by ISC at a triplet/singlet crossing point, allowing the system to decay at a sideways NO-bonded intermediate on the singlet state (denoted MS2) [67]. A key feature of this mechanism is that the MS2 metastable state can also absorb the same blue light photon as GS, as revealed by the computed electronic absorption spectra at the TD-DFT level [41,42]. This second photon absorption at MS2 triggers electronic transitions to 1MLCT states from which non-radiative decays by ISC and IC lead most probably to the population of 3MS2. From this intermediate, only a small adiabatic barrier needs to be overcome on the lowest triplet PES to reach a O-bonded triplet intermediate (denoted 3MS1), in which the Ru–O–N is bent. Another triplet/singlet crossing lying nearby allows for ISC to the O-bonded MS1 product [67], restoring the collinear conformation of the Cl, Ru, O, and N atoms. This final photoproduct does not absorb in the blue light spectral range, allowing for very high photoconversion yield.

2.2. Comparison with the trans-[RuBr(NO)(py)4]2+ and trans-(Cl,Cl)[RuCl2(NO)(tpy)]+ Complexes

2.2.1. Thermal Isomerization Pathway

The stationary points characterizing the singlet PES along the thermal isomerization pathway were located for all three complexes. They correspond to the GS, MS2, and MS1 minima and the two transition states (TS1 and TS2) connecting them, as shown in Figure 2. Relative energies are given in Table 1. All three complexes display similar potential energy profiles with small changes (<0.1 eV) in the relative energies of the three isomers. The energy barriers GS to MS2 and MS2 to MS1 are also comparable in all the complexes with deviations below 0.15 eV. Overall, these results confirm the thermal stability of the stable N-bonded isomer GS for each complex.

2.2.2. Adiabatic Photoisomerization Pathway

A similar work performed for the singlet PES was carried out for the lowest triplet PES, allowing a characterization of the adiabatic photoisomerization pathway for the three complexes. The stationary points found on the lowest triplet PES correspond to the 3GS, 3MS2, and 3MS1 minima and the two transition states (3TS1 and 3TS2) connecting them, as shown in Figure 2. Relative energies for these various structures are given in Table 2. As for the thermal ground-state isomerization pathways, the adiabatic triplet photoisomerization pathways are very similar in the three complexes, all involving the same stationary points and comparable relative energies. From a structural point of view, all the optimized triplet structures are analogous in the three complexes. The electronic nature of the lowest triplet state is very similar as well: it can be described as a 3MLCT state of Ru(d)→NO(π*) character in all three complexes. A minor difference is found in complex (2), for which the bromide ligand presents slightly higher spin densities than the chloride ligand. As far as the energy profiles are concerned, they are qualitatively similar in the three complexes, involving a large adiabatic potential energy barrier (ca. 1 eV) along the photoisomerization pathway. Thus, in all these systems, the N to O linkage photoisomerization cannot occur adiabatically. Note that the three triplet states 3GS, 3MS2, and 3MS1 are lowered in energy (by a few tenths of eV; see Table 2) relative to GS in complex (3).

2.2.3. Nonadiabatic Photoisomerization Pathway

Because of the large barrier occurring along the adiabatic triplet photoisomerization pathway, funnels for efficient non-radiative decay paths between the lowest triplet excited state and the singlet ground state were located by optimizing minimum energy crossing points (MECP) between these two states. As shown in Figure 2, four MECP were located in complex (1). The first crossing (MECP1) provides a photostabilizing pathway for GS, as the main relaxation path from this funnel is expected to regenerate the initial stable N-bonded isomer. The second one (MECP2) allows the system to undergo a partial nonadiabatic photoisomerization from GS to MS2. The third crossing (MECP3) provides a photostabilizing pathway for MS2, and the last one (MECP4) allows the system to evolve to the final O-bonded isonitrosyl photoproduct MS1. All these four funnels were also found for complexes (2) and (3). All these critical crossing points are located at similar relative energies compared to the reference complex (1), providing similar radiationless pathways for these two complexes (Table 3). The energy required to reach MECP2 is of particular importance in the nonadiabatic photoisomerization mechanism, as it represents the highest activation energy that the system has to overcome along the photoisomerization pathway. This crossing is located at 0.67 and 0.65 eV above 3GS in complexes (1) and (2), respectively. However, a slightly higher energy is required at 0.90 eV for complex (3).

2.2.4. Absorption Properties

In the proposed sequential two-photon photoisomerization mechanism described in Section 2.1, the absorption properties of GS, MS2, and MS1 are all of crucial importance. Thus, we report here the TD-DFT absorption spectra computed for each isomer of the three complexes. Figure 3, Figure 4 and Figure 5 display the spectra of GS, MS2, and MS1 for complexes (1), (2) and (3), respectively. Further details about the transition wavelengths and oscillators strengths are given in Supporting Information.
In all three complexes, the first significant absorption band of GS appears near the experimental irradiation wavelengths (400–500 nm). It is, however, noticeable that the intensity of this band is much lower in complex (3). In addition, at the first significant absorption maximum of GS around 400 nm, the absorption of MS1 vanishes in the first complex, whereas it becomes more intense going to complex (2) (ε ≈ 2600 M−1·cm−1) and to complex (3) (ε ≈ 3800 M−1·cm−1). The first absorption band of MS1 in the red region (650–700 nm) corresponds to a spectral range where the GS isomer does not absorb for any of the three complexes. As far as the absorption of MS2 is concerned, it shows two absorption bands, which strongly overlap with that of the first absorption bands of GS at 400–450 nm and of MS1 at 650–700 nm in the first complex. However, this overlap is very poor in complex (3), as MS2 weakly absorbs around 400 nm and shows no absorption at 700 nm.

3. Discussion

Our results show that the photoconversion trend experimentally observed in the family of ruthenium nitrosyl compounds presented in this work can be rationalized making use of the absorption properties of the different isomers (GS, MS2, and MS1) of each complex, rather than changes in the photoisomerization pathway. Indeed, the potential energy profiles for the three complexes are very similar and the slight changes in the relative energies of the critical points along the photoisomerization pathway cannot account for the large variation of photoconversion yields observed. As pointed out in refs. [41,42], and as recalled in Section 2.1, one of the key factors necessary for the photoreactivity of this kind of compounds is the overlap between the absorption bands of the GS and MS2 isomers. This is a requirement for the description of a sequential two-photon mechanism, where, initially, GS absorbs one photon and, subsequently, MS2 absorbs a second one. In the particular case of complex (1), which shows the highest photoconversion yield, two remarkable characteristics of the spectra permit the rationalization of the mechanism and experimental yield: a) there is a spectral region near the experimental irradiation wavelengths (400–500 nm) where an important superposition of the absorption bands of GS and MS2 exists; and b) MS1 does not absorb in that specific region. While the first property is obviously a sine qua non condition for a sequential two-photon absorption mechanism, the second property is necessary to achieve high photoconversion yields under continuous irradiation. If the system is progressively transformed from GS to MS1 without MS1 being consumed, as it does not absorb at the irradiation wavelength, a quantitative conversion of GS into MS1 can be reached.
In the case of complex (2), there is also a spectral region (400–500 nm) where an important overlap between the absorption bands of GS and MS2 exists. But in contrast to complex (1), MS1 also presents an absorption band in that same region overlapping those of GS and MS2. We believe that this is the main reason for lowering by about half the photoconversion yield upon substituting the chloride by the bromide ligand.
Another aspect that can be discussed regards the optimal irradiation wavelength used for photoconversion. During the experimental irradiation times, MS1 is formed and consumed continuously, and, depending on the chemical rates for the forward and backward isomerizations, different photoconversion yields will be attained once the photostationary state is reached. The fact that the optimal irradiation wavelength found for the complex with the bromide ligand (488–496 nm) [53] is higher with respect to the one found for the complex with the chloride ligand (473–476 nm) [51,52,53] can be explained comparing the different spectra. The optimal irradiation wavelength for that complex should be one where GS and MS2 absorb significantly compared to MS1. The important region of the spectra where both GS and MS2 isomers absorb but not MS1 in complex (1) is blue-shifted with respect to the corresponding spectral region in complex (2) (i.e., the region where GS and MS2 both absorb and MS1 shows a weak absorption).
Regarding the low photoconversion yield observed for complex (3), it can also be rationalized following the same line of thinking. As shown in Figure 5, in the spectral range of 350–450 nm, MS1 presents an important absorption band, while the isomers GS and MS2 barely absorb in that region. This is consistent with the very low photoconversion yield observed for this complex. Moreover, it appears that, for this particular complex, there is no wavelength that would trigger an efficient photoisomerization in the 350–500 nm range, as there is no region where both GS and MS2 would absorb and MS1 would not. In addition, note that for this complex, a slightly higher 3GSMS2 barrier (via MECP2) is encountered.

4. Computational Details

Gas-phase geometry optimizations of all the stationary points found on the lowest singlet and triplet PES were carried out using DFT with the Gaussian 09 quantum package [69], with the exception of the transition states in complex (3) (see Table 1 and Table 2), for which the Zero Temperature String (ZTS) method [70] was used [42]. The nature of these stationary points (i.e., minima or transition states) was determined by harmonic frequency analyses. These geometry optimizations were performed using the standard hybrid functional B3LYP [71], including Grimme’s dispersion correction [72] (denoted B3LYP-D3), with a double-ζ Ahlrichs-type basis set with a set of p polarization functions for the hydrogen atoms [73], a triple- ζ Ahlrichs-type basis set with one set of d polarization functions for all heavy atoms [73], and a Stuttgart relativistic effective small-core potential [74] with its associated basis set including two sets of f and one set of g polarization functions [75] for ruthenium.
The UV-Vis absorption spectra of GS, MS1, and MS2 were computed at the TD-DFT level using the TPSSh functional [76] for complexes (1) and (2) (Figure 3 and Figure 4), as in ref. [41]. For complex (3), the BHandHLYP [77] functional was used (Figure 5) as in ref. [42], along with the TPSSh functional for the sake of comparison (Figure S1 in Supporting Information). The Tamm-Dancoff approximation (TDA) [78] and the same basis sets as described above were used. In these TD-DFT calculations the resolution-of-identity (RI) approximation for hybrid functionals [79] was employed to calculate the Coulomb energy term using the Def2-TZV auxiliary basis set and the exchange term by the so-called ‘chain-of-spheres exchange’ algorithm [80]. All the TD-DFT calculations were performed with the ORCA 3.0.3 quantum package [81].

5. Conclusions

We have theoretically investigated the complex N→O linkage photoisomerization pathways along with the absorption properties of all the stable and metastable isomers involved in a series of three ruthenium nitrosyl complexes exhibiting very different photoconversion yields upon continuous blue-light irradiation. Our results unambiguously point towards an increasing absorption of the isonitrosyl photoproduct at the irradiation wavelength used to trigger the N→O linkage photoisomerization to explain the decrease of the photoconversion yield going from trans-[RuCl(NO)(py)4]2+, to trans-[RuBr(NO)(py)4]2+, and to trans-(Cl,Cl)[RuCl2(NO)(tpy)]+.

Supplementary Materials

The following are available online: Table S1: Cartesian coordinates for all the stable and metastable isomers; Table S2: TD-TPSSh absorption spectra of trans-[RuCl(NO)(py)4]2+; Table S3: TD-TPSSh absorption spectra of trans-[RuBr(NO)(py)4]2+; Table S4: TD-BHandHLYP absorption spectra of trans-(Cl,Cl)[RuCl2(NO)(tpy)]+; Figure S1: TD-TPSSh absorption spectra of trans-(Cl,Cl)[RuCl2(NO)(tpy)]+.

Acknowledgments

We acknowledge the financial support of the ITN-EJD-TCCM PhD Fellowship as part of the Marie Skłodowska-Curie Actions (European Union’s Horizon 2020 research and innovation programme under the Marie Curie Skłodowska-Curie grant agreement No.642294). This work was granted access to the HPC resources of CALMIP supercomputing center under the allocation 2017-[1133].

Author Contributions

J.S.G. performed all the DFT and TD-DFT calculations, participated in the interpretation of the results, and participated in the manuscript preparation. F.T. participated in the discussions and interpretation of the results. F.A., J.-L.H., and I.M.D. were all involved in the interpretation of the results and in the manuscript preparation. M.B.-P. wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Photochromism, Molecules and Systems; Dürr, H.; Bouas-Laurent, H. (Eds.) Elsevier: Amsterdam, The Netherlands, 1990. [Google Scholar]
  2. Bouas-Laurent, H.; Dürr, H. Organic photochromism. Pure Appl. Chem. 2001, 73, 639–665. [Google Scholar] [CrossRef]
  3. Crano, J.C.; Guglielmetti, R.J. Organic Photochromic and Thermochromic Compound; Plenum Press: New York, NY, USA; London, UK, 1998. [Google Scholar]
  4. Irie, M. Diarylethenes for memories and switches. Chem. Rev. 2000, 100, 1685–1716. [Google Scholar] [CrossRef] [PubMed]
  5. Berkovic, G.; Krongauz, V.; Weiss, V. Spiropyrans and spirooxazines for memories and switches. Chem. Rev. 2000, 100, 1741–1753. [Google Scholar] [CrossRef] [PubMed]
  6. Yokoyama, Y. Fulgides for memories and switches. Chem. Rev. 2000, 100, 1717–1739. [Google Scholar] [CrossRef] [PubMed]
  7. Görner, H.; Fischer, C.; Gierisch, S.; Daub, J. Dihydroazulene/vinylheptafulvene photochromism: Effects of substituents, solvent, and temperature in the photorearrangement of dihydroazulenes to vinylheptafulvenes. J. Phys. Chem. 1993, 97, 4110–4117. [Google Scholar] [CrossRef]
  8. Jockusch, S.; Turro, N.J.; Blackburn, F.R. Photochromism of 2H-naphto[1,2-b]pyrans: A spectroscopic investigation. J. Phys. Chem. A 2002, 106, 9236–9241. [Google Scholar] [CrossRef]
  9. Mitchell, R.H. The metacyclophanediene-dihydropyrene photochromic π switch. Eur. J. Org. Chem. 1999, 2695–2703. [Google Scholar] [CrossRef]
  10. Coppens, P.; Novozhilova, I.; Kovalevsky, A. Photoinduced linkage isomers of transition-metal nitrosyl compounds and related complexes. Chem. Rev. 2002, 102, 861–883. [Google Scholar] [CrossRef] [PubMed]
  11. Bitterwolf, T.E. Photochemical nitrosyl linkage isomerism/metastable states. Coord. Chem. Rev. 2006, 250, 1196–1207. [Google Scholar] [CrossRef]
  12. Rack, J.J. Electron transfer triggered sulfoxide isomerization in ruthenium and osmium complexes. Coord. Chem. Rev. 2009, 253, 78–85. [Google Scholar] [CrossRef]
  13. McClure, B.A.; Rack, J.J. Isomerization in photochromic ruthenium sulfoxide complexes. Eur. J. Inorg. Chem. 2010, 3895–3904. [Google Scholar] [CrossRef]
  14. Sylvester, S.O.; Cole, J.M.; Waddell, P.G. Photoconversion bonding mechanism in ruthenium sulfur dioxide linkage photoisomers revealed by in situ diffraction. J. Am. Chem. Soc. 2012, 134, 11860–11863. [Google Scholar] [CrossRef] [PubMed]
  15. Feringa, B.L. The art of building small: From molecular switches to molecular motors. J. Org. Chem. 2007, 72, 6635–6652. [Google Scholar] [CrossRef] [PubMed]
  16. Szaciłowski, K. Digital Information Processing in Molecular Systems. Chem. Rev. 2008, 108, 3481–3548. [Google Scholar] [CrossRef] [PubMed]
  17. Andréasson, J.; Pischel, U. Smart Molecules at Work-Mimicking Advanced Logic Operations. Chem. Soc. Rev. 2010, 39, 174–188. [Google Scholar] [CrossRef] [PubMed]
  18. Szymanski, W.; Beierle, J.M.; Kistemaker, H.A.V.; Velema, W.A.; Feringa, B.L. Reversible photocontrol of biological systems by the incorporation of molecular photoswitches. Chem. Rev. 2013, 113, 6114–6178. [Google Scholar] [CrossRef] [PubMed]
  19. Pischel, U.; Andréasson, J.; Gust, D.; Pais, V. Information processing with molecules—Quo vadis? ChemPhysChem 2013, 14, 28–46. [Google Scholar] [CrossRef] [PubMed]
  20. Kim, T.; Zhu, L.; Al-Kaysi, R.O.; Bardeen, C. Organic photomechanical materials. ChemPhysChem 2014, 15, 400–414. [Google Scholar] [CrossRef] [PubMed]
  21. Asadirad, A.M.; Boutault, S.; Erno, Z.; Branda, N.R. Controlling a polymer adhesive using light and a molecular switch. J. Am. Chem. Soc. 2014, 136, 3024–3027. [Google Scholar] [CrossRef] [PubMed]
  22. Boggio-Pasqua, M.; Bearpark, M.J.; Hunt, P.A.; Robb, M.A. Dihydroazulene/vinylheptafulvene photochromism: A model for one-way photochemistry via a conical intersection. J. Am. Chem. Soc. 2002, 124, 1456–1470. [Google Scholar] [CrossRef] [PubMed]
  23. Boggio-Pasqua, M.; Ravaglia, M.; Bearpark, M.J.; Garavelli, M.; Robb, M.A. Can diarylethene photochromism be explained by a reaction path alone? A CASSCF study with model MMVB dynamics. J. Phys. Chem. A 2003, 107, 11139–11152. [Google Scholar] [CrossRef]
  24. Migani, A.; Gentili, P.L.; Negri, F.; Olivucci, M.; Romani, A.; Favaro, G.; Becker, R.S. The ring-opening reaction of chromenes: A photochemical mode-dependent transformation. J. Phys. Chem. A 2005, 109, 8684–8692. [Google Scholar] [CrossRef] [PubMed]
  25. Maurel, F.; Aubard, J.; Millie, P.; Dognon, J.P.; Rajzmann, M.; Guglielmetti, R.; Samat, A. Quantum chemical study of the photocoloration reaction in the napthoxazine series. J. Phys. Chem. A 2006, 110, 4759–4771. [Google Scholar] [CrossRef] [PubMed]
  26. Boggio-Pasqua, M.; Bearpark, M.J.; Ogliaro, F.; Robb, M.A. Photochemical reactivity of 2-vinylbiphenyl and 2-vinyl-1,3-terphenyl: The balance between nonadiabatic and adiabatic photocyclization. J. Am. Chem. Soc. 2006, 128, 10533–10540. [Google Scholar] [CrossRef] [PubMed]
  27. Boggio-Pasqua, M.; Bearpark, M.J.; Robb, M.A. Towards a mechanistic understanding of the photochromism of dimethyldihydropyrenes. J. Org. Chem. 2007, 72, 4497–4503. [Google Scholar] [CrossRef] [PubMed]
  28. Tomasello, G.; Ogliaro, F.; Bearpark, M.J.; Robb, M.A.; Garavelli, M. Modeling the photophysics and photochromic potential of 1,2-dihydronaphthalene (DHN): A combined CASPT2//CASSCF-topological and MMVB-dynamical investigation. J. Phys. Chem. A 2008, 112, 10096–10107. [Google Scholar] [CrossRef] [PubMed]
  29. Tomasello, G.; Bearpark, M.J.; Robb, M.A.; Orlandi, G.; Garavelli, M. Significance of a zwitterionic state for fulgide photochromism: Implications for the design of mimics. Angew. Chem. Int. Ed. 2010, 49, 2913–2916. [Google Scholar] [CrossRef] [PubMed]
  30. Nenov, A.; Schreier, W.J.; Koller, F.O.; Braun, M.; de Vivie-Riedle, R.; Zinth, W.; Pugliesi, I. Molecular model of the ring-opening and ring-closure reaction of a fluorinated indolylfulgide. J. Phys. Chem. A 2012, 116, 10518–10528. [Google Scholar] [CrossRef] [PubMed]
  31. Raymo, F.M. Computational insights on the isomerization of photochromic oxazines. J. Phys. Chem. A 2012, 116, 11888–11895. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, F.; Morokuma, K. Multiple pathways for the primary step of the spiropyran photochromic reaction: A CASPT2//CASSCF study. J. Am. Chem. Soc. 2013, 135, 10693–10702. [Google Scholar] [CrossRef] [PubMed]
  33. Boggio-Pasqua, M.; Garavelli, M. Rationalization and design of enhanced photoinduced cycloreversion in photochromic dimethyldihydropyrenes by theoretical calculations. J. Phys. Chem. A 2015, 119, 6024–6032. [Google Scholar] [CrossRef] [PubMed]
  34. Daniel, C.; Gourlaouen, C. Chemical Bonding Alteration Upon Electronic Excitation in Transition Metal Complexes. Cood. Chem. Rev. 2017, 344, 131–149. [Google Scholar] [CrossRef]
  35. Ciofini, I.; Daul, C.A.; Adamo, C. Phototriggered Linkage Isomerization in Ruthenium-Dimethylsulfoxyde Complexes: Insights from Theory. J. Phys. Chem. A 2003, 107, 11182–11190. [Google Scholar] [CrossRef]
  36. Göttle, A.J.; Dixon, I.M.; Alary, F.; Heully, J.-L.; Boggio-Pasqua, M. Adiabatic versus nonadiabatic photoisomerization in photochromic ruthenium sulfoxide complexes: a mechanistic picture from density functional theory calculations. J. Am. Chem. Soc. 2011, 133, 9172–9174. [Google Scholar] [CrossRef] [PubMed]
  37. Vieuxmaire, O.P.J.; Piau, R.E.; Alary, F.; Heully, J.-L.; Sutra, P.; Igau, A.; Boggio-Pasqua, M. Theoretical investigation of phosphinidene oxide polypyridine ruthenium(II) complexes: Toward the design of a new class of photochromic compounds. J. Phys. Chem. A 2013, 117, 12821–12830. [Google Scholar] [CrossRef] [PubMed]
  38. Göttle, A.J.; Alary, F.; Dixon, I.M.; Heully, J.-L.; Boggio-Pasqua, M. Unravelling the S→O linkage photoisomerization mechanisms in cis- and trans-[Ru(bpy)2(DMSO)2]2+ using density functional theory. Inorg. Chem. 2014, 53, 6752–6760. [Google Scholar] [CrossRef] [PubMed]
  39. Li, H.; Zhang, L.; Wang, Y.; Fan, X. Theoretical Studies on the Photoisomerization Mechanism of Osmium(II) Sulfoxide Complexes. RSC Adv. 2015, 5, 58580–58586. [Google Scholar] [CrossRef]
  40. Li, H.; Zhang, L.; Zheng, I.; Li, X.; Fan, X.; Zhao, Y. Photoisomerization Mechanism of Ruthenium Sulfoxide Complexes: Role of the Metal-Centered Excited State in the Bond Rupture and Bond Construction Processes. Chem. Eur. J. 2016, 22, 14285–14292. [Google Scholar] [CrossRef] [PubMed]
  41. Sanz García, J.; Alary, F.; Boggio-Pasqua, M.; Dixon, I.M.; Malfant, I.; Heully, J.-L. Establishing the Two-Photon Linkage Isomerization Mechanism in the Nitrosyl Complex trans-[RuCl(NO)(py)4]2+ by DFT and TDDFT. Inorg. Chem. 2015, 54, 8310–8318. [Google Scholar] [CrossRef] [PubMed]
  42. Sanz García, J.; Alary, F.; Boggio-Pasqua, M.; Dixon, I.M.; Heully, J.-L. Is photoisomerization required for NO photorelease in ruthenium nitrosyl complexes? J. Mol. Model. 2016, 22, 284. [Google Scholar] [CrossRef] [PubMed]
  43. Woike, Th.; Haussühl, S. Infrared-Spectroscopic and Differential Scanning Calorimetric Studies of the Two Light-Induced Metastable States in K2[Ru(NO2)4(OH)(NO)]. Solid State Commun. 1993, 86, 333–337. [Google Scholar] [CrossRef]
  44. Fomitchev, D.V.; Coppens, P. X-ray Diffraction Analysis of Geometry Changes upon Excitation: The Ground-State and Metastable-State Structures of K2[Ru(NO2)4(OH)(NO)]. Inorg. Chem. 1996, 35, 7021–7026. [Google Scholar] [CrossRef] [PubMed]
  45. Fomitchev, D.V.; Coppens, P.; Li, T.; Bagley, K.A.; Chen, L.; Richter-Addo, G.B. Photo-induced metastable linkage isomers of ruthenium nitrosyl porphyrins. Chem. Commun. 1999, 2013–2014. [Google Scholar] [CrossRef]
  46. Da Silva, S.C.; Franco, D.W. Metastable Excited State and Electronic Structure of [Ru(NH3)5NO]3+ and [Ru(NH3)4(OH)NO]2+. Spectrochim. Acta A 1999, 55, 1515–1525. [Google Scholar] [CrossRef]
  47. Gorelsky, S.I.; Lever, A.B.P. Metastable States of Ruthenium (II) Nitrosyl Complexes and Comparison with [Fe(CN)5NO]2−. Int. J. Quantum Chem. 2000, 80, 636–645. [Google Scholar] [CrossRef]
  48. Ferlay, S.; Schmalle, H.W.; Francese, G.; Stoeckli-Evans, H.; Imlau, M.; Schaniel, D.; Woike, T. Light-Induced Metastable States in Oxalatenitrosylruthenium(II) and Terpyridinenitrosylruthenium(II) Complexes. Inorg. Chem. 2004, 43, 3500–3506. [Google Scholar] [CrossRef] [PubMed]
  49. Schaniel, D.; Woike, T.; Boskovic, C.; Güdel, H.-U. Evidence for two light-induced metastable states in Cl3[Ru(NH3)5NO]H2O. Chem. Phys. Lett. 2004, 390, 347–351. [Google Scholar] [CrossRef]
  50. Zangl, A.; Klüfers, P.; Schaniel, D.; Woike, T. Photoinduced Linkage Isomerism of {RuNO}6 Complexes with Bioligands and Related Chelators. Dalton Trans. 2009, 6, 1034–1045. [Google Scholar] [CrossRef] [PubMed]
  51. Schaniel, D.; Cormary, B.; Malfant, I.; Valade, L.; Woike, T.; Delley, B.; Krämer, K.W.; Güdel, H.-U. Photogeneration of Two Metastable NO Linkage Isomers with High Populations of up to 76% in trans-[RuCl(py)4(NO)][PF6]2·1/2H2O. Phys. Chem. Chem. Phys. 2007, 9, 3717–3724. [Google Scholar] [CrossRef] [PubMed]
  52. Cormary, B.; Malfant, I.; Buron-Le Cointe, M.; Toupet, L.; Delley, B.; Schaniel, D.; Mockus, N.; Woike, T.; Fejfarová, K.; Petříček, V.; Dušek, M. [Ru(py)4Cl(NO)](PF6)2·0.5H2O: A Model System for Structural Determination and Ab Initio Calculations of Photo-Induced Linkage NO Isomers. Acta Cryst. B 2009, 65, 612–623. [Google Scholar] [CrossRef] [PubMed]
  53. Cormary, B.; Ladeira, S.; Jacob, K.; Lacroix, P.G.; Woike, T.; Schaniel, D.; Malfant, I. Structural Influence on the Photochromic Response of a Series of Ruthenium Mononitrosyl Complexes. Inorg. Chem. 2012, 51, 7492–7501. [Google Scholar] [CrossRef] [PubMed]
  54. Khadeeva, L.; Kaszub, W.; Lorenc, M.; Malfant, I.; Buron-Le Cointe, M. Two-Step Photon Absorption Driving the Chemical Reaction in the Model Ruthenium Nitrosyl System [Ru(py)4Cl(NO)](PF6)2·1/2H2O. Inorg. Chem. 2016, 55, 4117–4123. [Google Scholar] [CrossRef] [PubMed]
  55. Tassé, M.; Mohammed, H.S.; Sabourdy, C.; Mallet-Ladeira, S.; Lacroix, P.G.; Malfant, I. Synthesis, Crystal Structure, Spectroscopic, and Photoreactive Properties of a Ruthenium(II)-Mononitrosyl Complex. Polyhedron 2016, 119, 350–358. [Google Scholar] [CrossRef]
  56. Ford, P.C.; Bourassa, J.; Miranda, K.; Lee, B.; Lorkovic, I.; Boggs, S.; Kudo, S.; Laverman, L. Photochemistry of Metal Nitrosyl Complexes. Delivery of Nitric Oxide to Biological Targets. Coord. Chem. Rev. 1998, 171, 185–202. [Google Scholar] [CrossRef]
  57. Tfouni, E.; Krieger, M.; McGarvey, B.R.; Franco, D.W. Structure, Chemical and Photochemical Reactivity and Biological Activity of Some Ruthenium Amine Nitrosyl Complexes. Coord. Chem. Rev. 2003, 236, 57–69. [Google Scholar] [CrossRef]
  58. Szundi, I.; Rose, M.J.; Sen, I.; Eroy-Reveles, A.A.; Mascharak, P.K.; Einarsdóttir, Ó. A New Approach for Studying Fast Biological Reactions Involving Nitric Oxide: Generation of NO Using Photolabile Ruthenium and Manganese NO Donors. Photochem. Photobiol. 2006, 82, 1377–1384. [Google Scholar] [CrossRef] [PubMed]
  59. Bitterwolf, T.E. Photolysis of [Ru(bipy)2(NO)Cl](PF6)2 in Frozen Ionic Glass Matrices. Evidence for Nitrosyl Linkage Isomerism and NO-Loss in a Physiologically Relevant Nitric Oxide Source. Inorg. Chem. Commun. 2008, 11, 772–773. [Google Scholar] [CrossRef]
  60. Rose, M.J.; Mascharak, P.K. Photoactive Ruthenium Nitrosyls: Effects of Light and Potential Application as NO Donors. Coord. Chem. Rev. 2008, 252, 2093–2114. [Google Scholar] [CrossRef] [PubMed]
  61. Rose, M.J.; Fry, N.L.; Marlow, R.; Hink, L.; Mascharak, P.K. Sensitization of Ruthenium Nitrosyls to Visible Light via Direct Coordination of the Dye Resorufin: Trackable NO Donors for Light-Triggered NO Delivery to Cellular Targets. J. Am. Chem. Soc. 2008, 130, 8834–8846. [Google Scholar] [CrossRef] [PubMed]
  62. Giglmeier, H.; Kerscher, T.; Klüfers, P.; Schaniel, D.; Woike, T. Nitric-Oxide Photorelease and Photoinduced Linkage Isomerism on Solid [Ru(NO)(terpy)(L)]BPh4 (L = glycolate dianion). Dalton Trans. 2009, 9113–9116. [Google Scholar] [CrossRef] [PubMed]
  63. Ostrowski, A.D.; Ford, P.C. Metal Complexes as Photochemical Nitric Oxide Precursors: Potential Applications in the Treatment of Tumors. Dalton Trans. 2009, 10660–10669. [Google Scholar] [CrossRef] [PubMed]
  64. Fry, N.L.; Mascharak, P.K. Photoactive Ruthenium Nitrosyls as NO Donors: How to Sensitize Them toward Visible Light. Acc. Chem. Res. 2011, 44, 289–298. [Google Scholar] [CrossRef] [PubMed]
  65. Akl, J.; Sasaki, I.; Lacroix, P.G.; Malfant, I.; Mallet-Ladeira, S.; Vicendo, P.; Farfán, N.; Santillan, R. Comparative Photo-Release of Nitric Oxide from Isomers of Substituted Terpyridinenitrosyl-Ruthenium(II) Complexes: Experimental and Computational Investigations. Dalton Trans. 2014, 43, 12721–12733. [Google Scholar] [CrossRef] [PubMed]
  66. deBoer, T.R.; Mascharak, P.K. Recent Progress in Photoinduced NO Delivery With Designed Ruthenium Nitrosyl Complexes. Adv. Inorg. Chem. 2015, 67, 145–170. [Google Scholar]
  67. Talotta, F.; Heully, J.-L.; Alary, F.; Dixon, I.M.; González, L.; Boggio-Pasqua, M. Linkage Photoisomerization Mechanism in a Photochromic Ruthenium Nitrosyl Complex: New Insights from a MS-CASPT2 Study. J. Chem. Theory Comput. in press.
  68. Freitag, L.; González, L. Theoretical Spectroscopy and Photodynamics of a Ruthenium Nitrosyl Complex. Inorg. Chem. 2014, 53, 6415–6426. [Google Scholar] [CrossRef] [PubMed]
  69. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  70. Weinan, E.; Ren, W.; Vanden-Eijnden, E. Simplified and improved string method for computing the minimum energy paths in barrier-crossing events. J. Chem. Phys. 2007, 126, 164103. [Google Scholar] [CrossRef] [PubMed]
  71. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  72. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  73. Schäfer, A.; Horn, H.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets for atoms Li to Kr. J. Chem. Phys. 1992, 97, 2571–2577. [Google Scholar] [CrossRef]
  74. Andrae, D.; Häußermann, U.; Dolg, M.; Stoll, H.; Preuß, H. Energy-adjusted ab initio pseudopotentials for the second and third row transition elements. Theor. Chim. Acta 1990, 77, 123–141. [Google Scholar] [CrossRef]
  75. Martin, J.M.L.; Sundermann, A. Correlation consistent valence basis sets for use with the Stuttgart–Dresden–Bonn relativistic effective core potentials: The atoms Ga–Kr and In–Xe. J. Chem. Phys. 2001, 114, 3408–3420. [Google Scholar] [CrossRef]
  76. Tao, J.; Perdew, J.P.; Staroverov, V.N.; Scuseria, G.E. Climbing the Density Functional Ladder: Nonempirical Meta–Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401. [Google Scholar] [CrossRef] [PubMed]
  77. Becke, A.D. A new mixing of Hartree-Fock and local density-functional theories. J. Chem. Phys. 1993, 98, 1372–1377. [Google Scholar] [CrossRef]
  78. Hirata, S.; Head-Gordon, M. Time-Dependent Density Functional Theory Within the Tamm−Dancoff Approximation. Chem. Phys. Lett. 1999, 314, 291–299. [Google Scholar] [CrossRef]
  79. Neese, F. An improvement of the resolution of the identity approximation for the formation of the Coulomb matrix. J. Comput. Chem. 2003, 24, 1740–1747. [Google Scholar] [CrossRef] [PubMed]
  80. Neese, F.; Wennmohs, F.; Hansen, A.; Becker, U. Efficient, approximate and parallel Hartree-Fock and hybrid DFT calculations. A 'chain-of-spheres' algorithm for the Hartree-Fock exchange. Chem. Phys. 2009, 356, 98–109. [Google Scholar] [CrossRef]
  81. Neese, F. The ORCA program system. WIREs Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are not available from the authors.
Figure 1. The series of ruthenium nitrosyl complexes studied in this work. From left to right: trans-[RuCl(NO)(py)4]2+ (1); trans-[RuBr(NO)(py)4]2+ (2); and trans-(Cl,Cl)[RuCl2(NO)(tpy)]+ (3) (py: pyridine, tpy: terpyridine).
Figure 1. The series of ruthenium nitrosyl complexes studied in this work. From left to right: trans-[RuCl(NO)(py)4]2+ (1); trans-[RuBr(NO)(py)4]2+ (2); and trans-(Cl,Cl)[RuCl2(NO)(tpy)]+ (3) (py: pyridine, tpy: terpyridine).
Molecules 22 01667 g001
Figure 2. Potential energy profiles for the singlet ground state (blue curve) and the lowest triplet state (red curve) along the N→O linkage photoisomerization pathway derived from density functional theory (DFT) calculations of complex (1) [41]. The reaction coordinate is represented as the Ru-N-O angle. GS, MS2, and MS1 correspond to the N-bonded, sideways NO-bonded, and O-bonded isomers in the singlet state, respectively. 3GS, 3MS2, and 3MS1 correspond to the N-bonded, sideways NO-bonded, and O-bonded isomers in the triplet state, respectively. TS and 3TS denote transition states lying on the singlet and triplet potential energy surfaces (PES), respectively. ISC and IC stand for intersystem crossing and internal conversion. Double-cone pictograms represent the minimum energy crossing points (MECP) between the singlet and triplet states (in black are the two ones relevant to N→O linkage photoisomerization). The black arrows indicate the sequential steps of the N→O linkage photoisomerization mechanism. The orange dashed arrows and orange double cones represent photostabilizing non-radiative decay paths to GS and MS2. Pyridine ligands have been omitted in the structures for clarity.
Figure 2. Potential energy profiles for the singlet ground state (blue curve) and the lowest triplet state (red curve) along the N→O linkage photoisomerization pathway derived from density functional theory (DFT) calculations of complex (1) [41]. The reaction coordinate is represented as the Ru-N-O angle. GS, MS2, and MS1 correspond to the N-bonded, sideways NO-bonded, and O-bonded isomers in the singlet state, respectively. 3GS, 3MS2, and 3MS1 correspond to the N-bonded, sideways NO-bonded, and O-bonded isomers in the triplet state, respectively. TS and 3TS denote transition states lying on the singlet and triplet potential energy surfaces (PES), respectively. ISC and IC stand for intersystem crossing and internal conversion. Double-cone pictograms represent the minimum energy crossing points (MECP) between the singlet and triplet states (in black are the two ones relevant to N→O linkage photoisomerization). The black arrows indicate the sequential steps of the N→O linkage photoisomerization mechanism. The orange dashed arrows and orange double cones represent photostabilizing non-radiative decay paths to GS and MS2. Pyridine ligands have been omitted in the structures for clarity.
Molecules 22 01667 g002
Figure 3. Time-dependent DFT (TD-DFT) absorption spectra of the three linkage isomers (GS in orange, MS2 in black, and MS1 in green) of trans-[RuCl(NO)(py)4]2+. Broadening model used: Gaussian function with a half-width at half-height of 0.2 eV.
Figure 3. Time-dependent DFT (TD-DFT) absorption spectra of the three linkage isomers (GS in orange, MS2 in black, and MS1 in green) of trans-[RuCl(NO)(py)4]2+. Broadening model used: Gaussian function with a half-width at half-height of 0.2 eV.
Molecules 22 01667 g003
Figure 4. TD-DFT absorption spectra of the three linkage isomers (GS in orange, MS2 in black, and MS1 in green) of trans-[RuBr(NO)(py)4]2+. Broadening model used: Gaussian function with a half-width at half-height of 0.2 eV.
Figure 4. TD-DFT absorption spectra of the three linkage isomers (GS in orange, MS2 in black, and MS1 in green) of trans-[RuBr(NO)(py)4]2+. Broadening model used: Gaussian function with a half-width at half-height of 0.2 eV.
Molecules 22 01667 g004
Figure 5. TD-DFT absorption spectra of the three linkage isomers (GS in orange, MS2 in black, and MS1 in green) of trans-(Cl,Cl)[RuCl2(NO)(tpy)]+. Broadening model used: Gaussian function with a half-width at half-height of 0.2 eV.
Figure 5. TD-DFT absorption spectra of the three linkage isomers (GS in orange, MS2 in black, and MS1 in green) of trans-(Cl,Cl)[RuCl2(NO)(tpy)]+. Broadening model used: Gaussian function with a half-width at half-height of 0.2 eV.
Molecules 22 01667 g005
Table 1. Relative energies (eV) for all stationary points on the singlet potential energy surface for each complex.
Table 1. Relative energies (eV) for all stationary points on the singlet potential energy surface for each complex.
Relative Energytrans-[RuCl(NO)(py)4]2+
(1)
trans-[RuBr(NO)(py)4]2+
(2)
trans-(Cl,Cl)[RuCl2(NO)(tpy)]+
(3)
ΔE(MS2GS)1.561.511.48
ΔE(MS1GS)1.861.821.91
ΔE(TS1GS)2.442.302.30 1
ΔE(TS2MS2)1.571.491.44 1
1 Values obtained with Zero Temperature String (ZTS) calculations [42].
Table 2. Relative energies (eV) for all stationary points on the triplet potential energy surface for each complex.
Table 2. Relative energies (eV) for all stationary points on the triplet potential energy surface for each complex.
Relative Energytrans-[RuCl(NO)(py)4]2+
(1)
trans-[RuBr(NO)(py)4]2+
(2)
trans-(Cl,Cl)[RuCl2(NO)(tpy)]+
(3)
ΔE(3GSGS)1.411.320.98
ΔE(3MS2GS)2.252.191.85
ΔE(3MS1GS)2.081.981.70
ΔE(3TS13GS)1.061.050.98 2
ΔE(3TS23MS2)0.250.20 10.16 2
1 Without explicit dispersion. 2 Values obtained with Zero Temperature String (ZTS) calculations [42].
Table 3. Relative energies (eV) for selected critical points in the studied complexes.
Table 3. Relative energies (eV) for selected critical points in the studied complexes.
Relative Energytrans-[RuCl(NO)(py)4]2+
(1)
trans-[RuBr(NO)(py)4]2+
(2)
trans-(Cl,Cl)[RuCl2(NO)(tpy)]+
(3)
ΔE(MECP13GS)0.000.000.01
ΔE(MECP23GS)0.670.650.90
ΔE(MECP33MS2)0.030.030.02
ΔE(MECP43MS1)0.170.180.34

Share and Cite

MDPI and ACS Style

Sanz García, J.; Talotta, F.; Alary, F.; Dixon, I.M.; Heully, J.-L.; Boggio-Pasqua, M. A Theoretical Study of the N to O Linkage Photoisomerization Efficiency in a Series of Ruthenium Mononitrosyl Complexes. Molecules 2017, 22, 1667. https://doi.org/10.3390/molecules22101667

AMA Style

Sanz García J, Talotta F, Alary F, Dixon IM, Heully J-L, Boggio-Pasqua M. A Theoretical Study of the N to O Linkage Photoisomerization Efficiency in a Series of Ruthenium Mononitrosyl Complexes. Molecules. 2017; 22(10):1667. https://doi.org/10.3390/molecules22101667

Chicago/Turabian Style

Sanz García, Juan, Francesco Talotta, Fabienne Alary, Isabelle M. Dixon, Jean-Louis Heully, and Martial Boggio-Pasqua. 2017. "A Theoretical Study of the N to O Linkage Photoisomerization Efficiency in a Series of Ruthenium Mononitrosyl Complexes" Molecules 22, no. 10: 1667. https://doi.org/10.3390/molecules22101667

Article Metrics

Back to TopTop