+ All Categories
Home > Documents > Comment on “Exothermic Rate Restrictions in Long-Range ... · the Photoinduced Electron Transfer...

Comment on “Exothermic Rate Restrictions in Long-Range ... · the Photoinduced Electron Transfer...

Date post: 26-Jul-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
3
Published: June 02, 2011 r2011 American Chemical Society 7858 dx.doi.org/10.1021/jp111064q | J. Phys. Chem. A 2011, 115, 78587860 COMMENT pubs.acs.org/JPCA Comment on “Exothermic Rate Restrictions in Long-Range Photoinduced Charge Separations” Gonzalo Angulo, Arnulf Rosspeintner, and Eric Vauthey* ,Institute of Physical Chemistry, Polish Academy of Sciences, 01-224 Warsaw, Poland, and Physical Chemistry Department, Sciences II University of Geneva, Geneva 1211, Switzerland G omes et al. have recently reported in this journal on their experimental observation of exothermic rate restrictions (i.e., Marcus inverted region behavior) in long-range photo- induced charge separation in rigid matrices. 1,2 Although the Marcus inverted region has been well established in charge recombination reactions, 3,4 its observation for photoinduced bimolecular charge separation in solution (rigid or not) has so far not been reported unambiguously. 515 The authors performed uorescence quenching experiments between immobile reactants in glycerol:methanol (9:1) mixtures at 255 K. This medium is supposed to be exible enough to allow for charge separation and accommodate the reaction products thereof, whereas it eciently hinders the mutual translational diusion of the reactants. By doing so, the problem of diusion limited reactions is notably simplied, because only static quench- ingwill be observed. The experimental methods used in these studies were steady-state and nanosecond time-resolved (by means of time correlated single photon counting) uorescence spectroscopy. The experimental results comprise the change of uorescence quantum yields and uorescence time traces at dierent quencher concentrations spanning a reasonable range of free energy for electron transfer with 8 uorophore/quencher pairs. These data are analyzed using a model for remote electron transfer with an exponential distance dependence (eq 3 in ref 1), which eventually leads to a uorescence intensity decay (eq 4 in ref 1) depending on only three parameters: r, the contact distance of the reactants, which the authors estimate using Connolly surfaces; k ET 0 , the rate constant of electron transfer; and β, its distance decay factor (k ET 0 and β being obtained from the ts to the experiments). The data analysis consists basically of rescaling the experimental time traces with the uorescence quantum yields (obtained from the steady-state experiment) and tting the model to them. Eventually one identical β is obtained for all systems studied and thus the free energy dependence of electron transfer is completely contained in the k ET 0 values. The so observed existence of exothermic rate restrictions in photoinduced charge separations in rigid mediais then extensively discussed by the authors. Unfortunately, however, the authors omitted a self-consis- tency test on their results. If the time traces, I(t), and thus the kinetics, are well accounted for by the applied model and the obtained parameters, the steady-state quenching experiments (i.e., the change of the uorescence quantum yield, φ, with quencher concentration, c) also ought to be well reproduced by the time integrals of the model. We may briey explain this obvious and necessary self-consistency. φ ¼ Z ¥ 0 Z ¥ 0 I ðλ;t Þ dλ dt ¼ C Z ¥ 0 I ðλ ¼ λ em ;t Þ dt ð1Þ I ðt ¼ 0Þ¼ 1 "c ð2Þ φðc ¼ 0Þ φðcÞ ¼ Z ¥ 0 I ðc ¼ 0;t Þ dt Z ¥ 0 I ðc;t Þ dt ¼ τ Z ¥ 0 I ðc;t Þ dt ð3Þ The rst equation merely states that the full kinetics is contained in the steady-state spectrum and that the quantum yield is, except for some scaling factor, C, when only one specic emission wavelength, λ em , is observed instead of the entire emission spectrum, given by the time integral of the uorescence decay (as long as the eect of the dynamic Stokes shift is negligible). The second equation states that the initial population just after excitation is independent of the quencher concentration. This is true if there is no ground-state complex formation or changes in the properties of the medium that could aect the radiative properties of the uorophore upon addition of the quencher. Finally, the third equation states that the ratio of steady-state uorescence intensities at increasing quencher concentration, c, is equal to the ratio of the time integrals of the uorescence decays at the same concentrations. It can be easily seen that the last equation is simply the outcome of the consequent application of the former two equations. We tested the self-consistence of the used model and of the obtained parameters by simulating the steady-state results, given in the Supporting Information of ref 2, with the electron transfer parameters given in Table 1 of ref 2. To this end, the same reactivity model (eq 4 from ref 1) was applied to evaluate I(c,t) and eventually integrated numerically to give φ(c). The results of this attempt are shown in Figure 1. It can clearly be seen that there is a signicant discrepancy between the experimental and simulated data. Irrespective of the inherent reasons for this discrepancy (inappropriateness of the scaling procedure, wrong model, wrong parameters), we are convinced that the appropri- ateness of the model and the adjoint parameter set ought to be tested on, and should equally well describe, both data sets. 1619 As a consequence, the obtained parameters in general and the electron transfer rate constant, k ET 0 , in particular cannot be correct. We thus conclude that the inverted region for this kind of reactions still remains unobserved. Additionally, we make some further comments: It should be noted that the Perrin equation is not correctly written in ref 1 and if used as such will lead to erroneous Received: November 19, 2010 Revised: April 5, 2011
Transcript
Page 1: Comment on “Exothermic Rate Restrictions in Long-Range ... · the Photoinduced Electron Transfer Reactions of Ruthenium(II) Poly-pyridine Complexes with Phenolate Ions. J. Phys.

Published: June 02, 2011

r 2011 American Chemical Society 7858 dx.doi.org/10.1021/jp111064q | J. Phys. Chem. A 2011, 115, 7858–7860

COMMENT

pubs.acs.org/JPCA

Comment on “Exothermic Rate Restrictions in Long-RangePhotoinduced Charge Separations”Gonzalo Angulo,† Arnulf Rosspeintner,‡ and Eric Vauthey*,‡

†Institute of Physical Chemistry, Polish Academy of Sciences, 01-224 Warsaw, Poland, and ‡Physical Chemistry Department,Sciences II University of Geneva, Geneva 1211, Switzerland

Gomes et al. have recently reported in this journal on theirexperimental observation of exothermic rate restrictions

(i.e., Marcus inverted region behavior) in long-range photo-induced charge separation in rigid matrices.1,2 Although theMarcus inverted region has been well established in chargerecombination reactions,3,4 its observation for photoinducedbimolecular charge separation in solution (rigid or not) hasso far not been reported unambiguously.5�15

The authors performed fluorescence quenching experimentsbetween immobile reactants in glycerol:methanol (9:1) mixturesat 255 K. This medium is supposed to be flexible enough to allowfor charge separation and accommodate the reaction productsthereof, whereas it efficiently hinders the mutual translationaldiffusion of the reactants. By doing so, the problem of diffusionlimited reactions is notably simplified, because only “static quench-ing” will be observed. The experimental methods used in thesestudies were steady-state and nanosecond time-resolved (bymeans of time correlated single photon counting) fluorescencespectroscopy. The experimental results comprise the change offluorescence quantum yields and fluorescence time traces atdifferent quencher concentrations spanning a reasonable rangeof free energy for electron transfer with 8 fluorophore/quencherpairs. These data are analyzed using a model for remote electrontransfer with an exponential distance dependence (eq 3 in ref 1),which eventually leads to a fluorescence intensity decay (eq 4 in ref 1)depending on only three parameters: r, the contact distance of thereactants, which the authors estimate usingConnolly surfaces; kET

0 ,the rate constant of electron transfer; and β, its distance decayfactor (kET

0 and β being obtained from the fits to the experiments).The data analysis consists basically of rescaling the experimentaltime traces with the fluorescence quantum yields (obtained fromthe steady-state experiment) and fitting the model to them.Eventually one identical β is obtained for all systems studied andthus the free energy dependence of electron transfer is completelycontained in the kET

0 values. The so observed “existence ofexothermic rate restrictions in photoinduced charge separationsin rigid media” is then extensively discussed by the authors.

Unfortunately, however, the authors omitted a self-consis-tency test on their results. If the time traces, I(t), and thus thekinetics, are well accounted for by the applied model and theobtained parameters, the steady-state quenching experiments(i.e., the change of the fluorescence quantum yield, φ, withquencher concentration, c) also ought to be well reproduced bythe time integrals of the model. We may briefly explain thisobvious and necessary self-consistency.

φ ¼Z ¥

0

Z ¥

0Iðλ;tÞ dλ dt ¼ C

Z ¥

0Iðλ¼λem;tÞ dt ð1Þ

Iðt¼0Þ ¼ 1 "c ð2Þ

φðc¼0ÞφðcÞ ¼

Z ¥

0Iðc¼0;tÞ dt

Z ¥

0Iðc;tÞ dt

¼ τZ ¥

0Iðc;tÞ dt

ð3Þ

The first equation merely states that the full kinetics is containedin the steady-state spectrum and that the quantum yield is, exceptfor some scaling factor, C, when only one specific emissionwavelength, λem, is observed instead of the entire emissionspectrum, given by the time integral of the fluorescence decay(as long as the effect of the dynamic Stokes shift is negligible).The second equation states that the initial population just afterexcitation is independent of the quencher concentration. This istrue if there is no ground-state complex formation or changes inthe properties of the medium that could affect the radiativeproperties of the fluorophore upon addition of the quencher.Finally, the third equation states that the ratio of steady-statefluorescence intensities at increasing quencher concentration, c,is equal to the ratio of the time integrals of the fluorescencedecays at the same concentrations. It can be easily seen that thelast equation is simply the outcome of the consequent applicationof the former two equations.

We tested the self-consistence of the used model and of theobtained parameters by simulating the steady-state results, givenin the Supporting Information of ref 2, with the electron transferparameters given in Table 1 of ref 2. To this end, the samereactivity model (eq 4 from ref 1) was applied to evaluate I(c,t)and eventually integrated numerically to give φ(c). The results ofthis attempt are shown in Figure 1. It can clearly be seen thatthere is a significant discrepancy between the experimental andsimulated data. Irrespective of the inherent reasons for thisdiscrepancy (inappropriateness of the scaling procedure, wrongmodel, wrong parameters), we are convinced that the appropri-ateness of the model and the adjoint parameter set ought to betested on, and should equally well describe, both data sets.16�19

As a consequence, the obtained parameters in general and theelectron transfer rate constant, kET

0 , in particular cannot becorrect. We thus conclude that the inverted region for this kindof reactions still remains unobserved.

Additionally, we make some further comments:• It should be noted that the Perrin equation is not correctlywritten in ref 1 and if used as such will lead to erroneous

Received: November 19, 2010Revised: April 5, 2011

Page 2: Comment on “Exothermic Rate Restrictions in Long-Range ... · the Photoinduced Electron Transfer Reactions of Ruthenium(II) Poly-pyridine Complexes with Phenolate Ions. J. Phys.

7859 dx.doi.org/10.1021/jp111064q |J. Phys. Chem. A 2011, 115, 7858–7860

The Journal of Physical Chemistry A COMMENT

results. The proper equation is given as follows:20

φðc¼0ÞφðcÞ ¼ exp

43πRc

3c

� �ð4Þ

where Rc is the critical radius (in Å) and c is the quencherconcentration (in Å�3).

• The critical radii, which, incidentally, are different in the twomanuscripts do not reproduce the experimental Perrin plots.In addition, the way they have been extracted is unclear,especially considering that they do not reproduce thePerrin plots.

• In the Correction not only the time-resolved data but also allsteady-state data are not the same as the original ones. Thisfact was neither explained nor pointed out in the Correction.

’AUTHOR INFORMATION

Corresponding Author*E-mail: [email protected]. Phone: þ41 22 379 65 37. Fax:þ41 22 379 65 18.

’REFERENCES

(1) Gomes, P. J. S.; Serpa, C.; Nunes, R. M. D.; Arnaut, L. G.;Formosinho, S. J. Exothermic Rate Restrictions in Long-Range Photo-induced Charge Separations in Rigid Media. J. Phys. Chem. A 2010, 114,2778–2787.

(2) Gomes, P. J. S.; Serpa, C.; Nunes, R. M. D.; Arnaut, L. G.;Formosinho, S. J. Exothermic Rate Restrictions in Long-Range Photo-induced Charge Separations. J. Phys. Chem. A 2010, 114, 10759–10760(Addition/Correction).

Figure 1. Comparison between experimental (gray line and circles) and simulated (black lines) Perrin plots. The experimental data are calculated on thebasis of the linear fits to the Perrin plots shown in the Supporting Information of ref 2 while the simulated data have been calculated using eq 4 and itsnumeric integral with the parameters given in ref 2.

Page 3: Comment on “Exothermic Rate Restrictions in Long-Range ... · the Photoinduced Electron Transfer Reactions of Ruthenium(II) Poly-pyridine Complexes with Phenolate Ions. J. Phys.

7860 dx.doi.org/10.1021/jp111064q |J. Phys. Chem. A 2011, 115, 7858–7860

The Journal of Physical Chemistry A COMMENT

(3) Suppan, P. The Marcus inverted region. Top. Curr. Chem. 1992,163, 95–130.(4) Mataga, N.; Asahi, T.; Kanda, Y.; Okada, T.; Kakitani, T. The

bell-shaped energy gap dependence of the charge recombination reac-tion of geminate radical ion pairs produced by fluorescence quenchingreaction in acetonitrile solution. Chem. Phys. 1988, 127, 249–261.(5) Cho, K. C.; Che, C. M.; Ng, K. M.; Choy, C. L. Electron transfer

between azurin and metalloporphyrins. J. Phys. Chem. 1987, 91,3690–3693.(6) Chen, J. M.; Ho, T. I.; Mou, C. Y. Experimental investigation of

excited-state electron-transfer reaction: effects of free energy and solventon rates. J. Phys. Chem. 1990, 94, 2889–2896.(7) McCleskey, T. M.; Winkler, J. R.; Gray, H. B. Driving-force

effects on the rates of bimolecular electron-transfer reactions. J. Am.Chem. Soc. 1992, 114, 6935–6937.(8) Turr�o, C.; Zaleski, J. M.; Karabatsos, Y. M.; Nocera, D. G.

Bimolecular Electron Transfer in the Marcus Inverted Region. J. Am.Chem. Soc. 1996, 118, 6060–6067.(9) Sinha, S.; De, R.; Ganguly, T. Electron Transfer Reactions in the

Excited Singlet States of Dimethyl Substituted Phenol 2-NitrofluoreneSystems: Evidence for the Marcus Inverted Region and ConcurrentOccurrence of Energy Transfer Processes. J. Phys. Chem. A 1997, 101,2852–2858.(10) Thanasekaran, P.; Rajendran, T.; Rajagopal, S.; Srinivasan, C.;

Ramaraj, R.; Ramamurthy, P.; Venkatachalapathy, B. Marcus InvertedRegion in the Photoinduced Electron Transfer Reactions of Ruthenium-(II) Polypyridine Complexes with Phenolate Ions. J. Phys. Chem. A1997, 101, 8195–8199.(11) Li, C.; Hoffman,M. Z. Comment onMarcus Inverted Region in

the Photoinduced Electron Transfer Reactions of Ruthenium(II) Poly-pyridine Complexes with Phenolate Ions. J. Phys. Chem. A 1998, 102,6052–6053.(12) Chakraborty, A.; Chakrabarty, D.; Hazra, P.; Seth, D.; Sarkar,

N. Photoinduced intermolecular electron transfer between Coumarindyes and electron donating solvents in cetyltrimethylammonium bro-mide (CTAB) micelles: evidence for Marcus inverted region. Chem.Phys. Lett. 2003, 382, 508–517.(13) Fukuzumi, S.; Nishimine, M.; Ohkubo, K.; Tkachenko, N. V.;

Lemmetyinen, H. Driving Force Dependence of Photoinduced ElectronTransfer Dynamics of Intercalated Molecules in DNA. J. Phys. Chem. B2003, 107, 12511–12518.(14) Kumbhakar, M.; Nath, S.; Pal, H.; Sapre, A. V.; Mukherjee, T.

Photoinduced intermolecular electron transfer from aromatic amines tocoumarin dyes in sodium dodecyl sulphate micellar solutions. J. Chem.Phys. 2003, 119, 388–399.(15) Kumbhakar, M.; Nath, S.; Mukherjee, T.; Pal, H. Intermole-

cular electron transfer between coumarin dyes and aromatic amines inTriton-X-100 micellar solutions: Evidence for Marcus inverted region.J. Chem. Phys. 2004, 120, 2824–2834.(16) Swallen, S. F.; Weidemaier, K.; Tavernier, H. L.; Fayer, M. D.

Experimental and Theoretical Analysis of Photoinduced ElectronTransfer: Including the Role of Liquid Structure. J. Phys. Chem. 1996,100, 8106–8117.(17) Rosspeintner, A.; Kattnig, D. R.; Angulo, G.; Landgraf, S.;

Grampp, G. The Rehm-Weller Experiment in View of Distant ElectronTransfer. Chem.—Eur. J. 2008, 14, 6213–6221.(18) Rosspeintner, A.; Kattnig, D. R.; Angulo, G.; Landgraf, S.;

Grampp, G.; Cuetos, A. On the Coherent Description of Diffusion-Influenced Fluorescence Quenching Experiments. Chem.—Eur. J. 2007,13, 6474–6483.(19) Angulo, G.; Kattnig, D. R.; Rosspeintner, A.; Grampp, G.;

Vauthey, E. On the Coherent Description of Diffusion-InfluencedFluorescence Quenching Experiments II: Early Events. Chem.—Eur. J.2010, 16, 2291–2299.(20) Valeur, B. Molecular Fluorescence: Principles and Applications;

Wiley-VCH: New York, 2001.


Recommended