Elucidation of the Adsorption Distribution of Cationic Porphyrin on the

Feb 14, 2018 - Petty, Ganguly, Rankine, Baucum, Gillan, Eddy, Léon, and Müller. 2018 122 (8), pp 4670–4680. Abstract: Molecular silver clusters ar...
0 downloads 0 Views 1MB Size
Subscriber access provided by READING UNIV

Article

The Elucidation of the Adsorption Distribution of Cationic Porphyrin on the Inorganic Surface by Energy Transfer as a Molecular Ruler Ayumi Nakayama, Junya Mizuno, Yuta Ohtani, Tetsuya Shimada, and Shinsuke Takagi J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b12104 • Publication Date (Web): 14 Feb 2018 Downloaded from http://pubs.acs.org on February 18, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The Elucidation of the Adsorption Distribution of Cationic Porphyrin on the Inorganic Surface by Energy Transfer as a Molecular Ruler Ayumi Nakayama, Junya Mizuno, Yuta Ohtani, Tetsuya Shimada and Shinsuke Takagi* Department of Applied Chemistry, Graduate Course of Urban Environmental Sciences, Tokyo Metropolitan University, Minami!ohsawa 1!1, Hachioji, Tokyo 192!0397, Japan.

ABSTRACT. The adsorption distribution of porphyrin dye on the inorganic surface was examined by the use of Förster resonance energy transfer (FRET).

Because the efficiency of

FRET depends on the distance between energy donor and acceptor, FRET could be used as a spectroscopic molecular ruler to estimate the adsorption distribution of dyes on the inorganic surface.

Saponite and cationic porphyrin were used as an anionic inorganic nanosheet and an

adsorbing dye (an energy acceptor) on the clay surface.

The edge of saponite was modified by

pyrene that is an energy donor, through silane coupling.

From the estimation of FRET

efficiency, the adsorption distribution of cationic porphyrin is turned out to be biased to the center of the nanosheet surface.

The factor to determine such adsorption distribution of cationic

porphyrin was discussed based on the electrostatic interactions between anionic nanosheet and cationic porphyrin.

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 25

Introduction The understanding and control of the assembly structure of molecules on the solid surface is a quite important subject in the field of surface, supramolecular and materials chemistry.

In

general, the assembly structure on the surface is difficult to control and tends to be irregular. Because of unfavorable interactions between transition dipoles in the assembly, the photochemical functionality of molecule would be altered and be often lost. For example, in the case of dyes on the surface, the excited lifetime becomes shorter and its photo-activity is drastically decreased in many cases. In the research field of two-dimensional materials such as inorganic nanosheets, the assembly structure control on their surface have been developed especially in recent years.1-12 For example, assembly structures with high-density without an aggregation were reported for layered silicate-organic dye systems.

This phenomenon was observed when the distance

between cationic sites in the guest molecule matches well to that of anionic sites on the clay surface.

Thus, this phenomenon is called “size-matching effect”. 13, 14

By the use of this

technique, the efficient photochemical reaction systems such as an artificial light harvesting systems have been reported. 15-18 of dyes have been reported.

Recently, unique effects of nanosheet on emission property By the complex formation with clay, the strong emission

enhancement behaviors of dyes were observed.19, induced emission”.

21-23

20

This effect is called “Surface-fixation

Despite those unique phenomena of nanosheet-dyes complexes,

researches on the adsorption distribution of dyes on the clay surface are few, because of its difficulty in observation.

Some researchers discussed the adsorption distribution of dyes by

ACS Paragon Plus Environment

2

Page 3 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

observing photochemical processes such as an electron transfer between two types of molecule. 24-26

A. J. Bard et al. reported that two types of dyes tend to segregate each other. 26 However, in

those systems, two types of molecule interact each other and interfere the original single-component adsorption structure, thus, it is difficult to know the intrinsic adsorption assembly structure so far. In this study, by the use of system where the energy donor is fixed at the edge of the clay, the estimation of dye (energy acceptor) adsorption assembly structure on the surface was examined.

The energy transfer reaction was used as molecular ruler in this system.

Because

energy donor and acceptor molecules are spatially well separated, the intrinsic assembly structure of adsorbing molecules can be discussed in the present system. Saponite clay was used as a typical inorganic nanosheet.

The particle shape is

approximately quite thin circular cylinder where the thickness and diameter are 0.96 nm and around 50 nm.1

Because there are silanol groups at the edge of the clay sheet, the selective

modification by the energy donor molecule is possible through the silane coupling reaction. 27-29 In the present research, the edge of the clay sheet is modified by pyrene derivative as an energy donor and cationic porphyrin is used as adsorbing molecule (energy acceptor) on the clay surface. By observing the energy transfer efficiency for the sample where the loading level of porphyrin was changed, the adsorption distribution of porphyrin on the clay surface was estimated. In this research, to estimate the adsorption distribution, distinctive three models were assumed as shown in Fig. 1.

Those are i) random adsorption, ii) from edge to center and iii) from center to edge of

porphyrins on the clay surface.

The clarification of adsorption assembly structure on the

inorganic surface is essential to construct photochemical reaction system such as artificial light harvesting systems. 30-39

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 4 of 25

Page 5 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Analysis UV-Visible absorption spectra were obtained on JASCO V-650 spectrophotometer. Fluorescence spectra were recorded on JASCO FP-6500 spectrophotometer. TG/DTA measurements were carried out with a SHIMADZU DTA-60H analyzer to determine the water content of the clay and p-TMPyP.

Synthesis of pyrene modified clay nanosheet Preparation

of

pyrene

modified

silane

coupling

agent

(N-(3-(dimethylethoxysilyl)propyl)-3-(1-pyrenyl)propanamide&APES-Py'

Silane-coupling agent, 3-aminopropyldimethylethoxysilane (APES) was used to modify the edge of the clay.

APES and 1-pyrenebuthylic acid were condensed by amide bond formation.

DMTMM was selected as a condensation agent. In a 50 mL flask, DMTMM 122.3 mg was dissolved in 2 mL distilled water.

Then alumina

6 g was added to the solution. In another flask, 1-pyrenebutyric acid 116.6 mg was dissolved in dichloromethane 8 mL and then APES 75 µL was added to the solution. The obtained dichloromethane solution was added to the DMTMM-alumina solid, and was stirred at room temperature for 4.5 h, according to a previously reported method. 40 The solution was filtrated by OmniporeTM Membrane Filters 0.1 µm, JVWP04700, the alumina on the filter was washed with dichloromethane (ca. 10 mL). The filtrate was evaporated, and the obtained solid was purified by column chromatography (alumina, ethyl acetate : toluene (1 : 1 (v/v))). The yield of N-(3-(dimethylethoxysilyl)propyl)-3-(1-pyrenyl)propanamide&APES-Py'was 45% (77.6 mg).

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 25

Elemental analysis for C27H33NO2Si: Found: C, 74.99; H, 7.44; N, 2.99. Calculated: C, 75.13; H, 7.71; N, 3.24. The 1H-NMR of APES-Py is shown in Fig. S2. Modification of clay nanosheet with APES-Py The condensation reaction between terminal hydroxyl group at clay edge sites and APES-Py was performed according to a previous report. 41 A toluene solution (6 mL) of APES-Py 3.54 mg was refluxed for 18.5 h with SSA101.5 mg. After the reaction, the modified clay was filtrated by OmniporeTM Membrane Filters 0.1 µm, JVWP04700. To remove the unreacted APES-Py, the obtained solid was washed with 10 mL of ethanol 14 times. The pyrene modified clay nanosheet (Py-SSA) was dried in vacuum for 24 h. The yield was 66% (68.9 mg).

The characterization is

described in Results and Discussion. Measurement of energy transfer reaction TG/DTA measurements were carried out with a SHIMADZU DTA-60H analyzer to determine the water content of p-TMPyP and Py-SSA, then the stock solutions were prepared. The -5

concentration of p-TMPyP and Py-SSA stock solution were 1.0!10 M and 1.0 ! 10-5 eq L-1, respectively. The sample, where the total volume was 3000 µL, the concentration of Py-SSA was 4.0!10-6 eq L-1 and the loading level of p-TMPyP was 0 to 50 % vs CEC, was prepared in a quartz cell.

Fluorescence spectra were measured for each sample, where the excitation

wavelength is 275 nm.

The energy transfer efficiency was calculated as described later.

Results and Discussion

ACS Paragon Plus Environment

6

Page 7 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 8 of 25

Page 9 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 25

Because the total energy transfer rate constant should increase as the number of p-TMPyP increased, the tendency observed in Fig. 3 and 4 is reasonable.

On the other hand, the plot in

Fig. 4 includes important information on the position distribution of p-TMPyP on the clay surface.

For example, if p-TMPyP adsorbed on the clay surface from edge to center, sudden

increase of "ET at low loading level region of p-TMPyP should be observed in Fig. 4.

In next

section, the adsorption structure of p-TMPyP will be discussed by comparing observed "ET and theoretical one for possible three adsorption distribution models. Estimation of p-TMPyP adsorption distribution on the clay surface To discuss the adsorption distribution of p-TMPyP, possible three models were assumed. Three adsorption models were i) random adsorption, ii) from edge to center and iii) from center to edge adsorption of p-TMPyP on the clay surface (shown in Fig. 1). As an example, the procedure to estimate the theoretical energy transfer efficiency (!"# ) for the case i) random adsorption, is described below.

The calculations of !"# for type ù) and iii) are described in

SI. To estimate the adsorption distribution of p-TMPyP on the clay surface, the calculation model which theoretically estimates the relationship between the number of adsorbed molecules and the energy transfer efficiency (!"# ), was constructed.

The structural model of clay

nanosheet / pyrene / p-TMPyP complex for the calculation of !"# is shown in Fig. 5.

It is

supposed that i) the shape of SSA is a circular cylinder whose diameter and thickness are 50 nm and 0.96 nm, respectively, ii) p-TMPyP adsorbs on the both side of the clay nanosheet as a hexagonal arrangement, where the average intermolecular distance is 2.4 nm,13,14

iii) pyrene

locates at the edge of nanosheet where the distance between the center of pyrene and the

ACS Paragon Plus Environment

10

Page 11 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

nanosheet edge is 0.42 nm that is radial of pyrene.

The area of a diamond composed of four

negative charges shown in Fig. 5 is defined as an adsorption site. The energy transfer rate constant (kET) between the adsorbed porphyrin and the edge modified pyrene was calculated by Eq. 2.

0"# $

1222..34%2.5 6 7+ .E................-8/ %89.: ; 4 ? @A B+ CD

, where #D is the fluorescence quantum yield of the energy donor (pyrene), $D is the excited lifetime of the donor (101 ns)46,47 , R is the distance between donor molecule and acceptor molecule, % is the orientation parameter, nsol is the refractive index of the bulk medium (1.33, water), NA is the Avogadro constant and J is the integral overlap between the fluorescence spectra of donor and the absorption spectra of acceptor. The J value was calculated from observed fluorescence and absorption spectra of pyrene and p-TMPyP, respectively (J = 1.0 ! 10-13 M-1 cm3 ). #Dwas measured by the relative method used anthracene as the standard sample (#D = 0.63). The orientation parameter $2 was supposed to be 2/3 that is the average value of the possible orientation.

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 12 of 25

Page 13 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The theoretical energy transfer efficiency !"# HIJKL M 84 when 2n porphyrins adsorb on clay nanosheet was evaluated by eq. 3.

!"# HIJKL M 84 $

P( NQR !"#-N/ .O-N/

.....-3)

, where !"#-N/ is the energy transfer efficiency for the arrangement i, O-N/ is the appearance probability for the arrangement i and ND is the number of all arrangement i including both sides of clay nanosheet.

!"#-N/ was calculated according to eq. 4.

!"#-N/ $

0"#-N/ %

B+ S 0"#-N/

.....-T/

, where 0"#-N/ is the energy transfer rate constant for the arrangement i, representing the sum of energy transfer rate constants between porphyrin adsorbed within FG and the excited edge modified pyrene. The appearance probability O-N/ for arrangement i was estimated as follows. The number of adsorption sites per one side of clay nanosheet was defined as @U . The number of arrangements when n porphyrins adsorbed on one side of clay nanosheet is

NsCn.

Therefore, the

number of all arrangements i including both sides of clay nanosheet @+ is -P< VW /6 .

For

example, when the porphyrins adsorb on the clay nanosheet whose diameter is 50 nm at the 1% adsorption density vs. CEC, n and @U is 4 ( = 392 (total number of adsorption site) ! 0.01) and 392, respectively.

Then the number of arrangement pattern @+ was about 1 !1018 in this case.

For the random adsorption, the appearance probability O-N/ is described as 1/ @+ because each arrangement should be appeared with same probability.

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 14 of 25

Page 15 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Discussion on the adsorption distribution of p-TMPyP on the clay surface It is known that the surface of clay nanosheet is uniformly flat and the substitution site by Al3+, that determines negative charge distribution, is uniformly distributed. 1, 13, 48-50 Nevertheless, the adsorption deviation of p-TMPyP that is biased to center of the clay sheet was observed in the present system. nanosheet?

Why does the adsorption of p-TMPyP distribute near the center of clay

It has been reported that the motivation of adsorption is electrostatic and

hydrophobic interaction in the case of cationic porphyrins.51 Although the strength of both interaction seems not to depend on the position (center or near edge) on the clay surface, we aware that the total electrostatic interaction energy between p-TMPyP and negative charges in the clay structure should depend on the p-TMPyP’s position on the clay surface. The electrostatic potential between univalent cation and the clay nanosheet was calculated. The electrostatic potential map image is shown in Fig. 7, supposing that i) the diameter of clay is 50 nm, ii) the negative charge distribution is uniform and the charge density is 0.8 nm-2, iii) the dielectric constant " is 43.2 (the dielectric constant of interface between clay and water is calculated by electric image method52 ; "clay: 6, "water: 80.4), iv) counter cations are dissociated from the nanosheet surface, and iv) the height h from clay surface to univalent cation is assumed to be 0.3 nm.

Interestingly, the potential map shows the shape of mortar. This calculation

results clearly indicate that the electrostatic interaction could be a motivation for the center-oriented adsorption distribution of p-TMPyP, although the quantitative accuracy of Fig. 7 is not high due to the difficulty in the estimation of the dielectric constant and h.

Thus, the

adsorption distribution of p-TMPyP is rationalized to be biased to the center of the clay nanosheet surface.

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 16 of 25

Page 17 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

spectroscopic molecular ruler.

Such adsorption distribution of p-TMPyP is rationalized by the

Coulomb energy potential calculation for the clay surface. These findings are beneficial to construct organic-inorganic hybrids where the assembly structure of adsorbed molecules is controlled, especially for two or more species co-adsorbed systems.

ASSOCIATED CONTENT Supporting Information. 1.! Purification of clay nanosheet 2.! Synthesis of pyrene modified clay nanosheet 3.! Control experiment of energy transfer reaction (between free pyrene to SSA-p-TMPyP complex) 4.! Details for theoretical calculation of energy transfer efficiencies This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author *[email protected]

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 25

ACKNOWLEDGMENT This work has been partly supported by a Grant-in-Aid for Scientific Research (B) (No. 24350100) and a Grant-in-Aid for Scientific Research on Innovative Areas “All Nippon Artificial Photosynthesis Project for Living Earth” (AnApple, No. 25107521).

REFERENCES 1.! Sasai, R.; Itoh, H.; Shindachi, I.; Shichi, T.; Takagi, K. Photochromism of Clay'Diarylethene Hybrid Materials in Optically Transparent Gelatin Films. Chem. Mater. 2001, 13, 2012–2016. 2.! Shichi, T.; Takagi, K. Clay Minerals as Photochemical Reaction Fields. J. Photochem. Photobiol., C 2000, 1, 113–130. 3.! Okada, T.; Ide, Y.; Ogawa, M. Organic–Inorganic Hybrids Based on Ultrathin Oxide Layers: Designed Nanostructures for Molecular Recognition. Chem. - Asian J. 2012, 7, 1980 – 199. 4.! Ogawa, M.; Kuroda, K. Preparation of Inorganic–Organic Nanocomposites through Intercalation of Organoammonium Ions into Layered Silicates. Bull. Chem. Soc. Jpn. 1997, 70, 2593–2618. 5.! Bujdak, J. Effect of the Layer Charge of Clay Minerals on Optical Properties of Organic Dyes. Appl. Clay Sci. 2006, 34, 58–73.

ACS Paragon Plus Environment

18

Page 19 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

6.! Bujdak, J.; Iyi, N.; Kaneko, Y.; Czımerova, A.; Sasai, R. Molecular Arrangement of Rhodamine 6G Cations in the Films of Layered Silicates: The Effect of the Layer Charge.

Phys. Chem. Chem. Phys. 2003, 5, 4680–4685 7.! Carbonaro, C. M.; Ricci, P. C.; Grandi, S.; Marceddu, M.; Corpino, R.; Salisa, M.; Aneddaab, A. On the Formation of Aggregates in Silica–Rhodamine 6G Type II Hybrids. RSC Adv. 2012, 2, 1905–1912. 8.! Bujdak, J.; Komadel, P. Interaction of Methylene Blue with Reduced Charge Montmorillonite. J. Phys. Chem. B, 1997, 101, 9065–9068. 9.! Li, X.; Sinks, L. E.; Rybtchinski, B.; Wasielewski, M. R.; Ultrafast Aggregate-to-Aggregate Energy Transfer within Self-assembled Light-Harvesting Columns of Zinc Phthalocyanine Tetrakis(Perylenediimide). J. Am. Chem. Soc. 2004, 126, 10810–10811. 10.! Rao, K. V.; Datta, K. K. R.; Eswaramoorthy, M.; George, S. J. Light-Harvesting Hybrid Assemblies. Chem.- Eur. J. 2012, 18, 2184–2194. 11.! Sato, H.; Ochi, M.; Kato, M.; Tamura, K.; Yamagishi, A. Energy Transfer in Hybrid Langmuir–Blodgett Films of Iridium Complexes and Synthetic Saponite: Dependence of Transfer Efficiency on the Interlayer Distance. New J. Chem. 2014, 38, 5715–5720. 12.! Yui, T.; Kameyama, T.; Sasaki, T.; Torimoto, T.; Takagi, K. Pyrene-To-Porphyrin Excited Singlet Energy Transfer in LBL-Deposited LDH Nanosheets. J. Porphyrins Phthalocyanines 2007, 11, 428–433. 13.! Takagi, S.; Shimada, T.; Fujimura, T.; Masui, D.; Tachibana, H.; Eguchi, M.; Inoue, H. Size-Matching Effect on Inorganic Nanosheets: Control of Distance, Alignment, and

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 25

Orientation of Molecular Adsorption as a Bottom-Up Methodology for Nanomaterials. Langmuir 2013, 29, 2108–2119. 14.! Takagi, S.; Shimada, T.; Eguchi, M.; Yui, T.; Yoshida, H.; Donald A. T.; Inoue, H. High-Density Adsorption of Cationic Porphyrins on Clay Layer Surfaces without Aggregation:( The Size-Matching Effect. Langmuir 2002, 18, 2265–2272. 15.! Tsukamoto, T.; Ramasamy, E.; Shimada, T.; Takagi, S.; Ramamurthy, V. Supramolecular Surface Photochemistry: Cascade Energy Transfer between Encapsulated Dyes Aligned on a Clay Nanosheet Surface. Langmuir 2016, 32, 2920–2927. 16.! Ohtani, Y.; Shimada, T.; Takagi, S. Artificial Light-Harvesting System with Energy Migration Functionality in a Cationic Dye/Inorganic Nanosheet Complex. J. Phys. Chem. C 2015, 119, 18896–18902. 17.! Ohtani, Y.; Kawaguchi, S.; Shimada, T.; Takagi, S. Energy Transfer among Three Dye Components in a Nanosheet–Dye Complex: An Approach To Evaluating the Performance of a Light-Harvesting System. J. Phys. Chem. C 2017, 121, 2052–2058. 18.! Tatsumi, D.; Tsukamoto, T.; Honna, R.; Hoshino, S.; Shimada, T.; Takagi, S. Highly Selective Photochemical Epoxidation of Cyclohexene Sensitized by Ru(II) Porphyrin–Clay Hybrid Catalyst. Chem. Lett. 2017, 46, 1311–1314. 19.! Villemure, G.; Detellier, C.; Szabo, A. Z. Fluorescence of Clay-Intercalated Methylviologen. J. Am. Chem. Soc. 1986, 108, 4658–4659.

ACS Paragon Plus Environment

20

Page 21 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

20.! Villemure, G.; Detellier, C.; Szabo, A. Z. Fluorescence of Methylviologen Intercalated into Montmorillonite and Hectorite Aqueous Suspensions. Langmuir 1991, 7, 1215–1221. 21.! Tokieda, D.; Tsukamoto, T.; Ishida, Y.; Ichihara, H.; Shimada, T.; Takagi, S. Unique Fluorescence Behavior of Dyes on the Clay Minerals Surface: Surface Fixation Induced Emission (S-FIE). J. Photochem. Photobiol., A 2017, 339, 67–79. 22.! Tsukamoto, T.; Shimada, T.; Takagi, S. Unique Photochemical Properties of p-Substituted Cationic Triphenylbenzene Derivatives on a Clay Layer Surface. J. Phys. Chem. C 2013, 117, 2774–2779. 23.! Ishida, Y.; Shimada, T.; Takagi, S. FSurface-Fixation Induced Emission” of Porphyrazine Dye by a Complexation with Inorganic Nanosheets. J. Phys. Chem. C 2014, 118 20466'20471. 24.! Konno, S.; Fujimura, T.; Ohtani, Y.; Shimada, T.; Takagi, S. Microstructures of the Porphyrin/Viologen Monolayer on the Clay Surface: Segregation or Integration? J. Phys. Chem. C 2014, 118, 20504–20510. 25.! Shimada, T.; Hamatani, S.; Onodera, S.; Inoue, Y.; Takagi, S. Investigation of Adsorption Behavior and Energy Transfer of Cationic Porphyrins on Clay Surface at Low Loading Levels by Picosecond Time-Resolved Fluorescence Measurement. Res. Chem. Intermed. 2013, 39, 269–278. 26.! Ghosh, P. K.; Bard, A. J. Photochemistry of Tris(2,2'-Bipyridyl)Ruthenium(II) in Colloidal Clay Suspensions. J. Phys. Chem. 1984, 88, 5519–5526.

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 25

27.! Lagaly, G.; Ziesmer, S. Colloid Chemistry of Clay Minerals: The Coagulation of Montmorillonite Dispersions. Adv. Colloid Interface Sci. 2003, 100 –102, 105–128. 28.! Felbeck, T.; Hoffmann, K.; Resch, G.; Lezhnina, M.; Kynast, U. Fluorescent Nanoclays: Covalent Functionalization with Amine Reactive Dyes from Different Fluorophore Classes and Surface Group Quantification. J. Phys. Chem. C 2015, 119, 12978–12987. 29.! Herrera, N. N.; Letoffe, J.-M.; Reymond, J.-P.; Bourgeat-Lami, E. Silylation of Laponite Clay Particles with Monofunctional and Trifunctional Vinyl Alkoxysilanes. J. Mater. Chem. 2005, 15, 863–871. 30.! Takeda, H.; Ohashi, M.; Tani, T.; Ishitani, O.; Inagaki, S. Enhanced Photocatalysis of Rhenium(I) Complex by Light-Harvesting Periodic Mesoporous Organosilica. Inorg. Chem. 2010, 49, 4554–4559. 31.! Mizoshita, N.; Yamanaka, K.; Hiroto, S.; Shinokubo, H.; Tani, T.; Inagaki, S. Energy and Electron Transfer from Fluorescent Mesostructured Organosilica Framework to Guest Dyes. Langmuir 2012, 28, 3987–3994. 32.! Oohora, K.; Mashima, T.; Ohkubo, K.; Fukuzumi, S.; Hayashi, T. Energy Migration Within Hexameric Hemoprotein Reconstituted with Zn Porphyrinoid Molecules. Chem. Commun. 2015, 51, 11138–11140. 33.! Hayashi, T.; Ogoshi, H. Molecular Modelling of Electron Transfer Systems by Noncovalently Linked Porphyrin–Acceptor Pairing. Chem. Soc. Rev. 1997, 26, 355–364.

ACS Paragon Plus Environment

22

Page 23 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

34.! Shibata, S.; Tsuge, K.; Sasaki, Y.; Ishizaka, S.; Kitamura, N. Directional Energy Transfer in Mixed-Metallic Copper(I)–Silver(I) Coordination Polymers with Strong Luminescence. Inorg. Chem. 2015, 54, 9733–9739. 35.! Sumino, A.; Dewa, T.; Kondo, M.; Morii, T.; Hashimoto, H.; Gardiner, A. T.; Cogdell, R. J.; Nango, M. Selective Assembly of Photosynthetic Antenna Proteins into a Domain-Structured Lipid Bilayer for the Construction of Artificial Photosynthetic Antenna Systems: Structural Analysis of the Assembly Using Surface Plasmon Resonance and Atomic Force Microscopy. Langmuir 2011, 27, 1092–1099. 36.! Sakamoto, R.; Nishikawa, M.; Yamamura, T.; Kume, S.; Nishihara, H. A New Special Pair Model Comprising Meso-Di-P-Anisylaminoporphyrin: Enhancement of Visible-Light Absorptivities and Quantification of Electronic Communication in Mixed-Valent Cation Radical. Chem. Commun. 2010, 46, 2028–2030. 37.! Saga, Y.; Shibata, Y.; Tamiaki, H. Spectral Properties of Single Light-Harvesting Complexes in Bacterial Photosynthesis. J. Photochem. Photobiol., C 2010, 11, 15–24. 38.! Karkas, M. D.; Verho, O.; Johnston E. V.; Akemark, B. Artificial Photosynthesis: Molecular Systems for Catalytic Water Oxidation. Chem. Rev. 2014, 114, 11863–12001. 39.! Gust, D.; Moore, T. A.; Moore, A. L. Solar Fuels via Artificial Photosynthesis. Acc. Chem. Res. 2009, 42, 1890–1898. 40.! Watanabe, Y.; Fuji, T.; Hiroki, K.; Tani, S.; Kunishima, M. Development of a Simple System for

Dehydrocondensation

Using

Solid-Phase

Adsorption

of

a

Water-Soluble

Dehydrocondensing Reagent (DMT-MM). Chem. Pharm. Bull. 2004, 52, 1223–1226.

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 25

41.! Bujdák, J.; Danko, M.; Chorvát, D., Jr.; Czímerová, A.; S)kora, J.; Lang, K. Selective Modification of Layered Silicate Nanoparticle Edges with Fluorophores. Appl. Clay Sci. 2012, 65-66, 152–157. 42.! Andrew, G. C.; Austin, D. D.; Zhiqiang, L.; Andreas, S.; Andrew, B.; Lars-Olof, P.; David, J. T.; Todd, B. M. Experimental and Theoretical Studies of the Photophysical Properties of 2and 2,7-Functionalized Pyrene Derivatives. J. Am. Chem. Soc. 2011, 133, 13349–13362 43.! The molar extinction coefficient (") of 1-pyrenebutyric acid at 275 nm was determined to be 5.1 !104 L mol-1 cm-1 in ethanol. 44.! The number of SSA particles = 750 [g m-2] ! 1.10!10-4 [eq L-1] ! 25!10-3 [L] / (9.97!10-4 [eq g-1] ! (# ! 252 ! 2!10-18) [m2]). 45.! The number of pyrene molecule at the edge of one SSA particle = 8.3!10-7 [mol/L] ! 25!10-3[L] ! 6.02!1023 [mol-1] / 5.3!1014 . 46.! Sandoval, C.; Arriagada, F. S.; De la Fuente, J. R.; Sanchez, S. A.; Morales, J.; Pizarro, N.; Nonell, S.; Gunther, G. Synthesis, Physicochemical and Photophysical Characterization of 4-(1-Pyrenyl)-buthyl-

-D-mannopyranoside. J. Braz. Chem. Soc. 2016, 27, 402-410.

47.! Matsui, J.; Mitsuishi, M.; Miyashita, T. A Study on Fluorescence Behavior of Pyrene at the Interface of Polymer Langmuir-Blodgett Films. J. Phys. Chem. B 2002, 106, 2468-2473. 48.! Tamura, K.; Yamagishi, A.; Kitazawa, T.; Sato, H. Harvesting Light Energy by Iridium(III) Complexes on a Clay Surface. Phys. Chem. Chem. Phys. 2015, 17, 18288–18293.

ACS Paragon Plus Environment

24

Page 25 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

ACS Paragon Plus Environment