exploring the blinking dynamics of eosin y

44
W&M ScholarWorks W&M ScholarWorks Undergraduate Honors Theses Theses, Dissertations, & Master Projects 4-2019 Exploring the Blinking Dynamics of Eosin Y Photosensitizers for Exploring the Blinking Dynamics of Eosin Y Photosensitizers for Dye-Sensitized Photocatalysis Using Single-Molecule Dye-Sensitized Photocatalysis Using Single-Molecule Spectroscopy Spectroscopy Pauline Lynch Follow this and additional works at: https://scholarworks.wm.edu/honorstheses Part of the Physical Chemistry Commons Recommended Citation Recommended Citation Lynch, Pauline, "Exploring the Blinking Dynamics of Eosin Y Photosensitizers for Dye-Sensitized Photocatalysis Using Single-Molecule Spectroscopy" (2019). Undergraduate Honors Theses. Paper 1299. https://scholarworks.wm.edu/honorstheses/1299 This Honors Thesis is brought to you for free and open access by the Theses, Dissertations, & Master Projects at W&M ScholarWorks. It has been accepted for inclusion in Undergraduate Honors Theses by an authorized administrator of W&M ScholarWorks. For more information, please contact [email protected].

Upload: others

Post on 22-Feb-2022

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Exploring the Blinking Dynamics of Eosin Y

W&M ScholarWorks W&M ScholarWorks

Undergraduate Honors Theses Theses, Dissertations, & Master Projects

4-2019

Exploring the Blinking Dynamics of Eosin Y Photosensitizers for Exploring the Blinking Dynamics of Eosin Y Photosensitizers for

Dye-Sensitized Photocatalysis Using Single-Molecule Dye-Sensitized Photocatalysis Using Single-Molecule

Spectroscopy Spectroscopy

Pauline Lynch

Follow this and additional works at: https://scholarworks.wm.edu/honorstheses

Part of the Physical Chemistry Commons

Recommended Citation Recommended Citation Lynch, Pauline, "Exploring the Blinking Dynamics of Eosin Y Photosensitizers for Dye-Sensitized Photocatalysis Using Single-Molecule Spectroscopy" (2019). Undergraduate Honors Theses. Paper 1299. https://scholarworks.wm.edu/honorstheses/1299

This Honors Thesis is brought to you for free and open access by the Theses, Dissertations, & Master Projects at W&M ScholarWorks. It has been accepted for inclusion in Undergraduate Honors Theses by an authorized administrator of W&M ScholarWorks. For more information, please contact [email protected].

Page 2: Exploring the Blinking Dynamics of Eosin Y
Page 3: Exploring the Blinking Dynamics of Eosin Y

Abstract

Dye-sensitized photocatalysis (DSP) is a promising way to harvest solar energy for

carbon-neutral fuel production, but a better understanding of how and why it is currently

inefficient is necessary. This thesis will delve into the complex excited-state dynamics of Eosin

Y (EY), a sensitizer for DSP, on glass substrates. By using single-molecule spectroscopy (SMS)

to understand the underlying photophysics at play, we can gain a more complete understanding

of the various photophysical events that contribute to inefficient DSP. In particular, SM blinking

dynamics give insight into kinetic models.

SM blinking measurements of EY molecules in air and in N2 were separated into on- and

off-time distributions and fit to heavy-tailed functions using the robust combined Maximum

Likelihood Estimation (MLE) and Kolmogorov Statistic (KS) method. Both on-time

distributions in air and N2 are power law distributed after an onset time. The off time distribution

in air is best fit to a lognormal function, consistent with the Albery model for dispersive electron

transfer. The off time distribution for N2, however, contained contributions from both an

exponential and lognormal function with an onset time, consistent with a model where both

intersystem crossing and dispersive electron transfer occur. The off-time distribution of an

individual molecule was lognormally distributed, consistent with dispersive ET kinetics.

Additionally, blinking dynamics were investigated as a function of laser excitation

power, revealing a power-dependence in the on times of individual EY molecules. Furthermore,

preliminary studies of EY on TiO2 exhibit significant visual differences from blinking dynamics

on glass. In future studies, the blinking dynamics and kinetics of EY on TiO2 will be explored in

order to gain a more complete understanding of ET in technologically relevant conditions for the

design and development of next-generation DSP solar cells.

Page 4: Exploring the Blinking Dynamics of Eosin Y

i

Table of Contents

List of Figures ................................................................................................................................. ii List of Tables .................................................................................................................................. ii Acknowledgements ........................................................................................................................ iii Chapter 1: Introduction ............................................................................................................... 1 1.1 Motivation ................................................................................................................................. 1 1.2 Single-Molecule Spectroscopy ................................................................................................. 4 1.3 Methods ..................................................................................................................................... 6 1.3.1 Sample Preparation .................................................................................................... 6 1.3.2 Confocal Microscopy ................................................................................................. 7

1.3.3 Blinking Analysis ....................................................................................................... 8 1.4 Outline ..................................................................................................................................... 11

Chapter 2: Unraveling the Excited-State Dynamics of EY Photosensitizers Using SMS .... 13 2.1 Introduction ............................................................................................................................. 13 2.2 Blinking Dynamics in Anoxic Environments ......................................................................... 14 2.3 Impact of Oxygen on Blinking Dynamics .............................................................................. 17 2.4 Discussion ............................................................................................................................... 19

2.4.1 Albery Model for Power Law and Lognormal Distributions ................................... 19 2.4.2 Kinetic Expressions for Triplet Blinking and Dispersive Electron Transfer ........... 21 2.4.3 Modeling Competitive Photophysical Processes ..................................................... 23

2.5 Conclusion .............................................................................................................................. 28 Chapter 3: Secondary Experiments .......................................................................................... 303.1 Power Dependent Blinking Dynamics .................................................................................... 30 3.2 Blinking Dynamics on TiO2 .................................................................................................... 32 References ..................................................................................................................................... 34

Page 5: Exploring the Blinking Dynamics of Eosin Y

ii

List of Figures

1.1 Schematic of DSP System ........................................................................................................ 2 1.2 Eosin Y Structure ...................................................................................................................... 3 1.3 Blinking Schematic and Trace .................................................................................................. 5 1.4 Confocal Microscopy Set-up .................................................................................................... 7 2.1 Dye Film Scan and Blinking Trace in N2 ............................................................................... 14 2.2 On- and Off-time Distributions in Air and N2 and Best-Fit Parameters ................................ 15 2.3 Dye Film Scan and Blinking Trace in Air .............................................................................. 17 2.4 1000 s Blinking Trace in Air .................................................................................................. 18 2.5 Proposed Kinetic Models ........................................................................................................ 20 2.6 Off-Time Probability Distribution of EY in N2 with Best-Fit Functions .............................. 23 2.7 1-ms and Binned Up Blinking Traces ..................................................................................... 25 3.8 On- and Off-time Distributions of an Individual Molecule .................................................... 28 3.1 On-time Distributions with Varying Excitation Power .......................................................... 31 3.2 TiO2 Dye Film Scans and Blinking Traces in N2 and Air ...................................................... 27

List of Tables 2.1 Best Fit Parameters for EY on Glass ...................................................................................... 14 3.1 Best Fit Parameters for Power-Dependent Blinking ............................................................... 28

Page 6: Exploring the Blinking Dynamics of Eosin Y

iii

Acknowledgements First, I would like to acknowledge the Charles Center at The College of William & Mary

for providing funding during 2018 through an Honors Fellowship, as well as the National

Science Foundation Award #1664828 for providing funding as well. This work was performed

in part using computing facilities at the College of William & Mary which were provided by

contributions from the National Science Foundation, the Commonwealth of Virginia

Equipment Trust Fund and the Office of Naval Research.

I would like to thank Dr. Kristin Wustholz for her mentorship and guidance during my

undergraduate career. She has been instrumental in my development as a scientist and researcher,

and has taught me much more than I ever could have expected when I first walked into her office

as a freshman. I would like to recognize Huw Richards for his significant contributions to the

numerical analysis, as well as John Li and Kelly Kopera for their work on preliminary TiO2

studies. A big thanks also goes to the entire Wustholz group, past and present, for their guidance,

support, and friendship. I am very fortunate to have been part of such a dynamic, smart,

passionate group, and they have made every day in lab an adventure. In particular, thank you to

Carolyn Farling for always being ready with words of encouragement or a laugh when I needed

itβ€” she is truly my lab soul-sister and I don’t know what I would have done without her. Thank

you to my family and friends for their unending love, support, and encouragement; this work has

been an undertaking that would have been impossible without it.

Page 7: Exploring the Blinking Dynamics of Eosin Y

1

Chapter 1: Introduction

1.1 Motivation

By 2050, the global energy demand is predicted to grow to at least 27.6 terawatts (TW),

with some estimates as high as 106 TW.1,2 The world increasingly relies on fossil fuels, usage of

which has increased in recent years and continues to rise, leading to worsening and irreversible

impacts on the environment and communities across the globe.3 As the contribution of burning

fossil fuels to climate change is better understood, there is an urgent need for renewable,

affordable, efficient, and carbon-neutral energy sources.1,4 Solar energy is an abundant,

renewable resource, and is particularly attractive because enough sunlight hits the surface of the

earth each year to exceed current and future energy needs.5 Unfortunately, current technology for

harnessing solar energy is inefficient and too expensive to implement widely. Research is needed

to fill this void and develop renewable energy technologies that can provide an alternative to

fossil fuels.

Dye-sensitized photocatalysis (DSP) (Fig. 1.1) shows promise as a renewable energy

source; it utilizes a chromophore, semiconductor, and catalyst to produce fuel from sunlight.2,6,7

DSP systems are heavily inspired by dye-sensitized solar cells (DSSCs).6 Both utilize an organic

dye adsorbed onto a semiconductor in order to capture the Sun’s energy. DSP systems are more

favorable than DSSCs because they addresses issues of energy storage and transportation by

producing carbon-neutral fuel.2 DSSCs and DSP systems utilize the same principles at the dye-

semiconductor interface, but utilize the excited electrons differently. In DSSC, photoexcited

electrons are converted directly into electricity. DSP, however, co-adsorbs an organic dye

molecule and a catalyst onto a semiconductor. The photoexcited electrons are first injected into

the conduction band (CB) of a semiconductor, and then their energy is used to catalyze fuel

Page 8: Exploring the Blinking Dynamics of Eosin Y

2

production reactions.6 Hydrogen fuel (H2) is produced via water splitting reactions, in which two

water molecules are broken into oxygen and hydrogen (2H2OΓ  O2 + 2H2). To produce O2 and

release four protons and four electrons,

the hydroxyl bonds of two water

molecules must be broken. The

protons and electrons subsequently

combine to produce H2 or can combine

with CO2 to produce hydrocarbon fuels

in synthetic photosynthesis reactions.2

Fujishima and Honda were the first to develop a photo-electrochemical cell that decomposed

water to oxygen and hydrogen using visible light without utilizing an external voltage.8 Since

then, many advances have been made to electrochemical systems.

Unfortunately, DSP systems are currently inefficient due to several reasons. First, they

must operate in aqueous environments in order to undergo water splitting for fuel production,

which complicates the electron-transfer (ET) dynamics at the dye-semiconductor interface.

Aqueous environments are inherently heterogeneous in nature. As a result, the ET kinetics can

differ widely from molecule to molecule. Additionally, the thermodynamics and kinetics of the

associated chemical reactions must be accounted for as well when considering the ET dynamics

and solar cell efficiency. DSP kinetics are further complicated due to the coupling of the short

timescales of photoexcitation (i.e., ps-ns) and long timescales of fuel production in the

electrocatalyst (i.e., ms-s). Because the chemical reactions associated with water splitting and

fuel production occur at a much slower rate than electron generation, a long-lived photo-excited

state is necessary for the electron to be injected into the electrocatalyst. Low yields for solar fuel

Figure 1.1: Schematic of a DSP system in which a dye molecule’s excited electron is injected into the conduction band (CB) of TiO2 and then into a catalyst for hydrogen fuel production.

Page 9: Exploring the Blinking Dynamics of Eosin Y

3

production in DSP can be attributed in part to charge recombination, where an excited electron

falls back down to the dye’s ground state and is unable to inject into the catalyst.9 As a result of

these complications and others, current DSPs exhibit low efficiencies and the underlying

photophysics contributing to DSP performance are not well understood.6,10

In order to efficiently produce solar fuel, DSP systems must be able to prolong the

lifetime of charge carriers.9,11 Ideally, ET from the dye-loaded semiconductor to the catalyst,

with no recombination, would enhance the long-lived excited state and allow time for catalysis to

occur.11 One way to extend the lifetime of the dye’s excited state is by employing chromophores

containing heavy atoms, which can undergo intersystem

crossing (ISC) to a long-lived triplet state. ISC is a

radiationless transition between electronic states of different

multiplicities and the presence of heavy atoms is known to

enhance ISC. The organic dye eosin y (EY) (Fig. 1.2) shows

promise as a sensitizer for DSP since it can undergo ISC to a

long-lived triplet state.12,13 Indeed, in 1984, Moser and

GrΓ€tzel studied aqueous EY and colloidal TiO2 systems and found that photoinduced charge

separation could achieve a lifetime of several microseconds in the semiconductor, allowing time

for the electrons to be injected into the catalyst.14 It has been used widely in catalytic water

splitting,12,13,15–19 as well as photoredox synthesis.20,21 However, little is known about ET

processes in DSP, and a recent review has expressed the necessity for thorough investigation into

ET in DSP systems.6 While it is well understood that triplet state population and depopulation

via ISC follows first order kinetics, the impacts of substrate, dye structure, and the heavy atom

effect on ET kinetics in DSP systems are not well understood. The goal of this research is to

Figure 1.2: Eosin Y (EY)

Page 10: Exploring the Blinking Dynamics of Eosin Y

4

elucidate the ET kinetics of EY under technologically relevant conditions using single-molecule

spectroscopy (SMS).

1.2 Single-Molecule Spectroscopy

SMS is an incredibly useful tool for understanding the full distribution of behaviors in

heterogeneous environments, like those occurring in DSP. By probing individual molecules

instead of ensemble averages, SMS provides the ability to reveal the various contributions of ET

kinetics to the overall behavior of DSP systems. Photophysics at the dye-semiconductor interface

can differ greatly between molecules as the result of changes in molecular orientation relative to

other molecules, the semiconductor surface or catalyst. Molecules will behave differently based

on their surroundings, and SMS provides the capability to examine how local environment

impacts ET kinetics. Each molecule has a distinct free energy, electronic coupling, and distance

from the semiconductor, all of which contribute to a molecules’ ET behavior in a solar cell.22

These details, otherwise buried within ensemble average techniques, are critical to

understanding photophysics within solar cells.

Single-molecule emitters have been studied using a method called blinking, a

phenomenon in which a molecule undergoes stochastic changes in emissive and non-emissive

intensity under continuous laser excitation. Blinking essentially measures changes in fluorescent

intensity as a function of time. Once an electron is promoted to an excited state via

photoexcitation (Pathway a in Fig. 1.3A), a molecule can undergo fluorescence and fall back to

the ground state (Pathway b in Fig. 1.3A) or transfer (Pathway c in Fig. 1.3A) into a β€˜dark,’ or

non-emissive, state (3 in Fig. 1.3A). From the dark state, the excited electron can undergo back

electron transfer (BET) (Pathway d in Fig. 1.3A) back to the ground state. Once back in the

Page 11: Exploring the Blinking Dynamics of Eosin Y

5

ground state, the ground state dye can be

excited again. Successive cycles of

excitation, fluorescence, charge transfer,

and BET produce a series of emissive and

non-emissive events, which together form

a blinking trace (Fig. 1.3B). A single

molecule can be photobleached, where

some photochemistry will occur and the

molecule will no longer be able to

fluoresce and it will remain in a non-

emissive state and will not recover

emission intensity. Emissive events are

called β€œon” times or events and non-

emissive events are called β€œoff” times or

events. Durations of on times and off times are associated with rate constants of dark state

population and BET, respectively. Robust statistical analysis of the probability distributions of

the durations of on and off times can reveal a distribution of rate constants.

Electrons can cycle through these three states (Fig. 1.3A), but this cycling is not

favorable for efficient DSP. Kinetic redundancy resulting from cycling through fluorescent

pathways prevents electrons from injecting into the semiconductor from the dark state (3 in Fig.

1.3A) and contributes to inefficient DSP. When an EY single-moelcule emitter undergoes ISC to

a long-lived triplet state, however, the electron can wait in this dark state until the appropriate

reactions have proceeded in the catalyst. This research aims to measure the blinking dynamics of

Figure 1.3: A) Three level schematic of blinking and B) blinking trace of EY with emissive intensity (black line). Red line indicates statistical analysis of emission intensity.

A

B

Page 12: Exploring the Blinking Dynamics of Eosin Y

6

EY in order to understand the kinetics contributing to favorable ISC, which will in turn improve

DSP efficiency.

Previous studies have demonstrated that heterogeneous excited-state dynamics can be

untangled using single-molecule spectroscopy.23,24,33,25–32 Single-molecule blinking dynamics

have been used to probe the inhomogeneities of ET occuring in dye-sensitized TiO2 systems.27–33

The Wustholz lab has demonstrated the validity of using SMS to measure the blinking dynamics

of rhodamine33 and anthraquinone dyes34 in order to gain a comprehensive understanding of ET

energetics. Prior results from the group have also shown that factors such as dye structure and

photodegredation contribute to the heterogeneous electron-transfer dynamics observed at the

dye-semiconductor interface in DSSC systems.32,35 Studying the kinetics on glass first is an

important control in understanding how the presence of a semiconductor impacts the

photophysics. Using the SMS and statistical analysis methods previously used by the Wustholz

group to understand chromophore blinking dynamics, this research aims to understand the

kinetics responsible for blinking in EY for the development and design of efficient DSP solar

cells.

1.3 Methods

1.3.1 Sample Preparation

Eosin Y (~99%) was used as received from Sigma-Aldrich and diluted with ethanol

(absolute anhydrous, 200 proof) obtained from Pharmco-Aaaper. Base on known pKa values of

EY,36 the predominant species of EY in ethanol is the dianionic form (EY2-). Glass coverslips

(Fisher Scientific, 12-545-102) were cleaned in a base bath for 12-24 h, rinsed in deionized water

(ThermoScientific, EasyPure II, 18.2 MΞ© cm), and dried with clean air (Wilkerson, X06-02-

Page 13: Exploring the Blinking Dynamics of Eosin Y

7

000). All dye solutions were prepared in ethanol using base-bathed glassware. For single-

molecule measurements, samples were prepared by spin-coating 35 Β΅L of a 5Γ—10-10 M EY in

ethanol solution onto a clean glass coverslip using a spin coater (Laurell Technologies, WS-400-

6NPP-LITE) operating at 3000 rpm. The resulting samples were mounted in a custom designed

flow cell and left open to ambient air (i.e., 𝜌!! β‰ˆ 160 torr) for oxic conditions or continuously

flushed with N2 (Airgas, 100%) at a rate of 0.2-0.5 scfh (Key Instruments, MR3A01AVVT)

during anoxic experiments.

1.3.2 Confocal Microscopy

Samples for single-molecule studies were placed on a nanopositioning stage (Physik

Instrumente, LP E-545) on top of a confocal microscope (Nikon, TiU). Laser excitation at 532

nm (Spectra Physics, Excelsior) was focused to a diffraction-limited spot using a high numerical

aperture (NA) 100Γ— oil-immersion objective (Nikon Plan Fluor, NA=1.3). An excitation power

of 0.37 Β΅W at the sample was used for

single-molecule measurements. Emission

from the sample was collected through the

objective, spectrally filtered using an edge

filter (Semrock, LP03-532RS-2S), and

focused to an avalanche photodiode

detector (APD) with a 50-Β΅m aperture

(MPD, PDM050CTB) to provide for

confocal resolution. A z-axis microscope

lock (Applied Science Instruments, MFC-Figure 1.4: Schematic of confocal microscope for SMS studies.

Page 14: Exploring the Blinking Dynamics of Eosin Y

8

2000) was used to maintain the focal plane of the objective during raster scans. A custom

LabView program was used to control the nanopositioning stage and collect corresponding

emission intensity using a 30-ms dwell time. Single-molecule emission was established based on

the observation of diffraction-limited spots, blinking dynamics, irreversible single-step

photobleaching, and concentration dependence of the spot density. The spot density was

approximately 5 molecules per 36 Β΅m2 for 5Γ—10-10 M EY in N2 and air. For dye scans in air

where the exact number of molecules was difficult to quantify (see Fig. 2.2), control experiments

in which N2 flow was turned on after ambient air scans to confirm a single-molecule film were

performed. Blinking was measured on the spots with the highest number of counts (i.e.

intensity), unless they were within ~1 Β΅m of each other, to ensure blinking was not performed on

one molecule more than once.

1.3.3 Blinking Analysis

Blinking dynamics were measured using a 10-ms integration time, defined as emission

intensity per 10-ms (also referred to as bin time or dwell time) for durations β‰₯ 200 s. Blinking

traces were analyzed using the change point detection (CPD) method. CPD can resolve up to 20

distinct intensity levels and corresponding durations. It reports statistically significant changes in

intensity, and it allows for multiple on and off intensity levels.37 An alternative method known as

thresholding has been implemented widely, where visual analysis gives an intensity level above

which are all emissive events and below which are all non-emissive events. However, studies

have shown that thresholding yields false short on-time events and that CPD gives a more

accurate description of emissive and non-emissive events, seen in the red line in Fig. 1.3B.38–41

Page 15: Exploring the Blinking Dynamics of Eosin Y

9

CPD also resolves two types of events called segments and intervals. A segment refers to

a given intensity level and an interval corresponds to successive segments that occur prior to a

switch between an emissive and non-emissive event. An on interval can be made up of several

successive on segments, which occur at different on intensities, whereas an off interval is

comprised of successive off segments. A single molecule has equal numbers of on and off

intervals. Segments are also called β€œtimes” and β€œevents” in this study, consistent with previous

reports.32,33 In CPD, the first and last segments are disregarded since they are arbitrarily set by

the experimental observation time. The lowest deconvolved intensity is designated as a non-

emissive or photobleaching event, depending on if emission is recovered. Intensities greater than

one standard deviation above the rms noise (for 10-ms integration time, ~4 counts) are

designated as emissive, or on, events. In this study blinking statistics are used as a method of

quantifying and comparing dynamics in air and in N2. An event rate is defined as the total

number of events per time of trace.

On and off times are compiled into a probability distribution and fit by test functions.

On-time distributions reveal information about electron injection and off-time distributions do

the same for charge recombination events. Much like first order kinetics are fit by an exponential

function, more complicated kinetic models are fit by a given function. For example, dispersive

ET kinetics have been shown to follow heavy-tailed distributions, such as power law, lognormal

and Weibull functions.32,33,42 While the least squares (LS) fitting method has been used widely to

model power-law and other heavy-tailed functions, it has been proven to be problematic with

power laws.43,44 Studies demonstrate that maximum likelihood estimation (MLE) gives a more

accurate estimation of the fitting parameters.32–34,42,43

Using this method, the power law, 𝑃 𝑑 = 𝐴𝑑!!, is normalized with an onset time for

Page 16: Exploring the Blinking Dynamics of Eosin Y

10

power-law behavior (𝑑!"#), giving:

𝑃 𝑑 = !!!!!"#

!!!"#

!!,𝛼 > 1 (1)

MLE finds the most likely 𝛼 and 𝑑!"# , where 𝛼 is another fitting parameter of power law.

However, MLE assumes that a power-law function will fit the dataset and does not indicate

whether the fit is accurate.43 Clauset and coworkers determined that MLE only gives accurate

fitting parameters when 𝑑!"# is found using the Kolmogorov-Smirnov (KS) statistic. Using the

statistically robust method developed by Clauset and coworkers, the probability distribution

functions (PDFs) were analyzed using a combination of the MLE and KS statistic methods.

The KS test involves transforming the dataset into a cumulative distribution function

(CDF) that gives the probability of an event occurring in a time less than or equal to 𝑑:

𝑆 𝑑 = 𝑃 𝑑! 𝑑𝑑! = !!

!!!"#

𝑑! ≀ 𝑑! (2)

in which 𝑆 𝑑 is the CDF and 𝑃 𝑑! is the PDF. In eq. 2, the CDF is determined from the

blinking data. An artificial CDF is created which fits the fitting parameters 𝛼 and 𝑑!"# exactly.

The KS statistic is the maximum distance between the real (𝑆(𝑑)!"#") and artificial (𝑆(𝑑)!"#)

datasets, given by:

𝐷 = max!!!!"# 𝑆(𝑑)!"#" βˆ’ 𝑆(𝑑)!"# (3)

where a perfect fit yields 𝐷 = 0. The best fit, therefore, is described by the 𝛼 and 𝑑!"# that

minimize D.

In order to find out the plausibility of the fit to the data, the data was analyzed using a

goodness-of-fit test developed by Clauset and coworkers.43 The best-fit parameters 𝛼 and 𝑑!"#

are used to simulate synthetic data sets, which are fit in the same way as the empirical data, and

the KS statistic for the synthetic data (𝐷!!"#!) is determined. The goodness-of-fit is quantified

Page 17: Exploring the Blinking Dynamics of Eosin Y

11

using the 𝑝 value, which represents the fraction of synthetic data sets with KS statistics larger

than that of the experimental data:

𝑝 = !!!!"#$!!!

(4)

where 𝑁! is the number of synthetic data sets (typically set to 104). 𝐷!"#$! here represents

deviations from statistical fluctuations, allowing us to determine if the fitting parameters actually

fit the experimental data, and are not artificial. Even if a power-law fit yields a high 𝑝 value, it

does not guarantee that the data are fit by a power-law function. Therefore, the MLE/KS method

developed by Riley et al. was used to fit blinking dynamics to other heavy-tailed functions

(Weibull and lognormal) in addition to power-law.42 Previous single-molecule studies have used

the MLE/KS analysis to determine lognormal distributions for rhodamine dyes on TiO2,

consistent with the Albery model for dispersive ET, in which activation barriers are normally

distributed.33,34,42,43

In this research, on- and off-time distributions were also fit to a lognormal function with

an onset time.Since analytical expressions for these fit parameters cannot be determined using

MLE, a numerical method was used to refine these values. Here, an initial guess for the mean of

the distribution is used to seed the numerical method and the best fit is determined by the set of

parameters that minimizes the KS statistic. Standard errors in the fit parameters are determined

from the second derivative of the log-likelihood with respect to the parameters. All data analyses

and fitting procedures were completed in Matlab (version R2018a).

1.4 Outline

In this thesis, SMS studies of EY photosensitizers under varying environmental

conditions are presented. An important first step in understanding the complex photophysics is

Page 18: Exploring the Blinking Dynamics of Eosin Y

12

studying blinking on glass as a control, before proceeding to more technologically relevant

substrates, such as TiO2. Therefore, in Chapter 2, the blinking dynamics of EY on glass substrate

are explored. First, the blinking dynamics in the absence of oxygen are examined, followed by

the more complex dynamics in the presence of oxygen. Using robust KS/MLE analysis, it is

shown that the complex excited-state dynamics can be modeled by multiple functions to reflect

competing photophysical pathways. As a result, the relative contributions of ISC and dispersive

electron transfer in EY blinking dynamics were unraveled. By understanding how competing

events contribute to the overall distribution of on- and off-times, a more complete understanding

of the underlying processes contributing to DSP inefficiencies is gained.

Chapter 3 explores several secondary studies that motivate future research and pose

interesting questions regarding the complex blinking dynamics of EY. Changes in laser

excitation power alter the on-time distribution of EY in the presence of oxygen, revealing a

power dependency of the blinking dynamics. Finally, future studies of the blinking dynamics of

EY on TiO2 are discussed. Preliminary dye film images and blinking traces indicate that the

excited-state dynamics of EY are significantly altered in the presence of a semiconductor,

encouraging more study of EY under technologically relevant conditions.

Page 19: Exploring the Blinking Dynamics of Eosin Y

13

Chapter 2: Unraveling the Excited-State Dynamics of EY Photosensitizers Using Single-Molecule Spectroscopy

2.1 Introduction

As mentioned in Chapter 1, EY shows great promise as a photosensitizer for DSP

systems because it can undergo ISC. It was hypothesized that these heavy atoms would make the

ISC pathway in EY more likely to occur, and the goal of this research was to probe the excited-

state dynamics of EY in order to gain a better understanding of the ET dynamics. Initially, the

aim was to analyze the blinking dynamics of EY on glass and TiO2 in N2, consistent with

previous studies.32–34 However, the project changed course significantly. During control

experiments and the process of finalizing experimental setup parameters, raster scans and

blinking traces were performed with and without N2 flow. Significant visual differences were

observed in the spot density and pixilation of dye-film images, as well as in the number and

duration of on and off events during blinking traces. The presence of oxygen is known to play a

role in the excited state dynamics of EY through quenching of the triplet state to produce singlet

oxygen,20,45–47 and preliminary observations indicated a stark modification of EY blinking

dynamics in ambient air relative to N2. It became clear that the observations in the blinking

traces resulted from this phenomena, and thus the scope of the study was expanded to include the

impact of oxygen on the photophysics of EY. In this chapter, the blinking dynamics of EY

photosensitizers in the absence and presence of oxygen are quantified and analyzed using robust

MLE/KS analysis in order to gain a more complete understanding of how triplet state quenching

impacts the excited-state dynamics and resulting on- and off-time distributions.

Page 20: Exploring the Blinking Dynamics of Eosin Y

14

2.2 Blinking Studies in Anoxic Environments

Fig. 2.1 shows a

representative false-colored

image of 5Γ—10!!" M EY

spun coat onto a glass slide

(A) and blinking trace (B)

in an anoxic environment.

CPD analysis (red line in

Fig. 2.1B) of the blinking

trace gives 14 intensity levels, 24 on events, and 2 off events, with an average event rate of 0.15

s-1. A photobleaching event is seen at 85.9 s. Blinking dynamics of EY with multiple emissive

intensities occurring before photobleaching is consistent with previous studies of xanthene and

anthraquinone dyes on glass.33,34

In order to examine the full range of photophysical behaviors of EY on glass in N2, the

blinking dynamics of 127 molecules were measured and analyzed. Fig. 2.2 presents the on-time

distribution that contains 1872 events, with an average on time of 3.5 s and values ranging from

0.02 s to 151.66 s. The off-time distribution contains 513 events, with an average off time of

17.15 s and values ranging from 0.04 s to 195.62 s. The average off time is significantly longer

than literature values for the triplet lifetime of EY (e.g. 55 Β΅s in water,48 3.6 ms in poly(methyl

methacrylate)47, and 1 ms on alumina49), which suggests that triplet-state decay is not the

primary mechanism responsible for blinking dynamics. The average event rate is 0.09Β± 0.01 s-1

molecule-1. Of the 127 molecules examined, 87 (69%) underwent single-step photobleaching

within the observation period (200 s). To understand the nature of the non-emissive state, the on-

Figure 2.1: (A) 6x6 Β΅m2 scan of 5x10-5 M EY on glass substrate in N2. Color scale represents counts per 30-ms. (B) 100-s excerpt of a blinking trace of a single EY molecule. Black line indicates fluorescence intensity. Red line indicates trajectory of CPD analysis.

Page 21: Exploring the Blinking Dynamics of Eosin Y

15

and off-time probability distributions were fit to

several functions that allow us to glean kinetic

information. For example, if blinking is the result of

the population and depopulation of the triplet state, it

will be consistent with first-order kinetics and the on-

and off-time probability distributions will be

exponentially distributed. Visual analysis of the

distributions in Fig. 2.2 tell us that these are not

described by exponential functions, since they are not

a straight line on the log-log axes, prompting us to

explore more complex photophysical kinetic models

and corresponding test functions. The compiled

distributions were fit to power-law, lognormal, and

Weibull functions using the MLE/KS method,

described in 1.3.3.

The resulting fit parameters of the on- and off-time probability distributions for EY in N2

are shown in Table 2.1. According to p-values alone, the on-time distribution is best fit by a

power-law function with 𝛼 = 2.46Β± 0.03 and p = 0.02. Although the positive p-value supports

the power-law fit, the given 𝑑!"# = 7.55 𝑠 means that the fit is only operative for 4% of the data,

so it is not an accurate description of the entire distribution. Our efforts to fit the on-time data to

Weibull and lognormal functions gave p-values equal to zero. Ultimately, analysis of on-time

distributions with the MLE/KS method demonstrates that power-law is operative after an onset

time of ~101 s, but does not fit the majority of the on-time data.

Figure 2.2: The (A) on- and (B) off-time probability distributions of events in air (blue) and in N2 (red) and their best fit parameters.

Page 22: Exploring the Blinking Dynamics of Eosin Y

16

The off-time distribution is also best fit to a power law with 𝛼 = 6.7Β± 0.3, 𝑑!"# =

107.43 s, and p = 0.74. Similar to on-time data, though, the power law does not describe the

majority (98%) of the data. Instead, a lognormal distribution is most representative of the off-

time data, as demonstrated by a p-value of 0.03. The lognormal distribution occurs when the

logarithm of the sampled variable (𝑑) is normally distributed according to:

𝑃 𝑑 = !!!!"

𝑒!!" ! !! !

!!! (5)

where the scale parameters πœ‡ and 𝜎 correspond to the geometric mean and standard deviation of

the variable’s natural logarithm, respectively. The off-time probability distribution for EY in N2

is best fit to a lognormal function corresponding to πœ‡ = 1.64Β± 0.08 and 𝜎 = 1.72Β± 0.05 (Fig.

2.2).

Overall, MLE/KS analysis of blinking dynamics of EY on glass in N2 shows that on

times are power-law distributed after an onset time and off times are weakly lognormally

distributed. These results are consistent with previous studies, which attribute power-law and

lognormal behavior to dispersive electron transfer between dyes and trap states on glass.32,33,50,51

In this case, however, statistically insignificant p-values indicate that dispersive electron transfer

is not the only mechanism responsible for the blinking dynamics observed. Previous ensemble-

Power Law: !!!

!!"#( !!!"#

)!!

Lognormal: !!"√!!

𝑒!(!"(!)!!)!

!!! Weibull: !!! !!!!!!

𝑒!!!!!!

𝑑!"#(𝑠) 𝛼 𝑝 πœ‡ 𝜎 𝑝 𝐴 𝐡 𝑝

ON N2 7.55 2.46 Β± 0.03 0.02 -0.01 Β± 0.04 1.66 Β± 0.03 0 0.640 Β± 0.005 2.31 Β± 0.07 0

Air 0.83 2.38 Β± 0.02 0.11 -1.17 Β± 0.01 1.05 Β± 0.01 0 0.820 Β± 0.002 0.55 Β± 0.01 0

OFF N2 107.43 6.7 Β± 0.3 0.74 1.64 Β± 0.08 1.72 Β± 0.05 0.03 0.617 Β± 0.008 12.1 Β± 0.8 0

Air 4.73 2.83 Β± 0.04 0.20 0.10 Β± 0.03 1.07 Β± 0.02 0.007 0.90 Β± 0.01 1.9 Β± 0.2 0

Table 2.1. Best-fit parameters and p-values for power-law, Weibull, and lognormal distributions. Errors represent one standard deviation. Physically significant fitting results are highlighted in gray.

Page 23: Exploring the Blinking Dynamics of Eosin Y

17

averaged research shows that upon photoexcitation, EY undergoes rapid ISC (i.e., π‘˜!"#~10! s-1)

to its lowest energy triplet state, which can be long-lived on solid substrates.20,47,49,52 Triplet

blinking could be responsible for blinking events at early times, even if the probability of a 1-s

off event occurring for a 1-ms triplet lifetime is minute (~10-44). If this were the case, the off-

time distribution would contain contributions from two competing processes: triplet-state

population and dispersive electron transfer. The presence of oxygen is known to significantly

impact the excited-state dynamics of EY through quenching of the triplet state to produce singlet

oxygen.20,45–47 Therefore, comparing the blinking dynamics of EY in N2 with that in air will give

insight into the hypothesis that ISC to and from the triplet state effects the photophysics of EY.

2.3 Impact of Oxygen on Blinking Dynamics

Fig. 2.3 shows a

representative false-colored

image of 5Γ—10!!" M EY

spun coat onto a glass slide

(A0 and blinking trace (B) in

an oxic environment.

Although there are a few

round spots, typically

indicative of a single molecule, image is pixelated in comparison to that in N2 (Fig. 2.1),

demonstrating that non-emissive events longer than the 30-ms bin time are observed frequently.

The blinking statistics of EY in air are also distinctive. CPD analysis of the blinking trace in Fig.

2.3B gives 11 intensity levels, 113 on events and 36 off events, with an average event rate of

Figure 2.3: (A) 6x6 Β΅m2 scan of 5x10-5 M EY on glass substrate in N2. Color scale represents counts per 30-ms. (B) 100-s excerpt of a blinking trace of a single EY molecule. Black line indicates fluorescence intensity. Red line indicates CPD analysis.

Page 24: Exploring the Blinking Dynamics of Eosin Y

18

0.75 s-1. While 69% of molecules in N2

underwent photobleaching within 200 s, none

in air demonstrated photobleaching within this

time frame. On the contrary, EY exhibited

persistent blinking dynamics for as long as

1000 s (Fig. 2.4). Our observations are

consistent with previous studies that show

increased photostability of EY in oxic

environments.46

Because of these long observation windows and high event rates, the probability

distributions of 17 molecules contained 6681 on events and 1810 off events. Emissive event

durations ranged from 0.02 s to 193.07 s with an average of 0.66 s. Non-emissive event durations

ranged from 0.06 s to 36.2 s with an average of 2.0 s. The average on and off times are an order

of magnitude shorter in air than in N2, and the average blinking rate of 0.90Β± 0.07 s-1molecule-1

is also an order of magnitude higher than that of N2.Like the distributions in N2, the air on- and

off-time probability distributions were fit using MLE/KS analysis, and fitting parameters are

summarized in Table 2.1. The distributions and best-fit parameters are given in Fig. 2.2. The on-

time distribution is fit best by a power law with a p-value of 0.11, corresponding to 𝛼 = 2.38Β±

0.02 and 𝑑!"# = 0.83 s, which describes 31% of the dataset. Fits to Weibull and lognormal

functions gave p-values of zero. The off-time distribution was best fit by power law,

corresponding to p=0.20, 𝛼 = 2.83Β± 0.04, 𝑑!"# = 4.73 s. However, the late onset time means

that power law only fits a small subset (10%) of the off-time data. Unlike the on times, a non-

Figure 2.4: Blinking dynamics of a single EY molecule in air, obtained using 532-nm continuous excitation. CPD analysis revealed that this molecule exhibits 14 statistically-significant intensities, 828 on events, 220 off events. No photobleaching event is detected within the 1000-s observation window.

Page 25: Exploring the Blinking Dynamics of Eosin Y

19

zero p-value of 0.007 suggests that a lognormal distribution with πœ‡ = 0.10Β± 0.03 and

𝜎 = 1.07± 0.02 is a plausible model for the off-time data.

Substantial differences in the single-molecule blinking dynamics are seen for EY in oxic

and anoxic environments. The increase of an order of magnitude in the average single-molecule

event rate and short event durations in air lead to significantly more events occurring in the

presence of oxygen. Additionally, most of the molecules in anoxic environments photobleached

within 200 s, whereas persistent blinking in EY in air was observed for durations up to 1000 s.

Both on-time distributions are modeled by power law functions after an onset time, and 𝛼 is

modified in N2 (𝛼 = 2.46Β± 0.03) versus air (𝛼 = 2.38Β± 0.02). The best-fit parameters of the

off-time probability distributions are significantly different for N2 (i.e., πœ‡ = 1.64Β± 0.08

and 𝜎 = 1.72Β± 0.05) as compared to air (i.e., πœ‡ = 0.10Β± 0.03 and 𝜎 = 1.07Β± 0.02). Overall,

the presence of oxygen has a significant impact on the blinking dynamics of EY, which is

consistent with a physical mechanism that involves the population and depopulation of the triplet

state.

2.4 Discussion

2.4.1 Albery Model for Power Law and Lognormal Distributions

Previous research has shown that power-law and lognormal distributions are consistent

with dispersive electron transfer between an excited dye and substrate (Fig. 2.5A).33 In this

expression, π‘˜!"#, π‘˜!", π‘˜!"#, and π‘˜!"# are rate constants for excitation, emission, and population

(injection) and depopulation (recombination) of the dark state, respectively. Injection and

recombination rate constants support the Albery model of electron transfer, which gives a

Gaussian distribution of activation barriers to electron transfer.53 In this model, the duration of an

Page 26: Exploring the Blinking Dynamics of Eosin Y

20

off event (𝜏!"") is equivalent to the lifetime of

the dark state (i.e., 𝜏!"# = 1/π‘˜!"#). The Albery

model predicts a lognormal distribution of off

events. The lifetime of the emissive state (𝜏!",!)

is more convoluted, as it is dependent on the

rate constant for injection (π‘˜!"#) , which is

partially responsible for depopulating the

excited state, as well as the population and

depopulation resulting from excitation (π‘˜!"#)

and emission (π‘˜!") . The expression for the

emissive state lifetime is:

𝜏!",! = π‘˜!"#!!"#

!!"#!!!"

!! (6)

Previous studies have shown that the on-time

probability distribution can be power-law or

lognormally distributed.32–34

The observation of power-law and

lognormal distributions for EY does support the

role of dispersive electron transfer for blinking dynamics, but this explanation fails to account for

the oxygen-dependent photophysics (i.e. triplet blinking) as well as the statistically insignificant

p-values (<0.05) corresponding to the lognormal fits of the off-time distributions. This suggests

that the Albery model does not give a complete picture of the blinking kinetics of EY, and that

the lognormal distribution is an inadequate description of the data. As a result, alternative kinetic

Figure 2.5: Energy level diagrams for blinking consisting of a singlet ground state (S0), an excited singlet state (S1), an excited triplet state (T1), and a dark state (D), with corresponding rate constants. (A) dispersive electron transfer; triplet blinking and dispersive electron transfer in the (B) presence and (C) absence of oxygen. Dotted arrows indicate dispersive ET and solid arrows indicate a single rate constant.

Page 27: Exploring the Blinking Dynamics of Eosin Y

21

models and expressions were explored to incorporate triplet blinking and dispersive electron

transfer to describe the data.

2.4.2 Kinetic Expressions for Triplet Blinking and Dispersive Electron Transfer

Single-molecules studies in polymer50 and crystal54 environments have utilized a four

electronic state system with a triplet state and a dark state to model power-law behavior. For EY,

both power law and lognormal behavior is observed, although p-values are modest. In Fig. 2.5B,

a four level system is proposed to take into account the impact of singlet oxygen formation on

the blinking dynamics of EY in oxic conditions. Here, photoexcitation of EY to its excited

singlet state (S1) is followed by rapid ISC to a triplet state (T1). The triplet is quenched in the

presence of triplet oxygen (3O2) to form singlet oxygen (1O2), with rate constant π‘˜! . If

depopulation of T1 via quenching is fast relative to population of T1 via ISC, then a small steady-

state concentration of T1 is present. Thus, the four-level system effectively reduces to a three-

level system (Fig. 2.5A). In this case, population of the dark state via dispersive electron transfer

will likely occur from S1 (i.e. the oxidation potential of EY in the excited singlet state, E0

(EY*/EY+), is -1.53 V vs. NHE)15,55 to localized trap states on glass.51

In this expression, 𝜏!"" is equal to the lifetime of the dark state (𝜏!"#) and we predict that

the off-time distribution will be lognormally distributed. The on times for Fig. 2.5B (𝜏!",!) are

dependent on the rate constant for population of the dark-state as well as the fractional

populations of the excited singlet and triplet states. The on-time probability distribution is a

convolution of the kinetics for emission as well as dark-state population and depopulation. On

time durations follow the expression:

𝜏!",! = π‘˜!"#!!"#

!!"#!!!"

!!"#!!!"#!!!"#

!! (7)

Page 28: Exploring the Blinking Dynamics of Eosin Y

22

The fast quenching of T1 by singlet oxygen formation effectively reduces the four level system to

a three-state model with dispersive electron-transfer for blinking, in which β€œon” corresponds to

𝑆! ⇄ 𝑆! and β€œoff” is the non-emissive radical cation state of EY.15

On the other hand, if there is no oxygen present, or π‘˜! is not significantly larger than π‘˜!"#,

then a more complex kinetic picture is at play (Fig. 2.5C). In this case, both triplet blinking and

dispersive electron transfer contribute to the kinetics, so the measured on times come from 𝜏!"#

as well as the lifetime of triplet state (i.e., 𝜏! = 1/π‘˜β€²!"#). The off-time distribution is then

predicted to be a combination of an exponential distribution (first order kinetics for triplet

blinking) and a lognormal distribution (dispersive electron transfer from S1 and/or T1).15 For Fig.

2.5C, the on times (𝜏!",!) depend on intersystem crossing to the triplet state (π‘˜!"#) and the

fractional population of the excited state according to:

𝜏!",! = π‘˜!"#!!"#

!!"#!!!"

!! (8)

The on-time probability distributions are very complex, and are not well represented by

any test function, consistent with the proposed models in Fig. 2.5B and 2.5C. The presence of a

small p-value for a lognormal distribution of off times for EY in N2 implies that the kinetics

modeled by Fig. 2.5C may be present. The statistically insignificant p-value could indicate more

complex behavior and an onset time for lognormal behavior. The current MLE/KS method does

not allow us to distinguish between competing photophysical processes in one probability

distribution. However, we hypothesize that the off-time distributions could be fit with an

exponential distribution, to account for triplet blinking at early times, and a lognormal

distribution, to account for dispersive electron transfer at later times. To test this theory, fitting

methods that allow for differentiation between competing photophysical processes were

explored.

Page 29: Exploring the Blinking Dynamics of Eosin Y

23

2.4.3 Modeling Competitive Photophysical Processes

Recently, a β€œscanning” MLE/KS method was developed by Mitsui and coworkers, in

which the onset time 𝑑!"# of lognormal behavior is progressively changed to determine

whether there are contributions from other functions in the data.56 They calculated the p-value as

a function of 𝑑!"# and showed that the off-time distribution of Atto 647N dyes on TiO2 is

described by a combination of exponential, power law, and lognormal functions. Similarly,

Clauset and coworkers developed a method that uses MLE/KS to determine the best 𝑑!"# such

that the data is fit by a lognormal function with an onset time. This method normalizes the

lognormal distribution from 𝑑!"# (instead of zero as is done for eq. 5) to yield:

𝑃 𝑑 = !!!!"

π‘’π‘Ÿπ‘“π‘ !" !!"#!!!!

!!𝑒!

!" ! !! !

!!! (9)

The data were analyzed using both methods in order to determine if more than one

function could be fit to the off-time probability distributions. The scanning MLE/KS method did

not yield a 𝑑!"# value that was statistically significant and physically relevant. As a result, we

used the approach where the best 𝑑!"# is

determined using eq. 9. Although all of

the on- and off-time distributions were

analyzed using this method, only the off

times in N2 gave a 𝑑!"# greater than the

first data point. The off time distribution

of EY in N2 is shown in Fig. 2.6, with

the best fit lognormal function

corresponding to 𝑑!"# = 0.14 s,

Figure 2.6: Off-time probability distribution (solid black line) of EY in N2 with best fit to a lognormal function after an onset time of 0.14 s (red dashed line) and an exponential function at early times (blue dotted line).

Page 30: Exploring the Blinking Dynamics of Eosin Y

24

πœ‡ = 1.63Β± 0.06, 𝜎 = 1.8Β± 0.1, and 𝑝 = 0.07. The statistically significant p-value suggests that

the lognormal distribution could be operative for the 98% of the data that occurs after the onset

time, 𝑑!"# = 0.14 s. The observation that a lognormal distribution does not fit off times

occurring before 0.14 s indicates that there is another photophysical process at play, such as

triplet state blinking. The first part of the distribution was fit to an exponential fit, consistent with

triplet state population and depopulation.

Because we observe that the off-time probability distribution for EY in N2 is best fit to a

lognormal distribution following an onset time, it suggests that the kinetic model shown in Fig.

2.5C is operative. Both dispersive electron transfer and triplet blinking occur in anoxic

environments, so the off-time durations (𝜏!"" ) result from both processes. The off-time

probability distribution contains an exponential distribution at early times, accounting for the

contribution of triplet state decay, and a lognormal distribution at later times, accounting for

electron transfer. It was considered that a power law could describe the tail of the distribution

well, as seen by the power-law fitting parameters from Table 2.1, but the third function was not

included due to two reasons. First, since a statistically significant p-value is seen for the

lognormal function after 𝑑!"# , the tail of the distribution is already fit by a test function.

Secondly, the physical meaning of a third function is unclear and could not be connected to the

proposed kinetic models in Fig. 2.5. As a result, using three functions to fit the distribution was

not justifiable.

The best fit of early times (i.e. before the onset time of the lognormal function, from 0.04

to 0.14 s) to an exponential decay (𝑃 𝑑 = !!𝑒!!/!)isshowninFig.2.6,andcorresponds to 𝜏 =

93 ms. Although the time constant is much longer than the reported triplet lifetime of EY on

alumina (1 ms),49 we suggest that triplet-state decay is occurring at timescales faster than the 10-

Page 31: Exploring the Blinking Dynamics of Eosin Y

25

ms bin time of our experiment. In order to test this hypothesis, blinking measurements of EY

with a 1-ms integration time were performed.

A recent study by Mitsui and coworkers progressively shortened the bin time for

measuring the single-molecule blinking dynamics of perylenediimide (DMP-PDI) in a matrix of

poly(methyl methacrylate) (PMMA).57 They observed bin time dependence in the resulting

probability distributions as they changed the bin time over five orders of magnitude (0.08-100

ms). By capturing shorter time events, they are able to see events that occur on different

timescales, and in their analysis they attribute early times to ISC and later times to charge

transfer. While the probability of seeing an event corresponding to triplet state decay is very

Figure 2.7: (A) Blinking trace of EY in air at 1-ms bin time and (B) corresponding trace binned up to 10-ms. (C) 1-ms bin time blinking trace of EY in N2 and (D) corresponding binned up trace.

A

C

B

D

Page 32: Exploring the Blinking Dynamics of Eosin Y

26

small for both 10- and 1-ms bin times, in theory the contribution of shorter time events should

become more apparent as the bin time is lowered.

Blinking dynamics of EY in both N2 and in air (Fig. 2.7) were using a 1-ms bin time. In

order to determine that the data was valid, the 1-ms blinking traces were β€œbinned up” to 10-ms.

Here, counts from 10-ms windows are summed, and this value is a data point on a 10-ms bin

time blinking trace. Binning up the data allows for comparison between the experimental 10-ms

data and the binned up 1-ms data. They should exhibit the same blinking statistics and their

probability distributions should be fit by the same test function. After performing CPD analysis

on the binned up traces, it was determined that the resulting traces did not reflect the dynamics

seen with an experimental 10-ms bin time. Changing the bin time is known to introduce artificial

events and can alter the resulting probability distributions and fits.58,59 The resulting signal-to-

noise was too low to prevent the introduction of binning artifacts. As a result, the 1-ms bin time

experiments were not included in the analysis of the EY blinking dynamics.

Analysis of the on- and off-time probability distributions of EY blinking dynamics using

the MLE/KS method unveil the contributions of the competing photophysical processes of triplet

state blinking and dispersive electron transfer. In anoxic environments, both processes contribute

to the blinking dynamics. However, in the presence of oxygen, the triplet state (T1) is rapidly

quenched by the formation of singlet oxygen species such that electron transfer is the primary

contributor to blinking dynamics. The relationship between the lognormal fitting parameters and

dispersive electron-transfer kinetics has been demonstrated in previous studies, where βˆ’πœ‡!"" is

proportional to the average rate constant for dark state depopulation (βˆ’πœ‡!"" = ln π‘˜!"" ) and

𝜎!"" corresponds to the energetic dispersion around the mean activation barrier.32 In this study,

we find that βˆ’πœ‡!"" of EY in air is βˆ’0.10Β± 0.03, similar to values for rhodamine dyes not

Page 33: Exploring the Blinking Dynamics of Eosin Y

27

containing heavy atoms.33 This supports the hypothesis that EY triplet blinking does not occur in

the presence of oxygen. The observation of persistent blinking dynamics of EY in air is

consistent with previous work and suggests that the short lifetime of singlet oxygen limits the

oxidative degradation of EY.46 It is also possible that fluctuations in the triplet state lifetime and

fluorescence quantum yield result from spectral diffusion, like during photoinduced

debromination events.46 These observations motivate more research in order to understand how

spectral diffusion impacts the excited-state kinetics of EY.

Observed 𝜎 values are very different for EY in air (𝜎!"" = 1.07± 0.02) versus N2

(𝜎!"" = 1.8± 0.1), which suggests that there are different photophysical processes at play. It is

likely that both static and dynamic heterogeneity of electron transfer is operative, which is

consistent with our observation of dispersive kinetics in an individual molecule of EY in air.

Static heterogeneity refers to individual molecules with homogeneous kinetics that contribute to

a heterogeneous environment, whereas dynamic heterogeneity involves individual molecules

with kinetics that evolve over time. In addition to analyzing the probability distributions of the

on and off events for all molecules measured in air, the probability distribution of an individual

molecule of EY in air was analyzed using the MLE/KS method. During a 500 s observation

window, CPD analysis gave 338 emissive and 99 non-emissive events, which were plotted in on-

and off-time probability distributions, respectively (Fig. 2.8). The on-time distribution is best fit

to a power law corresponding to 𝛼 = 1.94Β± 0.06, 𝑑!"# = 0.20 s, and p = 0.27. The off-time

distribution is best fit by a lognormal function with fitting parameters πœ‡ = βˆ’0.17Β± 0.09,

𝜎 = 0.92Β± 0.07, and 𝑝 = 0.67. All other test functions yielded statistically insignificant p-

values. A lognormal fit of off times for an individual molecule is consistent with dispersive

kinetics that incorporate dynamic heterogeneity; that is, the rate constants evolve over time.60–63

Page 34: Exploring the Blinking Dynamics of Eosin Y

28

Dispersive kinetics in an individual molecule

have been reported for multichromophoric

systems,62 molecules in crystals63 and in

polymer,57,64,65 but our observation is unique, as

it is one molecule on glass in the presence of

oxygen. Although consistent with dispersive

electron transfer, the fitting parameters

calculated in this analysis are significantly

modified from the multi-molecule distributions

analyzed in 2.3. The ability to analyze an

individual molecule’s blinking dynamics using

the MLE/KS method opens the door to further

untangle the single-molecule photophysics of

EY as a photosensitizer, and to investigate more

deeply the relationship between kinetics and

environment.

2.5 Conclusion

In this study, single-molecule spectroscopy was used to unveil the excited-state dynamics

of EY. We utilized the statistically robust MLE/KS method to quantify the blinking dynamics in

oxic and anoxic environments, and revealed the relative contributions of triplet-state decay and

dispersive electron transfer. The on-time probability distributions in N2 and air are power-law

distributed after about 1 s, and the fitting parameters are altered significantly in the presence of

Figure 2.8: On- (top) and off-time (bottom) probability distributions of an individual molecule of EY in air. Blue dots indicate events from CPD analysis and red lines indicate best fit functions for power law (top) and lognormal (bottom) distributions.

Page 35: Exploring the Blinking Dynamics of Eosin Y

29

oxygen. The off-time probability distribution for EY in N2 contains contributions from

exponential and lognormal functions. Based on this, a kinetic model that accounts for competing

triplet state population and dispersive electron transfer was proposed. The off-time distribution in

air is best fit to a lognormal function, consistent with dispersive electron transfer. The results in

this study reveal the underlying photophysical kinetics for EY sensitizers on glass substrate,

which is an important first step in the development and design of more efficient DSP systems, in

addition to its relevance in photodynamic therapy45,66 and cultural heritage applications.46,67,68

Page 36: Exploring the Blinking Dynamics of Eosin Y

30

Chapter 3: Secondary Experiments

3.1 Power Dependent Blinking Dynamics

Power-dependent blinking analysis was also performed, in which the power of laser

excitation was varied from 3.7 Β΅W to 21.2 Β΅W. The excitation power was varied with the

intention of photobleaching a molecule of EY in air, since none had exhibited single-step

photobleaching within the observation period at 0.37 Β΅W. However, a change in the blinking

dynamics of EY in air as the laser power increased from 3.7 Β΅W to 21.2 Β΅W was observed.

Visual inspection of the on-time distributions (Fig. 3.1) indicates that there is a power-

dependency in the blinking dynamics. All blinking traces were performed on distinct individual

molecules and had an observation period of 300 s, but exhibited a variation in the number of on

and off events. Indeed, the event rate changed drastically with laser power. Blinking traces were

taken at 3.7, 6.7, 9.9, 14.8, and 21.2 Β΅W, with event rates of 1.93 s-1, 2.35 s-1, 1.23 s-1, 0.85 s-1,

and 0.43 s-1, respectively. Additionally, the observation of a significant change in fitting

parameters to their on-time probability distributions (Table 3.1) indicates that the laser power

significantly alters excited-state behavior. The fitting parameters also differ from that of the

Power Law: !!!!!"!

( !!!"#

)!! Lognormal: !!" !!

𝑒!(!" ! !!)!

!!!

Weibull: !!

!!

!!!𝑒!

!!

!

𝑑!"#(𝑠) 𝛼 𝑝 % Data Fit πœ‡ 𝜎 𝑝 𝐴 𝐡 𝑝

0.37 Β΅W 0.20 1.94 Β± 0.06 0.27 82 -0.97 Β± 0.07 1.09 Β± 0.05 0 0.80 Β± 0.02 0.7 Β± 0.1 0

3.7 Β΅W 0.13 2.20 Β± 0.05 0.148 70 -2.31Β± 0.04 1.03 Β± 0.03 0 0.88 Β±0.013 0.18 Β± 0.05 0

6.7 Β΅W 0.27 2.63 Β± 0.09 0.398 12 -2.65 Β± 0.05 0.95 Β± 0.04 0 0.95 Β± 0.02 0.122 Β± 0.009 0

9.9 Β΅W 0.41 2.58 Β± 0.09 0.588 18 -1.95 Β± 0.06 1.06 Β±0.04 0.004 0.88 Β± 0.02 0.26 Β± 0.08 0

14.8 Β΅W 0.25 2.27 Β± 0.09 0.405 33 -1.74 Β± 0.07 1.04 Β± 0.05 0.16 0.87 Β± 0.02 0.3 Β± 0.1 0

21.2 Β΅W 0.2 2.29 Β± 0.1 0.837 38 -1.9 Β± 0.1 1.08 Β± 0.08 0.315 0.90 Β± 0.04 0.26 Β± 0.05 0.001

Table 3.1: Fitting parameters for on-time probability distributions of individual EY molecules with varying excitation power.

Page 37: Exploring the Blinking Dynamics of Eosin Y

31

individual molecule in air (included in

Table 3.1). Off-time probability

distributions were not analyzed due to

the low number of off times (for 3.7

Β΅W, 64 off segments, 78 for 6.7 Β΅W, 67

for 9.9 Β΅W, 43 for 14.8 Β΅W, 28 for 21.1

Β΅W).

By increasing the excitation

power, the rate constant ( π‘˜!"#)

increases as well.. As a result, the

lifetime of the singlet excited state (𝜏!",!) should decrease based on the kinetic expressions

proposed in 2.4.1 and in eq. 6. In the distributions in Fig. 3.1, the on-time distributions at higher

powers are skewed to shorter durations than the on-time distribution at 0.37 Β΅W (black),

supporting this hypothesis. For excitation powers less than 6.7 Β΅W, power law appears to be

operative for the majority of the data. Although power law yields a statistically significant p-

value for 6.7 Β΅W, it only fits 12% of the data. Other test functions yielded p-values equal to zero

at for 6.7 Β΅W. At higher powers, a lognormal fit yields non-zero p-values. Here, βˆ’πœ‡!" is

proportional to the average rate constant for on-state population (βˆ’πœ‡!" ∝ ln π‘˜!"), so one would

expect the distributions to shift towards shorter on events as βˆ’πœ‡!" decreases. This trend is not

readily apparent, but decisive conclusions cannot be drawn, because the observations are for an

individual molecule at each power. Further investigation, however, could give more clues to the

roles of static and dynamic heterogeneity at play in the excited-state dynamics of EY.

Figure 3.1: On-time distributsions of individual molecules of EY with varying excitation power: 0.37 Β΅W (black), 3.7 Β΅W (red), 6.7 Β΅W (blue), 9.9 Β΅W (green), 14.8 Β΅W (purple), 21.2 Β΅W (yellow).

Page 38: Exploring the Blinking Dynamics of Eosin Y

32

3.4 Blinking Dynamics of EY on TiO2

As with previous studies in the Wustholz lab,32–35 studies of dyes on both glass and TiO2

substrate are necessary to fully comprehend the complete range of photophysical activity of a

sensitizer for solar energy and DSP applications. The blinking studies of EY performed on glass

in the presence and absence of oxygen are critical first steps to understanding the complex

dynamics of EY operative in DSP systems. For EY, it is crucial to conduct studies in both air and

N2, as done on glass, in order to understand the role of substrate in modifying the blinking

dynamics. The kinetic models proposed in Chapter 2 (Fig. 2.4) could be significantly altered

when a semiconductor like TiO2 is introduced.

Preliminary studies have been conducted (Fig. 3.2), and visual analysis of the dye scans

and blinking traces indicates that TiO2 changes the kinetic activity significantly. The observed

changes in air (Fig. 3.2 C

and D) are particularly

intriguing, as they differ

greatly from the

dynamics observed on

glass (Fig. 2.3). On glass,

the dye film scans were

very pixelated, but on

TiO2 individual

molecules can be

discerned (Fig. 3.2C),

indicating that fewer non-

Figure 3.2: False-colored image of 1Γ—10!! M EY on TiO2 in (A) N2 with (B) corresponding blinking trace and (C) in air with (D) corresponding blinking trace.

Page 39: Exploring the Blinking Dynamics of Eosin Y

33

emissive events longer than the 30-ms bin time are observed as compared to glass (Fig. 2.3). In

the blinking trace in air, there are long non-emissive events, which suggests that EY more readily

injects into the dark state on TiO2. The dynamics observed in N2 (Fig. 3.2 A and B) are not

visually drastically altered from that on glass, so no conclusions can be drawn from these data

yet. A full dataset of molecules in air and N2 need to be collected and analyzed with the MLE/KS

method in order to give a comprehensive understanding of the excited-state photophysics of EY

under more technologically relevant conditions.

By probing the complex excited-state dynamics, this research has provided new insights

into the ET dynamics of EY photosensitizers. Through the analysis presented here, proposed

kinetic models and corresponding test functions give a more complete understanding of how EY

behaves. SMS and robust statistical analysis reveal contributions from both dispersive ET and

ISC, as well as altered kinetics in the presence of oxygen. Moving forward, studies on TiO2 will

reveal more technologically relevant kinetic models. This study is a crucial first step towards the

development and design of next-generation DSP solar cell technology.

Page 40: Exploring the Blinking Dynamics of Eosin Y

34

References

(1) Lewis, N. S.; Nocera, D. G. Powering the Planet: Chemical Challenges in Solar Energy Utilization. Proc. Nat. Acad. Sci. 2006, 103 (43), 15729–15735.

(2) Cook, T. R.; Dogutan, D. K.; Reece, S. Y.; Surendranath, Y.; Teets, T. S.; Nocera, D. G. Solar Energy Supply and Storage for the Legacy and Nonlegacy Worlds. Chem. Rev. 2010, 110, 6474–6502. https://doi.org/10.1021/cr100246c.

(3) Pachauri, R. K.; Meyer, L.; Hallegatte France, S.; Bank, W.; Hegerl, G.; Brinkman, S.; van Kesteren, L.; Leprince-Ringuet, N.; van Boxmeer, F. AR5 Synthesis Report- Climate Change 2014; Pachauri, R. K., Meyer, L., Eds.; IPCC: Geneva, Switzerland, 2015.

(4) Caldeira, K.; Jain, A. K.; Hoffert, M. I. Climate Sensitivity Uncertainty and the Need for Energy without CO2. Science (80-. ). 2003, 299, 2052–2054.

(5) Coling, G. World Energy Assessment: Energy, the Environment, and Human Health; United Nations Development Programme: New York, NY, 2008; Vol. 64. https://doi.org/10.2105/ajph.64.12.1166-b.

(6) Willkomm, J.; Orchard, K. L.; Reynal, A.; Pastor, E.; Durrant, J. R.; Reisner, E. Dye-Sensitised Semiconductors Modified with Molecular Catalysts for Light-Driven H2 Production. Chem. Soc. Rev 2016, 45, 9–23. https://doi.org/10.1039/c5cs00733j.

(7) Haque, S. A.; Palomares, E.; Cho, B. M.; Green, A. N. M.; Hirata, N.; Durrant, J. R. Charge Separation Versus Recombination in Dye-Sensitized Nanocrystalline Solar Cells: The Minimization of Kinetic Redundancy. J. Am. Chem. Soc 2005, 127, 3456–3462.

(8) Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38.

(9) Cowan, A. J.; Durrant, J. R. Long-Lived Charge Separated States in Nanostructured Semiconductor Photoelectrodes for the Production of Solar Fuels. Chem. Soc. Rev 2013, 42, 2281–2293. https://doi.org/10.1039/c2cs35305a.

(10) Zhang, X.; Peng, T.; Song, S. Recent Advances in Dye-Sensitized Semiconductor Systems for Photocatalytic Hydrogen Production. J. Mater. Chem. A. 2016, 4, 2365–2402. https://doi.org/10.1039/c5ta08939e.

(11) Reynal, A.; Willkomm, J.; Muresan, N. M.; Lakadamyali, F.; Planells, M.; Reisner, E.; Durrant, J. R. Distance Dependent Charge Separation and Recombination in Semiconductor/Molecular Catalyst Systems for Water Splitting. Chem. Commun. 2014, 50, 12768–12771. https://doi.org/10.1039/C4CC05143B.

(12) Shimidzu, T.; Iyoda, T.; Koide, Y. An Advanced Visible-Light-Induced Water Reduced with Dye-Sentsitized Semiconductor Power Catalyst. J. Am. Chem. Soc. 1985, 107, 35–41.

(13) Hashimoto, K.; Kawai, T.; Sakata, T. The Mechanism of Photocatalytic Hydrogen Production with Halogenated Fluorescein Derivatices. Nouv. J. Chim. 1984, 8, 693–700.

(14) Moser, J.; Graetzel, M. Photosensitized Electron Injection in Colloidal Semiconductors. J. Am. Chem. Soc. 1984, No. 12, 6557–6564. https://doi.org/10.1021/ja00334a017.

(15) Lazarides, T.; McCormick, T.; Du, P.; Luo, G.; Lindley, B.; Eisenberg, R. Making Hydrogen from Water Using a Homogeneous System Without Noble Metals. J. Am. Chem. Soc. Comm. 2009, 16, 9192–9194. https://doi.org/10.1021/ja903044n.

(16) Hartley, C. L.; DiRisio, R. J.; Screen, M. E.; Mayer, K. J.; McNamara, W. R. Iron Polypyridyl Complexes for Photocatalytic Hydrogen Generation. Inorg. Chem. 2016, 55, 8865–8870.

Page 41: Exploring the Blinking Dynamics of Eosin Y

35

(17) Min, S.; Lu, G. Enhanced Electron Transfer from the Excited Eosin Y to Mpg-C3N4 for Highly Efficient Hydrogen Evolution from Water. J. Phys. Chem. C 2012, 116, 19644–19652.

(18) Misawa, H. Photosensitizing Action of Eosin Y for Visible Light Induced Hydrogen Evolution from WAter. Chem. Lett. 1983, 12, 1021–1024.

(19) Wang, J.; Liu, Y.; Li, Y.; Xia, L.; Jiang, M.; Wu, P. Highly Efficient Visible-Light-Driven H2 Production Via an Esoin Y-Based Metal-Organic Framework. Inorg. Chem. 2018, 57, 7495–7498.

(20) Prasad Hari, D.; KΓΆnig, B. Synthetic Applications of Eosin Y in Photoredox Catalysis. Chem. Commun 2014, 50, 6688–6699. https://doi.org/10.1039/c4cc00751d.

(21) Srivastava, V.; Singh, P. P. Eosin Y Catalysed Photoredox Synthesis: A Review. RSC Adv. 2017, 7, 31377–31392.

(22) Ardo, S.; Meyer, G. J. Photodriven Heterogeneous Charge Transfer with Transition-Metal Compounds Anchored to TiO2 Semiconductor Surfaces. Chem. Soc. Rev. 2009, 38 (1), 115–164. https://doi.org/10.1039/b804321n.

(23) McNeil, I. J.; Ashford, D. L.; Luo, H. L.; Fecko, C. J. Power-Law Kinetics in the Photoluminescence of Dye-Sensitized Nanoparticle Films: Implications for Electron Injection and Charge Transport. J. Phys. Chem. C 2012, 116, 15888–15899.

(24) Grollman, R.; Quist, N.; Robertson, A.; Rath, J.; Purushothaman, B.; Haley, M. M.; Anthony, J. E.; Ostroverkhova, O. Single-Molecule Level Insight into Nanoscale Environment-Dependent Photophysics in Blends. J. Phys. Chem. C 2017, 121, 12483–12494.

(25) Guo, L.; Wang, Y.; Lu, H. P. Combined Single-Molecule Photon-Stamping Spectroscopy and Femtosecond Transient Absorption Spectroscopy Studies of Interfacial Electron Transder Dynamics. J. Am. Chem. Soc. 2010, 132, 1999–2004.

(26) Wu, X. Y.; Bell, T. D. M.; Yeow, E. K. L. Electron Transport in the Long-Range Charge-Recombination Dynamics of SIngle Encapsulated Dye Molecules on TiO2 Nanoparticle Films. Angew. Chem., Int. Ed. 2009, 48, 7379–7382.

(27) Wang, Y. M.; Wang, X. F.; Gosh, S. K.; Lu, H. P. Probing Single-Molecule Interfacial Electron Transfer Dynamics of Porphyrin on TiO2 Nanoparticles. J. Am. Chem. Soc. 2009, 131, 1479–1487.

(28) Biju, V.; Micic, M.; Hu, D. H.; Lu, H. P. Intermittent Single-Molecule Interfacial Electron Transfer Dyanmics. J. Am. Chem. Soc. 2004, 126, 9374–9381.

(29) Godin, R.; Sherman, B. D.; Bergkamp, J. J.; Chesta, C. A.; Moore, A. L.; Palacios, R. E.; Cosa, G. Charge-Transfer Dynamics of Fluorescent Dye-Sensitized Electrodes under Applied Biases. J. Phys. Chem. Lett 2015, 6, 2688–2693.

(30) Sevinc, P. C.; Dhital, B.; Govind Rao, V.; Wang, Y.; Lu, H. P. Probing Electric Field Effect on Covalent Interactions at a Molecule-Semiconductor Interface. J. Am. Chem. Soc 2016, 138, 1536–1542.

(31) Rao, V.; Dhital, B.; Peter Lu, H. Single-Molecule Interfacial Electron Transfer Dynamics of Porphyrin on TiO2 Nanoparticles: Dissecting the Interfacial Electric Field and Electron Accepting State Density Dependent Dynamics. Chem. Commun. 2015, 51, 16821–16824.

(32) Tan, J. A.; Rose, J. T.; Cassidy, J. P.; Rohatgi, S. K.; Wustholz, K. L. Dispersive Electron-Transfer Kinetics of Rhodamines on TiO 2 : Impact of Structure and Driving Force on Single-Molecule Photophysics. J. Phys. Chem. C 2016, 120, 20710–20720. https://doi.org/10.1021/acs.jpcc.6b01960.

Page 42: Exploring the Blinking Dynamics of Eosin Y

36

(33) Wong, N. Z.; Ogata, A. F.; Wustholz, K. L. Dispersive Electron-Transfer Kinetics from Single Molecules on TiO2 Nanoparticle Films. J. Phys. Chem. C 2013, 117, 21075–21085. https://doi.org/10.1021/jp405899v.

(34) Tan, J. A.; Garakyaraghi, S.; Tagami, K. A.; Frano, K. A.; Crockett, H. M.; Ogata, A. F.; Patterson, J. D.; Wustholz, K. L. Contributions from Excited-State Proton and Electron Transfer to the Blinking and Photobleaching Dynamics of Alizarin and Purpurin. 2016. https://doi.org/10.1021/acs.jpcc.6b09818.

(35) Cassidy, J. P.; Tan, J. A.; Wustholz, K. L. Probing the Aggregation and Photodegradation of Rhodamine Dyes on TiO2. J. Phys. Chem. C 2017, 121, 15610–15618. https://doi.org/10.1021/acs.jpcc.7b04604.

(36) Roberto Batistela, V.; Silva Pellosi, D.; Dutra de Souza, F.; Ferreira da Costa, W.; Maria de Oliveira Santin, S.; Roberto de Souza, V.; Caetano, W.; Paulo MoisΓ©s de Oliveira, H.; Spacino Scarminio, I.; Hioka, N. PKa Determinations of Xanthene Derivates in Aqueous Solutions by Multivariate Analysis Applied to UV-Vis Spectrophotometric Data. Spectrochim. Acta A Mol. Biomol. Spec. 2011, 79, 889–897. https://doi.org/10.1016/j.saa.2011.03.027.

(37) Watkins, L. P.; Yang, H. Detection of Intensity Change Points in Time-Resolved Single-Molecule Measurements. J. Phys. Chem. B. 2005, 109, 617–628. https://doi.org/10.1021/jp0467548.

(38) Wustholz, K. L.; Bott, E. D.; Kahr, B.; Reid, P. J. Memory and Spectral Diffusion in Single-Molecule Emission. J. Phys. Chem. C 2008, 112, 7877–7885. https://doi.org/10.1021/jp711687j.

(39) Bott, E. D.; Riley, E. A.; B, K.; Reid, P. J. Proton-Transfer Mechanism for Dispersed Decay Kinetics of Single Molecule Isolated in Potassium Hydrogen Phthalate. ACS Nano 2009, 3, 2403–2411.

(40) Riley, E. A.; Bingham, C.; Bott, E. D.; Kahr, B.; Reid, P. J. Two Mechanisms for Fluorescence Intermittency of Single Violamine R Molecules. Phys. Chem. Chem. Phys 2011, 13, 1879–1887.

(41) Bott, E. D.; Riley, E. A.; Kahr, B.; Reid, P. J. Unraveling the Dispersed Kinetics of Dichlorofluorescein in Potassium Hydrogen Phthalate Crystals. J. Phys. Chem. A 2010, 114, 7331–7337.

(42) Riley, E. A.; Hess, C. M.; Whitham, P. J.; Reid, P. J. Beyond Power Laws: A New Approach for Analyzing Single Molecule Photoluminescence Intermittency. J. Chem. Phys 2012, 136, 184508. https://doi.org/10.1063/1.4717618.

(43) Clauset, A.; Rohilla Shalizi, C.; J Newman, M. E. Power-Law Distributions in Empirical Data. SIAM 2009, 51, 661–703.

(44) Hoogenboom, J. P.; den Otter, W. K.; Offerhaus, H. L. Accurate and Unbiased Estimation of Power-Law Exponents from Single-Emitter Blinking Data. J. Chem. Phys. 2006, 125 (20), 204713.

(45) Amat-Guerri, F.; LΓ³pez-GonzΓ‘lez, M. M. C.; MartΓ­nez-Utrilla, R.; Sastre, R. Singlet Oxygen Photogeneration by Ionized and Un-Ionized Derivatives of Rose Bengal and Eosin Y in Diluted Solutions. J. Photochem. Photobio. A. 1990, 53, 199–210.

(46) Alvarez-Martin, A.; Trashin, S.; Cuykx, M.; Covaci, A.; De Wael, K.; Janssens, K. Photodegradation Mechanisms and Kinetics of Eosin-Y in Oxic and Anoxic Conditions. Dye. Pigment. 2017, 145 (October), 376–384. https://doi.org/10.1016/j.dyepig.2017.06.031.

Page 43: Exploring the Blinking Dynamics of Eosin Y

37

(47) Garland, P. B.; Moore, C. H. Phosphorescence of Protein-Bound Eosin and Erythrosin. A Possible Probe for Measurements of Slow Rotational Mobility. Biochem. J 1979, 183, 561–572.

(48) Nishimura, Y.; Sakuragi, H.; Tokumaru, K. Xanthene Dye-Sensitized Electron Transfer to Methylviologen in Aqueous Organic Solution. Effects of Organic Solvents and Heavy Atoms in the Dyes. Bull. Chem. Soc. Jpn. 1992, 65, 2887–2893.

(49) Levin, P. P.; Costa, S. M. B. Kinetics of Oxygen Induced Delayed Fluorescence of Eosin Adsorbed on Alumina. The Dependence on Dye and Oxygen Concentrations. Chem. Phys. Lett. 2000, 320, 194–201.

(50) Zondervan, R.; Kulzer, F.; Orlinskii, S. B.; Orrit, M. Photoblinking of Rhodamine 6G in Poly(Vinyl Alcohol): Radical Dark State Formed through the Triplet. J. Phys. Chem. A 2003, 107, 6770–6776. https://doi.org/10.1021/jp034723r.

(51) Yeow, E. K. L.; Melnikov, S. M.; Bell, T. D. M.; De Schryver, F. C.; Hofkens, J. Characterizing the Fluorescence Intermittency and Photobleaching Kinetics of Dye Molecules Immobilized on a Glass Surface. J. Phys. Chem. A 2006, 110, 1726–1734. https://doi.org/10.1021/jp055496r.

(52) Penzkofer, A.; Beidoun, A.; Daiber, M. Intersystem-Crossing and Excited-State Absorption in Eosin Y Solutions Determined by Picosecond Double Pulse Transient Absorption Measurements. J. Lumin. 1992, 51, 297–314.

(53) John Albery, W.; Bartlett, P. N.; Paul Wilde, C.; Darwent, J. R. A General Model for Dispersed Kinetics in Heterogeneous Systems. J. Am. Chem. Soc 1985, 107, 1854–1858.

(54) Barbon, A.; Bott, E. D.; Brustolon, M.; Fabris, M.; Kahr, B.; Kaminsky, W.; Reid, P. J.; Wong, S. M.; Wustholz, K. L.; ZanrΓ©, R. Triplet States of the Nonlinear Optical Chromophore DCM in Single Crystals of Potassium Hydrogen Phthalate and Their Relationship to Single-Molecule Dark States. J. Am. Chem. Soc 2009, 131, 11548–11557. https://doi.org/10.1021/ja903284y.

(55) Yin, M.; Li, Z.; Kou, J.; Zou, Z. Mechanism Investigation of Visible Light-Induced Degradation in a Heterogeneous TiO2/Eosin Y/Rhodamine B System. Environ. Sci. Technol. 2009, 43, 8361–8366. https://doi.org/10.1021/es902011h.

(56) Mitsui, M.; Unno, A.; Mori, K. Methodology for Discriminating between Competitive Photophysical Processes in Photoblinking: Application to the Fluorescence Blinking of Single Dye Molecules Adsorbed on TiO2. Chem. Lett. 2017, 46, 866–869. https://doi.org/10.1246/cl.170127.

(57) Mitsui, M.; Unno, A.; Azechi, S. Understanding Photoinduced Charge Transfer Dynamics of Single Perylenediimide Dyes in a Polymer Matrix by Bin-Time Dependence of Their Fluorescence Blinking Statistics. J. Phys. Chem. C 2016, 120, 15070–15081. https://doi.org/10.1021/acs.jpcc.6b04114.

(58) Crouch, C. H.; Sauter, O.; Wu, X.; Purcell, R.; Querner, C.; Drndic, M.; Pelton, M. Facts and Artifacts in the Blinking Statistics of Semiconductor Nanocrystals. Nano Lett. 2010, 10, 1692–1698. https://doi.org/10.1021/nl100030e.

(59) Amecke, N.; Heber, A.; Cichos, F. Distortion of Power Law Blinking with Binning and Thresholding. J. Chem. Phys. 2014, 140, 1143061–1143068. https://doi.org/10.1063/1.4868252.

(60) Kuno, M.; Fromm, D. P.; Johnson, S. T.; Gallagher, A.; Nesbitt, D. J. Modeling Distributed Kinetics in Isolated Semiconductor Quantum Dots. Phys. Rev. B Condens. Matter Mater. Phys. 2003, 67, 125304.

Page 44: Exploring the Blinking Dynamics of Eosin Y

38

(61) Pelton, M.; Grier, D. G.; Guyot-Sionnest, P. Characterizing Quantum-Dot Blinking Using Noise Power Spectra. Appl. Phys. Lett. 2004, 85, 819–821.

(62) Hoogenboom, J. P.; van Dijk, E.; Hernando, J.; van Hulst, N. F.; Garcia-Parajo, M. F. Power-Law-Distributed Dark States Are the Main Pathway for Photobleaching of Single Organic Molecules. Phys. Rev. Lett. 2005, 95, 097401.

(63) Wustholz, K. L.; Bott, E. D.; Isborn, C. M.; Li, X.; Kahr, B.; Reid, P. J. Dispersive Kinetics from the Single Molecules Oriented in Single Crystals of Potassium Acid Phthalate. J. Phys. Chem. C 2007, 111, 9146–9156.

(64) Clifford, J. N.; Bell, T. D. M.; Tinnefield, P.; Heilemann, M.; Melnikov, S. M.; Hotta, J.; Silwa, M.; Dedecker, P.; Sauer, M.; Hofkens, J. Fluorescence of Single Molecules in Polymer Films: Sensitivity of Blinking to Local Environment. J. Phys. Chem. B. 2007, 111, 6987–6991.

(65) Rao, V.; Dhital, B.; He, Y.; Lu, H. P. Single-Molecule Interfacial Electron Transfer Dynamics of Porphyrin on TiO2 Nanoparticles: Dissecting the Complex Electronic Coupling Dependent Dynamics. J. Phys. Chem. C 2014, No. 118, 20209–20221.

(66) Niedre, M.; Patterson, M. S.; Wilson, B. C. Direct Near-Infrared Luminescence Detection of Singlet Oxygen Generated by Photodynamic Therapy in Cells in Vitro and Tissues in Vivo. Photochem. Photobio. 2002, 75, 382–391.

(67) Casadio, F.; Leona, M.; Lombardi, J. R.; Van Duyne, R. Identification of Organic Colorants in Fibers, Paints, and Glazes by Surface Enhanced Raman Spectroscopy. Acc. Chem. Res. 2010, 43, 782–791.

(68) Centeno, S. A.; Hale, C.; Caro, F.; Cesaratto, A.; Shibayama, N.; Delaney, J.; Dooley, K.; van der Snickt, G.; Janssens, K.; Stein, S. A. Van Gogh’s Irises and Roses: The Contribution of Chemical Analyses and Imaging to the Assessment of Color Changes in the Red Lake Pigments. Herit. Sci. 2017, 5, 18.