development of conjugated polymer nanoparticles for

71
Clemson University Clemson University TigerPrints TigerPrints All Theses Theses August 2021 Development of Conjugated Polymer Nanoparticles for Development of Conjugated Polymer Nanoparticles for Applications in Superresolution Imaging and Sensors Applications in Superresolution Imaging and Sensors Muskendol Novoa Delgado Clemson University, [email protected] Follow this and additional works at: https://tigerprints.clemson.edu/all_theses Recommended Citation Recommended Citation Novoa Delgado, Muskendol, "Development of Conjugated Polymer Nanoparticles for Applications in Superresolution Imaging and Sensors" (2021). All Theses. 3585. https://tigerprints.clemson.edu/all_theses/3585 This Thesis is brought to you for free and open access by the Theses at TigerPrints. It has been accepted for inclusion in All Theses by an authorized administrator of TigerPrints. For more information, please contact [email protected].

Upload: others

Post on 25-Apr-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Development of Conjugated Polymer Nanoparticles for

Clemson University Clemson University

TigerPrints TigerPrints

All Theses Theses

August 2021

Development of Conjugated Polymer Nanoparticles for Development of Conjugated Polymer Nanoparticles for

Applications in Superresolution Imaging and Sensors Applications in Superresolution Imaging and Sensors

Muskendol Novoa Delgado Clemson University, [email protected]

Follow this and additional works at: https://tigerprints.clemson.edu/all_theses

Recommended Citation Recommended Citation Novoa Delgado, Muskendol, "Development of Conjugated Polymer Nanoparticles for Applications in Superresolution Imaging and Sensors" (2021). All Theses. 3585. https://tigerprints.clemson.edu/all_theses/3585

This Thesis is brought to you for free and open access by the Theses at TigerPrints. It has been accepted for inclusion in All Theses by an authorized administrator of TigerPrints. For more information, please contact [email protected].

Page 2: Development of Conjugated Polymer Nanoparticles for

DEVELOPMENT OF CONJUGATED POLYMER NANOPARTICLES FOR APPLICATIONS IN SUPERRESOLUTION IMAGING AND SENSORS

A Thesis Presented to

the Graduate School of Clemson University

In Partial Fulfillment of the Requirements for the Degree

Master of Science Chemistry

by Muskendol Novoa Delgado

August 2021

Accepted by: Dr Jason McNeill, Committee Chair

Dr. Carlos Garcia Dr. Leah Casabianca

Dr. Dvora Perahia

Page 3: Development of Conjugated Polymer Nanoparticles for

ii

ABSTRACT

Utilization of conjugated polymer nanoparticles (CPNs) as fluorescent probes has

attracted lots of scientific interest due to their exceptional brightness and photostability,

particularly there has been considerable interest in developing a photoswitchable

nanoparticle based on CPNs. Previous efforts by other groups of researchers, either did not

exhibit single-step switching or suffer from poor on/off contrast because of incomplete

quenching of the nanoparticle fluorescence. In this work, superresolution imaging via

controlled reversible generation and recombination of hole polarons as fluorescence

quenchers inside CPNs has been demonstrated resulting in a localization precision of

~0.6nm, which is about 4 times better than the typical resolution achieved.

We demonstrated a novel polydot nanoparticle made of PCBM doped PFBT with

photoswitching properties for super resolution microscopy. Another efforts were made to

continue improving photoswitching polydots, including the development of PCBM doped

MEH-CNPPV photoswitches trying to achieve better colloidal stability and narrower size

distributions, also electric field modulation studies were performed on the nanoparticles and

new protocols to produce DNA wrapped polydots were developed in this work, attempting to

extend the scope of the applications of CPNs in biosensors, biological imaging and to

demonstrate improved super resolution imaging.

Page 4: Development of Conjugated Polymer Nanoparticles for

iii

ACKNOWLEDGMENTS

During my time at Clemson University, I have made lots of friends and received help from

many people. I want to thank all of them, even though I only have space to mention a few

in this dissertation.

It has been a lot of fun working in Dr. Jason McNeill’s lab. I would like to thank him for

all the compounds, optics, electrodes that I purchased, for the time he spent with my

manuscripts, and for his sound advises about my research as well as career. I enjoyed

talking to him about science, research and live because he always has some good stories to

share. Also, he was understandable in times when I was ill and could not perform research.

In the McNeill’s group I got to meet great people. Yifei Jiang, Teeranan Nongnual,

Amirhossein Jafariyan, Liaoran Cao and Ming Lei have all given me great help in research.

I would also like to thank my committee, Dr. Carlos Garcia, Dr. Leah Casabianca, and

Dr. Dvora Perahia, for their patience and guidance. I appreciate the research advises they

gave me and their support.

Finally, I would like to thank my family and relatives back in Colombia, specially my

parents and Stefany Collazos for all their love and support through all the difficult

moments.

Page 5: Development of Conjugated Polymer Nanoparticles for

iv

TABLE OF CONTENTS

Page

TITLE PAGE ............................................................................................................... i

ABSTRACT ................................................................................................................ ii

ACKNOWLEDGMENTS .......................................................................................... iii

LIST OF FIGURES ..................................................................................................... v

CHAPTER

I. INTRODUCTION ..................................................................................... 1

Conjugated polymers and CPNs ........................................................... 1 Fluorescence spectroscopy ................................................................... 6

Single Molecule Spectroscopy .............................................................12

II. EXPERIMENTAL METHODS ................................................................17

Materials .............................................................................................17 Preparation of CPNs ............................................................................17

Characterization Methods ....................................................................19 Single Molecule Spectroscopy .............................................................26

III. RESULTS .................................................................................................32

Improved superresolution imaging using telegraph noise in organic semiconductor nanoparticles .......................................................................................32

Attempts to improve the photoswitching CPNs: PCBM Doped MEH-CNPPV Nanoparticles ................................................................................................44

Electric Field Modulation of CPN’s fluorescence. ...............................50 DNA wrapped polydots development ..................................................53

IV. GENERAL CONCLUSIONS AND FUTURE DIRECTIONS...................56

REFERENCES...........................................................................................................58

Page 6: Development of Conjugated Polymer Nanoparticles for

v

LIST OF FIGURES

Figure Page

1.1 Chemical structures of four common conjugated polymer backbones ............. 1

1.2 SEM images of nanoparticles of the conjugated polymer PFBT with different shapes and sizes. ........................................................................................................... 4

1.3 Illustration of Beer’s Law derivation ............................................................ 7

1.4 Jablonski diagram showing the interactions of singlet and triplet states in a fluorophore…... ............................................................................................................ 8

1.5 Illustration of single molecule localization. ..................................................15

2.1 Illustration of typical conjugated polymer nanoparticles preparation ..............19

2.2 A typical AFM image of PFBT nanoparticles. ..............................................21

2.3 Scattering intensity fluctuations caused by Brownian motions of particles and the corresponding autocorrelation function. ......................................................................23

2.4 Illustration of the single particle fluorescence microscopy setup. ...................27

3.1 Chemical structures of PFBT and PCBM. Number weighted particle size distributions of PCBM doped………………….34

3.2 AFM images of PCBM doped PFBT CPNs. ..............................................35

3.3 Fluorescence microscopy images showing the blinking behavior of 20% PCBM doped PFBT CPNs. 3D histogram of the localized centroid position of a blinking 20% PCBM doped PFBT CPN. Fluorescence intensity trajectories of 20% PCBM doped PFBT CPNs, acquired at two excitation intensities, showing different blinking duty cycles ...............................37

3.4 Proposed photoswitching mechanism in PCBM-doped CPN photoswitches….40

Page 7: Development of Conjugated Polymer Nanoparticles for

vi

List of Figures (Continued)

Figure Page

3.5 AFM image of unlabeled e.coli. Fluorescence microscopy images of unlabeled, PFBT CPNs labeled, 40% PCBM doped PFBT CPNs labeled E.coli ..................................................................................................41

3.6 Fluorescence microscopy images showing the blinking behavior of an E. coli labeled with 20% PCBM doped PFBT CPNs. A 3D fluorescence intensity map showing a few switched “on” CPNs on an E. coli. A scatter plot shows the superresolution image of an E. coli, constructed from the localized positions ofblinking CPNs. ....................................................................................43

3.7 Number weighted particle size distributions of PCBM doped MEH-CNPPV CPNs at various PCBM doping percentages, determined from DLS measurements. Normalized fluorescence spectra of PCBM doped PFBT CPNs at various PCBM doping percentages………………………………………………………………………………...46

3.8 Fluorescence microscopy images of 1% PCBM doped MEH-CNPPV CPNs, 10% PCBM doped MEH-CNPPV CPNs, images acquired at 400 W/cm2, 10 Hz framerate. .....47

3.9 Trajectories of PCBM doped MEH-CNPPV CPNs, acquired at several framerates and power intensities……………………………………………………….48

3.10 Fluorescence intensity trajectories of PCBM CPNs modulated by -2V-0V square wave pulses acquired at 10 Hz framerate, and excitation power of 100 W/cm2.. ...............52

3.1 Normalized fluorescence spectra for the samples and the control experiment using

an excitation wavelength of 445 nm……………………………………………………….5

Page 8: Development of Conjugated Polymer Nanoparticles for

1

CHAPTER ONE

INTRODUCTION

1.1 Conjugated polymers and CPNs

1.1.1 Conjugated polymers and applications

Since the observation of semiconducting properties of conjugated polymers in 1977

by Heeger, MacDiarmid, and Shirakawa,1 they have attracted great interest from the

scientific community for their fluorescent and semiconducting properties2-6 and particularly

their electro-optic applications such as polymer light emitting devices and photovoltaic cells

because of their light weight and flexibility.7,8 Conjugated polymers are a special class of

polymers, containing alternate single and double bonds along the polymer backbone. The

π electrons can be delocalized along the length of polymer backbone,9,10 which gives

conjugated polymers organic semiconductor behavior. Some examples of commonly used

conjugated polymer backbones are conjugated Polyfluorenes (PFs), polyphenylene

vinylenes (PPVs), polythiophenes (PTs) and polyphenyl ethynylenes (PPEs), (structures

shown in Fig. 1.1).

Figure 1.1: Chemical structures of four common conjugated polymer backbones.

Page 9: Development of Conjugated Polymer Nanoparticles for

2

Delocalization of the π and π* energy levels and the moderate energy gap (in the

range of 1.5 to 3 eV), allow efficient absorption of photons in the visible, near UV, and

near IR spectral ranges. Variation of the sidechain chemistry or the inclusion of

heteroatoms, such as O, N, or S in the backbone or the sidechains, allow tuning of

photophysical and chemical properties of the conjugated polymer.1,11,12,13 In principle, the

electronic states of a conjugated polymer molecule would be delocalized over the entire

polymer molecule (assuming a perfectly ordered structure at absolute zero), however,

under typical conditions structural features such as kinks, chemical defects, and bends in

the polymer chain lead to lengths of 4-10 repeating units over which conjugation is

unbroken.14,15 Conjugation length plays an important role in determining the photophysical

properties of the polymer, as the energy level of an electronic state is a function of the

space available for delocalization, as explained by Hückel theory. Longer conjugation

lengths have lower energy electronic states leading to red shifted absorption and emission

wavelengths; conversely, short conjugation lengths lead to blue shifted spectra. This is

important to consider when designing devices, as solution and film processing each have a

dramatic impact on conjugation length and spectral characteristics. 16

1.1.2 Conjugated polymer nanoparticles

Conjugated polymer nanoparticles (CPNs) are nanoparticles made from conjugated

polymers. These nanoparticles have emerged as a promising class of multifunctional

photoluminescent nanomaterials particularly as fluorescent probes in biomedical analysis

and other advanced fluorescence imaging and analysis applications,17 because of their

Page 10: Development of Conjugated Polymer Nanoparticles for

3

small sizes, tunable emission wavelengths, high absorption coefficients, and excellent

fluorescence efficiencies. Conjugated polymers nanoparticles (CPNs) are by far the

brightest fluorescent nanoparticles demonstrated (roughly 20-100 times brighter than

similar-sized colloidal semiconductor quantum dots under typical imaging conditions),18,19

due to the high extinction coefficient, high chromophore density, and (in some cases) high

fluorescence quantum yield of the conjugated polymer.20 CPNs have been demonstrated in

several fluorescence-based applications, such as multiphoton fluorescence imaging,21

single nanoparticle sensors,22 photoswitching nanoparticles,19,23,24 particle tracking25 and

photodynamic therapy.26 Particularly is worth noting that photoswitching CPNs have

shown promise on applications of CPNs to improve signal levels and resolution in

photoswitching-based ultraresolution fluorescence imaging.27,28

Conjugated polymer nanoparticles (CPNs) size range may vary between a few

nanometers to microns depending on the preparation methods, and consist of somewhat

closely packed chains of hydrophobic conjugated polymers, typically with a disordered or

semicrystalline structure.29-31 Suspended in water, or cast onto a surface from an aqueous

suspension, the overall shape of the particle is typically governed primarily by interfacial

tension, and thus approximately spherical for smaller (<50 nm dia.) particles,32 is worth

nothing that being among soft nanoparticles, CPNs have particle shapes ranging from

spherical to ellipsoidal (Fig. 2.2). Particularly for our studies, we will focus on smaller (<20

nm) nanoparticles. There is interest in knowing the detailed structure of CPNs, including

polymer chain conformation and chain packing, largely motivated by the expectation

(based on knowledge of systems with similar chromophore density, such as organic

Page 11: Development of Conjugated Polymer Nanoparticles for

4

semiconductors and light-harvesting complexes) that structure dictates the optical

properties and related dynamics of CPNs.20

Figure 1.2: SEM images of nanoparticles of the conjugated polymer PFBT with different

shapes and sizes. (a) 28nm spheres with an aspect ratio of 1.0. (b) 140nm long ellipsoids with

an aspect ratio of 1.5. (c) 200nm long ellipsoids with an aspect ratio of 1.2. (d) 1,000 nm spheres

with an aspect ratio of 1.0. The white bar represents 200 nm in each plot. Adapted with

permission from ref. 32 Copyright 2005 Macmillan Publisher Ltd.

1.1.3 Exciton dynamics and polarons in conjugated polymers

When a photon is absorbed by a conjugated polymer, an electron undergoes

excitation from the π to the π* band, generating a neutrally charged local excitation and

polarization or distortion of the polymer matrix. The primary neutral electronic excited

state in conjugated polymer is typically described as the Frenkel-type molecular exciton,

consisting of two or more chromophores quantum-mechanically coupled via transition

dipoles.33,34 The nature of exciton motion in the system varies with the chemical structure,

conformation, and processing conditions (as well as other factors such as temperature).

Excitons can undergo either radiative or nonradiative decay, move to another chromophore

via either a Dexter or Förster mechanism, or lose an electron to leave behind a hole polaron

Page 12: Development of Conjugated Polymer Nanoparticles for

5

(discussed below). Exciton mobility in the polymer occurs through a combination of

Dexter electron transfer and Förster energy transfer. As the emission spectrum from the

donor chromophore needs to overlap with the absorption spectrum of the acceptor

chromophore, there tends to be a red shift in the energy of an exciton as it diffuses through

the particle, observable in the emission spectrum. There have been extensive investigations

of exciton dynamics in conjugated polymer systems, including thin films,35-37 solutions,38,39

in a matrix,40, 41 and nanoparticles.42

Polarons are either electrons or holes (electron vacancies) that exist in an organic

semiconductor. 43, 44 Polarons can be generated through chemical doping, electrode

oxidation or reduction, direct injection from an electrode, or exciton dissociation. A

positive or negative charge is created and localized to a segment of the material, and cause

the distortion of the nearby structure, which yields a localized charge volume. Polarons are

mobile, so they can act as charge carriers. Polarons able to move to different locations in

the material, acting as charge carriers in such devices as organic field effect transistors and

solar cells.

Hole polarons in conjugated polymers are typically generated when an electron is

lost from the exciton, leaving behind a positively charged hole and a polarization of the

polymer matrix. The absorption spectrum of a hole polaron is typically red-shifted as

compared to the neutral polymer, resulting in an absorption spectrum with a high degree

of overlap with the emission of the neutral polymer, which in turn results in highly efficient

energy transfer to hole polarons, characterized by a large energy transfer radius (4-8 nm),

enabling a small number of polarons to cause a large number of excitons in the vicinity to

Page 13: Development of Conjugated Polymer Nanoparticles for

6

undergo FRET. As emission from polarons is typically red shifted and has a low quantum

yield, thus a relatively small hole polaron density can nearly completely quench

fluorescence from the conjugated polymer. 45, 46

Polarons are very efficient singlet exciton quenchers43, 47, 48 because of the high

degree spectral overlap between the emission of neutral polymer and absorption of

polarons, which facilitates efficient energy transfer to polarons. Since the hole polarons

can have a dramatic effect on the performance of conjugated polymer devices, there is

considerable interest in studying and understanding polaron dynamics and related

phenomena, such as the fluorescence twinkling in conjugated nanoparticles.49

1.2 Fluorescence spectroscopy

1.2.1 Absorption and fluorescence

When molecules containing π electrons absorb energy from ultraviolet or visible

light, the π electrons can be excited from the ground state S0 (HOMO, π) to an excited state

S1 (LUMO, π*) this event is called absorption. This is characterized by the Beer-Lambert

law (also known as Beer’s law). Beer’s Law can be derived by treating the molecule as an

opaque disk with a cross-sectional area σ (cm2, the effective absorption area). Fig. 1.3,

illustrates the derivation of Beer’s law, where an infinitesimal slab with length dz (cm) and

total area S (cm2) is taken, the molecule density of the sample is N (molecules/cm3), then

the total number of molecules in the slab is N*S*dz. And the fraction of the total area where

light gets absorbed due to each molecule would be (fractional area for each molecule) σ/S.

If the intensity of the light incident at the infinitesimal slab is Iz, then the light intensity

Page 14: Development of Conjugated Polymer Nanoparticles for

7

absorbed by the slab would be 𝑑𝐼 = 𝐼𝑧 ∗ (𝜎𝑠⁄ ) ∗ (𝑁 ∗ 𝑆 ∗ 𝑑𝑧) = 𝐼𝑧 ∗ 𝜎 ∗ 𝑁 ∗ 𝑑𝑧. Then, we

can derive the following equation,

𝑑𝐼

𝐼𝑧= −𝜎𝑁𝑑𝑧. (1)

Figure 1.3. Illustration of Beer’s Law derivation.

In equation 1, the negative sign represents the absorption of light. Integrating from

z = 0 to z = l, the equation becomes,

𝑙𝑛𝐼 − 𝑙𝑛𝐼0 = − ln𝐼0

𝐼= −𝜎𝑁𝑙. (2)

where I0 is the intensity entering the sample, I is the intensity of light leaving the sample,

and l is the light path length (cm). Since the definition of the absorbance of the sample A

is given by 𝐴 = 𝑙𝑜𝑔10(𝐼0

𝐼). We can combine this equation with equation 2 getting 𝐴 =

1

2.303𝜎𝑁𝑙. Then replacing molecule density N (molecules/cm3) with concentration C (M,

Page 15: Development of Conjugated Polymer Nanoparticles for

8

mol/L), and cross section σ (cm2) with molar extinction coefficient ε (M-1 cm-1), the

equation becomes,

𝐴 = 𝜀 ∗ 𝐶 ∗ 𝑙, (3)

where 𝜀 = (1

2.303)(

1

1000)𝜎𝑁𝐴, 𝐶 =

𝑁∗1000

𝑁𝐴, and NA is the Avogadro’s number. It is

worth noting that the molar extinction coefficient of a sample depends on the wavelength of

the incident light.

For a molecule with a ground state S0, the photon can excite the molecule to the

first excited singlet state (S1) as shown in Fig. 4. The molecule may also have a triplet state

T1.

Figure 1.4: Jablonski diagram showing the interactions of singlet and triplet states in a

fluorophore.

In Fig. 1.4, a Jablonski diagram is shown to illustrate the relationship between the states

S0, S1, and T1 which represent a singlet ground state, singlet excited state and triplet excited

state, respectively along with the rate constants for interchange between the states and

energy transfer from a donor molecule to an acceptor molecule.50

The various rate constants are:

kabs = Excitation of the molecule through the absorption of a photon

Page 16: Development of Conjugated Polymer Nanoparticles for

9

kIC = Internal conversion rate

kR = Radiative rate (Fluorescence)

kNR = Non-radiative decay

kPB = Irreversible Photobleaching

kISC = Intersystem crossing to the triplet state

kP = Phosphorescence

Internal conversion is the relaxation from higher vibrational states to the ground

vibrational state of the S1 state, with a time constant on the order of 10-12 s, ensuring that

in most cases transitions from the S1 state (to the ground state or the triplet state) start

from its lowest vibrational state.

After the molecule is excited, it spends a certain time in the S1 state before it decays

via one of several possible pathways. The time constant for residence in the excited state

is called the fluorescence lifetime, τF, and results from the combination of processes

depopulating it, in equation 4, 𝑘𝑛𝑟 is the non-radiative rate (𝑘𝑛𝑟 = 𝑘𝑖𝑐 + 𝑘𝑖𝑠𝑐 + 𝑘𝐸𝑇 + 𝑘𝑏),

which include all non-radiative relaxation processes:

𝜏𝑓 =1

𝑘𝑟+𝑘𝑛𝑟. (4)

The fluorescence quantum efficiency of fluorophores is characterized by

fluorescence quantum yield (ϕf), which is the ratio of number of photons, emitted to number

of photons absorbed. It can also be expressed in terms of rate constants as follow:

𝜙𝑓 =𝑘𝑟

𝑘𝑟+𝑘𝑛𝑟. (5)

It is worth noticing that for good fluorophores, 𝑘𝑟 + 𝑘𝑖𝑐 ≫ 𝑘𝑖𝑠𝑐 + 𝑘𝑏, and kr is similar or

ideally higher than kic, so that the quantum yield for most useful fluorescent tags ranges

Page 17: Development of Conjugated Polymer Nanoparticles for

10

from a few percent to close to 100% (for R6G).51 Experimentally, the fluorescence

quantum yield can be measured with steady-state fluorescence spectroscopy and the

fluorescence lifetime can be measured with time-resolved fluorescence methods. The

radiative rate constant then is estimated by 𝑘𝑟 =𝜙𝑓

𝜏𝑓.

1.2.2 Förster Resonance Energy Transfer

It is possible for many photophysical systems to transfer energy, such as in the light

harvesting complex,52 or from one species to another such as in DNA molecular

beacons.53,54 This non-radiative transfer from donor to acceptor is accomplished through

either the electron transfer mechanism known as Dexter energy transfer, or Förster

resonance energy transfer. The Dexter energy transfer requires wavefunction overlap

between donor and acceptor and is a short-range interaction (< 2 nm) that requires spin

conservation.50 Förster resonance energy transfer involves the interaction of transition

dipoles of donor and acceptor and is a longer-range interaction (up to 10 nm). Förster

theory assumes a weak interaction between the transition dipoles of the donor and acceptor,

with the application of Fermi’s Golden Rule to estimate the transition rate. The theory

predicts that the energy transfer rate depends on intermolecular distance to the inverse sixth

power,

𝑘𝐸𝑇 =1

𝜏𝐷(𝑅0

𝑟)6, (6)

where τD is the donor fluorescence lifetime in the absence of the acceptor, R is the donor-

acceptor distance, and R0 is the energy transfer radius or Förster radius, defined as the

distance at which the energy transfer efficiency is 50%. The energy transfer rate constant

Page 18: Development of Conjugated Polymer Nanoparticles for

11

(Eq. 6), being highly distance dependent, can be used a ‘molecular ruler’ by attaching the

donor and acceptor to the same molecule and observing the degree of energy transfer.55

Based on the assumption of weakly interacting transition dipoles and some other

assumptions, Förster developed an expression for the energy transfer radius in terms of

determined spectroscopic properties:

R06 =

9000ln(10)ϕDκ2

128π5NAn4 ∫ fD(λ)

0ε𝐴(λ)λ

4dλ, (7)

where D is the quantum yield of the donor in the absence of acceptor; 2 is the

configurational factor (accounts for relative orientation of the donor and acceptor dipoles,

usually 2/3 in solution); n is the index of refraction of the medium; FD() is the pure donor

emission spectrum, normalized for the area under the curve; εA is the pure acceptor

emission spectrum; and is wavelength.50

The energy transfer efficiency E for FRET can be expressed as,

𝐸 =𝑅0

6

𝑟6+𝑅06. (8)

Equation 8, shows the strong dependence of the energy transfer efficiency on the donor-

acceptor distance. If the distance r decreases lower than R0, the efficiency increases to near

unity quickly. Conversely, the efficiency decreases dramatically to near 0 when r is greater

than R0. The energy transfer efficiency is usually measured using the relative donor

fluorescence intensity, in the presence (F) and the absence (F0) of acceptor,

𝐸 = 1 −𝐹

𝐹0. (9)

Page 19: Development of Conjugated Polymer Nanoparticles for

12

1.3 Single Molecule Spectroscopy

Single molecule spectroscopy was first demonstrated with doped crystals at low

temperatures56 and later at room temperature with near field scanning optical microscopy

(NSOM).57 More recently, most single molecule research utilizes confocal and wide field

microscopy arrangements.58,59

Fluorescence spectroscopy measurements in bulk solution typically only gives the

averaged properties of a large number of single molecules. In contrast, single molecule

spectroscopy (SMS),60,61 provides more per-molecule based molecular information, such

as the molecule’s structure, kinetics, and local environment, which is crucially important

when in heterogeneous systems, such as polymers, biological systems, nanostructured

interfaces, and nanoparticles. Another key advantage of single molecule measurements is

the ability to unravel complex dynamics, such as the sequence of events, which is

particularly useful for systems with complex, multi-step kinetic schemes that does not

follow conventional synchronization-based kinetics methods such as T-jump, pump-probe,

or stopped-flow. Single molecule fluorescence spectroscopy has been applied in various

research fields such as conjugated polymers,6,62-65 motor protein movements,66 and protein

folding.67

1.3.1 Single Molecule Excitation

The signal level in single molecule fluorescence measurements depends on several

factors, including the excitation intensity, optical cross-section and fluorescence quantum

yield of the fluorescent molecule, collection efficiency, and detector quantum efficiency.

The excitation rate of a single molecule can be estimated by analogy with Beer’s Law (Eq.

Page 20: Development of Conjugated Polymer Nanoparticles for

13

3), while the molar absorptivity (ε) parameter is most often used to describe macro-scale

experiments, the parameter used far more often at the single molecule level is absorption

cross section:

𝜎 =2.303𝜀

𝑁𝐴(10)

where σ is the absorption cross section (cm2) and NA is Avogadro’s number. This cross

section defines the effective area that a single molecule absorbs at a particular wavelength

of light.

The fluorescence cross-section is defined as the product of peak absorption cross

section (σ) and fluorescence quantum yield (ϕf) (𝜎𝑓 = 𝜎 × 𝜙𝑓). In bulk solution, the

absorption cross section and fluorescence quantum yield can be measured through steady-

state UV-Vis absorption and fluorescence emission experiments to obtain the brightness of

fluorophores. While at single particle level, it is difficult to directly measure the absorption

and quantum yield of a single fluorophore. Instead, the fluorescence cross section can be

determined through emitted photon counts per second divided by the excitation intensity

in photons per second per cm2 at single particle level (𝜎𝑓 =𝑁𝑝ℎ𝑜𝑡𝑜𝑛

𝐼𝑒𝑥).68

1.3.2 Single molecule localization and super-resolution microscopy

The resolution of conventional light microscopy is limited due to light diffraction

as shown by Ernst Abbe in 1873.69 The lateral resolution of a conventional optical

microscope is given by, 𝑑 =𝜆

2𝑁𝐴, where λ is the excitation wavelength and NA is the

numerical aperture of the microscope objective. A typical oil immersion microscope

objective (NA, ~1.25) with visible lights excitation yields ~200-250 nm resolution. Since

Page 21: Development of Conjugated Polymer Nanoparticles for

14

most biomolecules are much smaller than 250 nm, it is usually not possible to resolve the

individual components of biomolecular assemblies or their internal dynamics. Super-

resolution microscopy has been developed to obtain images with higher resolution than the

diffraction limit. There are two main types of functional super-resolution microscopy

methods: saturation-based methods (e.g. STED, GSD) and single molecule localization

based methods (e.g. PALM, STORM).2 In 2014, Eric Betzig, W. E. Moerner and Stefan

Hell were awarded the Nobel Prize in Chemistry for “the development of super-resolved

fluorescence microscopy”.

The principle for single molecule localization microscopy is through isolating

emitters and fitting the 2D intensity profile of the emitter with a point spread function (PSF,

approximately a Gaussian function) to obtain the location of the fluorophore. The process

is illustrated in Fig. 5.70 The localization precision is given by:

∆𝑙𝑜𝑐 = √(𝑠2

𝑁) + (

𝑎2 12⁄

𝑁) + (

8𝜋𝑠2𝑏2

𝑎2𝑁2 ), (11)

where N is the number of collected photons, s is the Gaussian width of PSF, a is the

effective pixel size of the detector and b the standard deviation of background noise in units

of photons. In the right side, the first term represents the effect of the photon counting

noise, while the second term represents the effect of finite pixel size of the detector and the

third represents background noise, including autofluorescence, scattered laser light,

thermal noise in the detector, and readout noise.

To enhance resolution, the goal is to maximize the photons collected from single

fluorophores and minimize background noise. In general, the more photons collected, the

better the resolution. Ideally, if there are 10,000 photons collected from a single fluorescent

Page 22: Development of Conjugated Polymer Nanoparticles for

15

probe with negligible background noise, the localization of the center position can be

determined with 1-2 nm accuracy. Background noise usually comes from the sample

autofluorescence, dark current, readout noise and some residual fluorescence from other

nearby fluorophores. Therefore, for single molecule-based super-resolution imaging,

fluorophores with large number of photons emitted prior to bleaching, and high contrast

over background are needed, and for high-speed tracking, a high emission rate at or below

saturation is also required.

Figure 1.5. Illustration of single molecule localization: (a) 2D image data of a single

fluorophore (b) Gaussian function fitting to obtain the center position.

Super resolution fluorescence microscopy based on single molecule localization

provides a substantial improvement in resolution over conventional fluorescence

microscopy and has been successfully applied to many important systems. The techniques

typically require a photoswitching fluorescent dye with certain characteristics. In our lab

we pioneered the development of a novel class of fluorescent nanoparticles based on

fluorescent conjugated polymer molecules49, 71 and in previous work combined with some

Page 23: Development of Conjugated Polymer Nanoparticles for

16

of the findings of this thesis, a photoswitching polydot nanoparticle was developed,

probing to be much brighter and photostable than typical photoswitching dyes.24, 72

Page 24: Development of Conjugated Polymer Nanoparticles for

17

CHAPTER TWO

EXPERIMENTAL METHODS

2.1 Materials

The conjugated polymers, poly[(9,9-dioctylfluorenyl-2,7-diyl)-co-(1,4-benzo-{2,1’,3}-

thiadiazole)] (PFBT, MW 10,000, polydispersity 1.7), poly[2-methoxy-5-(2-

ethylhexyloxy)-1,4-(1-cyanovinylene-1,4-phenylene)] (MEH-CN-PPV, MW 27000-

28000, polydispersity, 4.7) were employed in this study and purchased from ADS Dyes,

Inc. (Quebec, Canada). Tetrahydrofuran (THF, HPLC grade, 99.9%), phenyl-C61-butyric

acid methyl ester (PCBM, 99.5%) and (3-Aminopropyl) trimethoxysilane (APS, 97%)

were purchased from Sigma-Aldrich (Milwaukee, WI). All chemicals were used as

received without further purification.

2.2 Preparation of CPNs

A nano-precipitation method is employed to prepare conjugated polymer nanoparticles

(CPNs), as reported previously.73,74 This method is modified from the reprecipitation

procedure developed by Kurokawa and collaborators,75 and it is chosen because of its ease

and good reproducibility. This procedure involves dissolving hydrophobic conjugated

polymers in a water-miscible solvent such as THF and fast injection of the polymer solution

into deionized (DI) water. The rapid change of solvent quality by mixing results in polymer

chain collapse, leading to the formation of polymer nanoparticles. Then, the removal of

THF yields a clear aqueous nanoparticle suspension. There is competition between polymer

aggregation and nanoparticle formation (chain collapse) during the reprecipitation process,

Page 25: Development of Conjugated Polymer Nanoparticles for

18

and the combination of lower concentration with faster mixing rate disfavors aggregation

and favors smaller particle generation. Therefore, we are using a 20 ppm injection

concentration of polymer solution in this research. Higher concentrations tend to lead to a

larger fraction of the polymer lost due to aggregation, and/or larger nanoparticle sizes.

For a typical PFBT nanoparticle preparation, 10 mg of PFBT polymer is dissolved in 10

g of inhibitor-free HPLC grade THF with stirring to produce a 1000 ppm PFBT stock

solution. Then the stock solution was diluted to 20 ppm, and 2 mL of the polymer solution

was injected rapidly into 8 mL of DI water via micropipette. The THF solvent was removed

under vacuum at 40 °C for ~6 hours, and finally the sample was filtered through a 100 nm

PVDF membrane filter (Millipore) to remove aggregates. The PFBT particle preparation

process is illustrated in Fig. 2.1. Typically, less than 10% of the polymer was removed by

the filtration, as determined by UV-Vis absorption spectroscopy, indicating that most of

the polymers formed nanoparticles. The resulting suspension was clear and stable for

weeks without aggregation. The formation of CPNs is governed by hydrophobic

interaction, interfacial tension as well as surface free energy. The size of the CPNs can be

controlled by multiple steps in the preparation procedure. Rapid mixing of THF and water

results in a sudden decrease in solvent quality and favors small particle formation. In

contrast, large aggregates are typically observed when conjugated polymer THF solution

is added drop-by-drop into water. In addition, increase of the precursor concentration

typically results in larger particle sizes and larger fraction of polymer loss during the

filtration due to aggregation.

Page 26: Development of Conjugated Polymer Nanoparticles for

19

Similar procedures can be employed to prepare doped CPNs. To prepare PCBM doped

PFBT nanoparticles, conjugated polymer PFBT and fullerene derivative PCBM were

dissolved in THF and diluted to 20 ppm. The solutions of PFBT and PCBM were mixed in

various ratios to form precursors of varying dopant percentages (5%, 10%, 20%, 40%). 2

mL of the precursor solution was rapidly injected into 8 mL of water under mild sonication.

Then the THF solvent was removed by partial vacuum evaporation. After evaporation, the

sample was filtered through a 100 nm membrane filter (Millipore) to remove aggregates.

The same methods and procedures described above, were applied to prepare CNPPV

nanoparticles and PCBM doped CNPPV nanoparticles adjusting the doping percentages.

Figure 2.1. Illustration of typical conjugated polymer nanoparticles preparation.

2.3 Characterization Methods

We employed a variety of spectroscopic and microscopic techniques to investigate the

photophysical properties of CPNs. Atomic force microscopy (AFM) and dynamic light

Page 27: Development of Conjugated Polymer Nanoparticles for

20

scattering (DLS) were used to determine the size distribution of CPNs. Steady-state

absorbance and emission of CPNs were studied through UV-Vis and fluorescence

spectroscopy, respectively. Single nanoparticle spectroscopy was utilized to investigate

single particle properties, such as kinetics, photon number, and single particle spectra.

2.3.1 Atomic force microscopy (AFM)

Atomic force microscopy (AFM) is one of the scanning probe microscopy family, and

is usually used to characterize the surface properties of a material by scanning the surface

with a small tip.76 There are several different AFM scanning modes, which includes contact

mode, non-contact mode, and intermittent contact mode (also known as “tapping” mode).

Tapping mode was employed for the AFM scanning in this study, since it is typically the

method of choice for soft samples and/or samples that are weakly adhered to the surface.

The AFM tip is attached to a cantilever, which oscillates in the Z direction while the tip is

scanned in the X and Y directions across the sample surface. The tapping mode scanning

is performed by oscillating the cantilever at or near the cantilever’s resonant frequency (in

the range of 70-200 kHz) with a piezoelectric ceramic. The signal used to control the

cantilever’s oscillation (and damping) in the Z feedback control loop is detected by a photo

diode through the reflection of a laser beam from AFM cantilever back. During tapping

mode scanning, the cantilever oscillation amplitude is kept constant by adjusting the tip

distance from the surface through the feedback controller, called proportional-integral-

derivative controller (PID). The surface morphology image is assembled with a series of

scanning lines while the piezo moves the tip along the sample.

Page 28: Development of Conjugated Polymer Nanoparticles for

21

In this study, isolated particles immobilized on coverslips were imaged with an Ambios

Q250 multimode AFM in tapping mode. The scan range of the instrument is 40 μm in the

XY direction, and 5.7 μm in the Z direction. The scanning parameters employed in this study

were scan area of 5 μm×5 μm, 500 lines/scan, and a scan rate of 0.5 Hz (per line) with a

pixel resolution of 10 nm. The tip used in this work is Q-WM190 (NanoAndMore) at

tapping mode with mean resonance frequency of 190 kHz. The general guidelines for

setting PID parameters depend on the surface; to image flat surfaces with low scan rates

over small area, low Integral and Proportional gains are adequate; conversely, higher gains

are needed for rough surfaces at fasting scanning with large scan area. For typical AFM

images, the density of particles is ~10 per μm2. A typical AFM and a line scan for CPNs is

shown in Fig. 2.2. Since the particles are roughly spherical, the particle height was taken

as particle diameter due to the higher vertical resolution (Z direction) compared to the

lateral resolution (X-Y direction) as well as the tip convolution effect that typically

broadens the image of small particles in the XY plane.

Figure 2.2 A typical AFM image of PFBT nanoparticles (left), size distribution of the CPNs.

0 5 10 15 20 25 30 350

20

40

60

80

100

Nanoparticle Height (nm)

Occ

urre

nce

Page 29: Development of Conjugated Polymer Nanoparticles for

22

In this work, AFM samples were prepared on glass coverslips. Those coverslips were

cleaned with following steps: soap water, DI water, Nochromix/H2SO4 solution, and DI

water. Then the coverslips were functionalized with amine groups by dropping 50 μL

0.0005 M freshly prepared APS/ethanol solution onto the coverslip for ~2min, followed by

rinsing with DI water. The coverslip was then submerged into a ~20 times diluted CPNs

solution for 40 min. Excess CPNs were removed with DI water and the coverslip was dried

with nitrogen gas. APS covered coverslip was scanned as a control sample.

2.3.2 Dynamic Light Scattering (DLS)

Dynamic light scattering (DLS), also known as photon correlation spectroscopy, is a

method for determining size distribution in solution of particles such as nanoparticles,

polymers, proteins, and colloids. DLS measurement involves illuminating the sample with

a laser. The scattered light from different particles in solution can interfere constructively

(in-phase) or destructively (out-of-phase) and results in the speckle pattern. As particles

undergo Brownian motions in solution, the intensity of the spots on the speckle pattern will

fluctuate overtime due to changes in the distances between particles. The intensity

fluctuations thus contain the time scale of the particle movements and can be used to

determine the translational diffusion coefficients of the particles through autocorrelation

analysis. Applying Stokes-Einstein equation, the hydrodynamic diameters of the particles

can be calculated from the translational diffusion coefficients. Since the diffusion

coefficient is also temperature dependent, DLS measurement is typically performed under

isothermal environment. In addition, at high particle concentrations, multiple scattering

Page 30: Development of Conjugated Polymer Nanoparticles for

23

and viscosity effect occur, which could distort size distributions obtained from DLS

measurement. Therefore, DLS is typically performed with diluted samples (sometimes

involves a series of dilutions and extrapolating the result to the limit of zero concentration).

When the particle size is small compared to the wavelength of the laser (< λ/10), the

scattering process is isotropic, known as Rayleigh scattering. However, when the particle

size is comparable to the wavelength of the laser, the scattering intensity shows strong

angle dependence, known as Mie scattering.77 As a result, at certain angles, the scattering

intensity of some particles will completely overwhelm the signal from the others. In

addition, the scattering angle also controls the decay rates 𝛤 in the autocorrelation function

(𝛤 = 𝑞2𝐷, where D is the translational diffusion coefficient, q is the wave vector,

depending on the scattering angle), thus could affect data analysis. For highly dispersed

samples (especially with large particles), it is necessary to perform DLS at multiple angles

in order to obtain complete information about the particle size distribution.

Figure 2.3. (a) Scattering intensity fluctuations caused by Brownian motions of small particles and

the corresponding autocorrelation function. (b) Scattering intensity fluctuations caused by

Brownian motions of large particles and the corresponding autocorrelation function.

Page 31: Development of Conjugated Polymer Nanoparticles for

24

In this work, DLS measurements of CPNs were performed using a Nanobrook Omni

(Brookhaven Instruments Corporation, NY) with BIC Particle Solutions Software (V 3.1).

The CPN suspensions were diluted to a peak absorbance of ~0.1. Typical conditions

employed for the DLS measurements were 25 °C, count rate of 500 kcounts/s, scattering

angle of 90°, and acquisition time of 180 seconds. Polystyrene spheres (Thermo Fisher, 24

nm) were used as a size standard. Typically, CPNs suspension consists of mostly small

nanoparticles with small populations of larger aggregates (>100 nm). The NNLS method

provided reliable fitting results of the particle size distributions, which were highly

consistent with the results from AFM measurements.

2.3.3 UV-Vis and fluorescence spectroscopy

The UV-Vis absorption spectra were collected with a Shimadzu UV-2101PC scanning

spectrophotometer using 1 cm cuvettes. The instrument is equipped with two light sources,

which allows the scanning wavelength from 190-900 nm. A photomultiplier tube (PMT) is

employed as the detector in the system for sensitive absorbance measurements. To

determine the molar extinction coefficients/absorption cross-section of conjugated

polymers, conjugated polymers were dissolved in THF and diluted to a peak absorbance of

~0.1. According to the Beer’s Law, the molar extinction coefficient is given by 𝜀 = 𝐴/𝑙𝑐,

where A is the absorbance of the sample, l is the sample path length (1 cm) and c is the

molar concentration of conjugated polymer. If we replace the molar concentration c in the

Beer’s law with molecule density N (molecules/cm3), we can calculate the absorption

cross-section 𝜎 using equation 𝜎 = 2.303 ∗ 𝐴/𝑙𝑁. The extinction coefficient and

Page 32: Development of Conjugated Polymer Nanoparticles for

25

absorption cross-section of CPNs were calculated from the molar extinction coefficients

absorption cross-section of the corresponding conjugated polymer times the number of

polymer molecules per nanoparticle Nnp. Nnp is determined from the mean nanoparticle

volume (calculated from nanoparticle diameters determined by AFM) and the polymer

molecular weight (assuming a polymer density of ~1 g/cm3).

The fluorescence emission spectra were obtained with a commercial fluorometer

(Quantamaster, PTI, Inc.). The instrument has xenon arc lamp as excitation source. The

monochromators for excitation and emission have 1/4 m focal length with grating of 1200

grooves/mm. The detector employed in the instrument is a photomultiplier tube (PMT,

model 814) in photon counting mode. All the slit widths used in this work are 0.5 mm,

yielding a wavelength resolution of 2 nm (4 nm/mm for a 1200 grooves/mm grating, 1/4

m focal length). The fluorescence quantum yields of CPN samples were determined with

the fluorescent dye fluorescein in 0.01 M sodium hydroxide (pH = 12) used as standard.

All the CPN samples and the fluorescein standard solution were diluted to yield an

absorbance of ~0.05 at an excitation wavelength of 473 nm. The calculation of sample

quantum yield is employed using the following equation with the literature quantum yield

value of 0.92 for the fluorescein standard,78, 79

𝜙𝑓(𝑥) =𝐴𝑠𝐹𝑥𝑛𝑥

2

𝐴𝑥𝐹𝑠𝑛𝑠2𝜙𝑓(𝑠). (12)

In this equation, 𝜙𝑓is the fluorescence quantum yield, A is the absorbance at the excitation

wavelength, F is the integrated area under the corrected fluorescence emission curve, and

n is the solvent refractive index. Subscripts s and x represent standard and unknown sample,

respectively. For consistency, fluorescein in aqueous NaOH (pH = 12) is used as standard

Page 33: Development of Conjugated Polymer Nanoparticles for

26

and 473 nm is employed as excitation wavelength in the fluorescence quantum yield

measurement of all the samples in this research.

With the fluorescence spectrum of donor and absorption spectrum of acceptor, the

Förster radius (R0) between energy transfer pairs can be calculated using the following

equation.

R06 =

9000ln(10)ϕDκ2

128π5NAn4 ∫ fD(λ)

0ε(λ)λ4dλ, (13)

where NA is Avogadro’s number; n is the refractive index of donor κ2, the orientation

factor is assumed as 2/3;50 ϕD, fD(λ), ε𝐴(λ) are the quantum yield of donor, normalized

fluorescence spectra of donor, and extinction coefficient of acceptor in THF solution,

respectively.

2.5 Single Molecule Spectroscopy

2.5.1 Single Particle Imaging

Single particle imaging was performed with a custom wide-field epifluorescence

microscope described as follows (Fig. 2.4). A 445 nm excitation laser (Thorlabs, LP450)

is passed through a liquid crystal noise eater (Thorlabs, NEL01) and then is guided to the

rear epi port of an inverted fluorescence microscope (Olympus IX-71) by an optical fiber.

The laser beam is reflected by a 500 nm long-pass dichroic (Chroma 500 DCLP) to a high

numerical aperture objective (Olympus Ach, 100×, 1.25 NA, Oil). At the sample plane, the

laser excitation exhibits a Gaussian profile with fwhm of ∼5 μm. Various laser intensities

were employed for the single particle imaging. Depending on the experiment, the typical

excitation power at the center of the laser spot ranges from 10-500 W/cm2. CPNs were

dispersed on a glass coverslip by the same method used for preparing AFM samples.

Page 34: Development of Conjugated Polymer Nanoparticles for

27

Figure 2.4. Illustration of the single particle fluorescence microscopy setup.

Fluorescence from the CPNs is collected by the objective lens, passes through a 500 nm

long-pass filter and then is focused onto a sCMOS detector (Andor, Neo sCMOS), yielding

a pixel pitch of 66 nm/pixel, as determined using a calibration slide. A piezoelectric

scanning stage (P-517.3CL, Polytec PI) is used to position the sample. The overall

fluorescence detection efficiency of the microscope is ~3%, determined from nile red

loaded polystyrene beads (Thermo Fisher) as standards. Depending on the experiment, a

variety of framerates ranging from 10 Hz to 1 kHz were used. At fast framerates, the region

of interest was set to the center of the sCMOS chip to maximize the performance of the

camera. When the number of photons detected per pixel is small comparing to the well-

depth, the typical camera settings are 11 bits per pixel, gain setting of 0.6, rolling shutter

mode. If the particles are bright and the number of photons detected per pixel is comparable

to the well-depth, the typical camera settings are 16 bits per pixel, gain setting of 1.6, rolling

Page 35: Development of Conjugated Polymer Nanoparticles for

28

shutter mode. The experimental gain factors are determined by analysis of photon counting

noise for a flat field, which are consistent with the factory values.

2.5.3 Single Particle Localization

The localization analysis of CPNs was performed using custom scripts written in Matlab

(Mathworks). To determine small movements in the fluorescence centroid associated with

polaron motion, the scripts first sum over all the frames, find the intensity of the brightest

pixel in the summed image and use 5-10% of the intensity as a threshold to distinguish

CPNs from the background noise. Then the scripts search for pixels above the threshold in

the summed image and compare their intensities to the neighboring pixels to roughly locate

the central pixel of the fluorescence spot of each CPN. Based on the rough locations of

CPNs, a 2D Gaussian function was fit to the fluorescence spot of each CPN, frame by

frame. Typically 9×9 pixels were used for the fitting (4 pixels on each side of the central

pixel). The scripts check the FWHM obtained from the fitting (should be around ~270 nm)

to make sure it is not from multiple CPNs that are close to each other. The centroid position

trajectory of a CPN was constructed from the centroid positions determined from the

corresponding fluorescence spot in all the frames.

A slightly different method was used to determine the position of photoswitching CPNs.

When imaging certain samples, sometimes background subtraction is needed to ensure

optimal localization accuracy. The background fluorescence was determined from the

frames with no CPN switched “on”. ~100 frames of the background fluorescence were

averaged and subtracted from each frame prior to analysis. After the subtraction, the scripts

Page 36: Development of Conjugated Polymer Nanoparticles for

29

first find the intensity of the brightest pixel of all frames, then use 5-15% of the intensity

as a threshold to distinguish CPNs from the background noise (a higher threshold might be

needed, if the autofluorescence is not effectively corrected). For each frame, the scripts

search for pixels above the threshold and compare their intensities to the neighboring pixels

to roughly locate the central pixel of the fluorescence spot of each CPN. Then the

fluorescence centroid positions of the CPNs were determined by nonlinear least-squares

fitting of a 2D Gaussian function to the fluorescence spots. Typically 9×9 pixels were used

for the fitting (4 pixels on each side of the central pixel). The scripts check the FWHM

obtained from the fitting to make sure it is not from multiple nearby CPNs. The localized

positions obtained from each frame were then used to construct the shape of the sample.

In general, there are typically a few percent of the CPNs showing minimal photoblinking,

consistent with large particles or aggregates. It was observed that the non-blinking particles

or aggregates in the image exhibit highly correlated slow movement in the centroid

position, due to drift and vibrations from the environment. For each frame, the non-blinking

particles were used as markers to measure at each frame how far the sample travelled

relative to the previous frame, allowing us to apply a vibration correction needed for the

experiment. The displacement was then subtracted from the localized positions of all the

CPNs in the current frame. The experimental localization precision can be measured in

multiple ways. For photoswitching CPNs, the experimental localization precision was

determined from the standard deviation of the localized position distribution (i.e., the per-

frame position uncertainty is obtained from a sequence of consecutive frames during an

“on” cycle). For CPNs showing clear movements in the fluorescence centroid caused by

Page 37: Development of Conjugated Polymer Nanoparticles for

30

polaron motion, the experimental localization precision was determined from the root mean

square displacement of the centroid position trajectory at lag time 0. Under typical imaging

conditions, due to the high brightness of CPNs, the number of photons used in the

localization is typically large (>2000). The experimentally determined localization

precision is consistent with the theoretical localization uncertainties determined from the

equation 𝜎 = 𝑠 √𝑁⁄ , where s is the standard deviation (STD) of the PSF and N is the

number of detected photons used in the fitting. When the number of photons used in the

localization is small (<1000), addition source of localization uncertainty should be

considered. Thompson and collaborators modeled the effect of pixel size and background

noise on the localization precision (assuming a gaussian PSF), which is given by the

following expression,70

𝜎 = √𝑠2

𝑁+𝛼2/12

𝑁+8𝜋𝑠4𝑏2

𝛼2𝑁2,(14)

where s is the STD of the PSF and N is the number of detected photons used in the fitting,

𝛼 is the pixel size, b is the background noise. STD of the single particle fluorescence spot

is 130 nm. The pixel size of the setup is 66 nm and the background noise (due primarily to

readout noise, autofluorescence, and scattered light) ranges from roughly 1.5 to 10,

depending on conditions. Assuming N=2000 and b=1.5, we obtained a localization

precision of ~3.0 nm. The first term (shot noise) on equation 14 contributed to 95% of the

uncertainty. The second (pixel size) and the third term (background noise) only contributed

to 5%. If we lower the number of detected photons to 500, the shot noise term still

contributed to 85% of the uncertainty, the background noise contributed to 14% and the

Page 38: Development of Conjugated Polymer Nanoparticles for

31

pixel size term only contributed to <1%. According to these results, the shot-noise-limited

expression, 𝜎 = 𝑠 √𝑁⁄ is fairly accurate for most of the conditions. Only when N<1000,

background noise needs to be considered. However, in some cases, due to the presence of

a high background caused by autofluorescence of the sample or residual fluorescence of

dyes or particles in the “off” state, background subtraction is required to perform single

particle localization. The subtraction procedure can introduce additional uncertainty (for

example, shot noise from autofluorescence), and thus equation 14, would not be a reliable

estimate of the overall localization uncertainty. In such cases, experimental uncertainty is

typically compared to computer-generated quasi-random data with noise characteristics

that closely match the experiment. In other words, a Monte Carlo approach to estimating

the expected localization uncertainty is performed, based on the standard Monte Carlo

approach for propagating uncertainty.80

Page 39: Development of Conjugated Polymer Nanoparticles for

32

CHAPTER THREE

RESULTS

3.1 Improved superresolution imaging using telegraph noise in organic

semiconductor nanoparticles

Reprinted and adapted with permission from Ref 24. Copyright (2017) American Chemical

Society.

In recent years, driven by the interest of studying biological structures and processes

beyond the diffraction limit, superresolution imaging techniques have undergone rapid

growth. Various photophysical processes, such as, photoactivation,27 photoswitching,28

stimulated emission81 and inter-system crossing28, have been utilized to overcome the

diffraction limit. The development of superresolution fluorescent probes, including

photoactivatable proteins,27, 82 photoswitchable dyes,83, 84 and nanoparticles85, 86 have

improved the spatial resolution of optical microscope to <10 nm, which enables us to image

cellular structures in an unprecedented level of detail.

To further improve the spatial resolution of superresolution microscopy,

photoswitchable fluorescent probes with high brightness, good reversibility, and high

on/off contrast are required. Owing to the exceptional brightness and photostability of

CPNs,19-22 there has been considerable interest in developing photoswitchable nanoparticle

based on CPNs.23, 86 Previous efforts in modifying CPNs or other fluorescent nanoparticles

for superresolution imaging typically involve incorporating or linking photochromic dyes

Page 40: Development of Conjugated Polymer Nanoparticles for

33

to the nanoparticles. However, these nanoparticles either do not exhibit single-step

switching or suffer from poor on/off contrast due to incomplete quenching of the

nanoparticle fluorescence. In this work, we demonstrate a novel method that utilizes

controlled reversible generation and recombination of hole polarons inside CPNs to

achieve superresolution imaging. Hole polarons are known to be highly efficient

fluorescence quenchers in conjugated polymer systems.45, 63, 87, 88 Nanoparticles of the

conjugated polymer PFBT doped with the fullerene derivative PCBM, rapidly establish a

large population of hole polaron charge carriers sufficient to nearly suppress nanoparticle

fluorescence. However, fluctuations in the number of charges lead to occasional bursts of

fluorescence, which is similar to the random telegraph signal noise observed in

semiconductor devices.89 The repeated, spontaneous generation of short, intense bursts of

fluorescence photons (3-5×104 photons detected per switching event, on average) are

roughly 1-2 orders of magnitude brighter than those of photoswitching dye molecules,

resulting in a localization precision of ~0.6 nm, about 4 times better than the typical

resolution obtained by localization of dye molecules.90

3.1.1 Nanoparticle Preparation and Characterization

Nanoparticles of the conjugated polymer PFBT doped with various percentages of

PCBM were prepared using nano-precipitation method described previously.71, 74, 91 20

ppm solutions of PFBT and PCBM in THF were prepared and filtered through a 0.45 μm

membrane filter (Millipore). The solutions of PFBT and PCBM were mixed in various

ratios to form precursors of varying dopant percentages (5%, 10%, 20%, 40%), 2 mL of

the precursor solution was rapidly injected into 8 mL of water under mild sonication. Then

Page 41: Development of Conjugated Polymer Nanoparticles for

34

the THF solvent was removed by partial vacuum evaporation. After evaporation, the

sample was filtered through a 100 nm membrane filter (Millipore) to remove aggregates.

Figure 3.1. (a) Chemical structures of PFBT and PCBM. (b) Number weighted particle size

distributions of PCBM doped PFBT CPNs at various PCBM doping percentages, determined from

DLS measurements. (c, d) Normalized absorption and fluorescence spectra (excited at 450 nm) of

PCBM doped PFBT CPNs at various PCBM doping percentages.

The size distribution of the nanoparticles was determined using AFM and DLS. No

significant difference in particle size was observed throughout the doping range, as shown

in Fig. 3.1 b and Fig. 3.2. The typical particle sizes determined from AFM and DLS are

10.2±3.5 nm, 14.4±4.4 nm in diameter, respectively. As PCBM doping percentage

increases, the nanoparticle absorption at low wavelength (< 400 nm) increases, which

corresponds to PCBM absorption, while the absorption peak of PFBT remains unchanged

at 450 nm (Fig. 3.1 c). The presence of PCBM effectively quenched the fluorescence of

Page 42: Development of Conjugated Polymer Nanoparticles for

35

PFBT--at 5% PCBM doping ratio, the fluorescence quantum yield dropped by 70% as

compared to undoped PFBT CPNs (Fig. 3.1 d).

Figure 3.2. (a) AFM image and (b) size distribution of 5% PCBM doped PFBT nanoparticles

(12.9±5.8 nm); (c) AFM image and (d) size distribution of 10% PCBM doped PFBT nanoparticles

(10.7±3.9 nm); (e) AFM image and (f) size distribution of 20% PCBM doped PFBT nanoparticles

(9.9±4.2 nm); (g) AFM image and (h) size distribution of 40% PCBM doped PFBT nanoparticles

(10.2±4.5 nm).

Page 43: Development of Conjugated Polymer Nanoparticles for

36

3.1.2 Single Particle Fluorescence and Localization Study

The single particle fluorescence study of PCBM doped PFBT CPNs was carried out with

an inverted microscope. The CPNs were dispersed on a glass coverslip and excited with a

445 nm laser. Sequences of fluorescence microscopy images were acquired at multiple

framerates (1Hz, 10Hz or 50Hz) using a sCMOS camera, for over 600 s (fig. 3.3a). It is

found that the blinking behavior of PCBM doped PFBT CPNs is highly sensitive to the

dopant concentration. At low PCBM doping level (<10%), the fluorescence intensity of the

doped CPNs jumps between multiple levels, indicating fluctuations in the quencher

population, however, most of the nanoparticles do not turn “off” during the experiment.

When the PCBM percentage is higher (>20%), we typically observe transitions between a

dark state and multiple “on” states in the single particle fluorescence (fig. 3.3 c, d). Upon

further increase of the PCBM doping level (>40%), the frequency and duration of “on”

events is further reduced. The single particle fluorescence study was mainly focused on

20% and 40% PCBM doped PFBT CPNs because the frequency and duration of “on”

events in this doping range appeared promising for single particle localization. For some

CPNs, we observed a distinctive initial decay in fluorescence intensity followed by “on”

and “off” blinking behavior, which is likely caused by polaron population establishing an

equilibrium. A similar phenomenon was also observed for undoped CPNs, discussion and

modeling of the initial fluorescence decay dynamics arising from polaron generation was

discussed previously in our group.20

Page 44: Development of Conjugated Polymer Nanoparticles for

37

Figure 3.3. (a) A sequence of fluorescence microscopy images showing blinking behavior of 20% PCBM doped PFBT CPNs. (b) 3D histogram of the localized centroid position of a blinking 20% PCBM doped PFBT CPN, determined frame by frame from a trajectory acquired at 200W/cm2, 1 Hz framerate. The inserted plot shows the centroid position histogram along X axis of the same CPN, which is fitted to a Gaussian distribution (σ = 2.1 nm). (c, d) Fluorescence intensity trajectories of 20% PCBM doped PFBT CPNs acquired at 10 Hz framerate, two excitation powers, (c) 200 W/cm2, (d) 800 W/cm2, showing different duty cycles.

Since polaron generation is a photo-driven process and polaron population in CPNs is

likely to be excitation intensity dependent, we used various excitation intensities to

modulate the blinking behavior of 20% PCBM doped PFTB CPNs. For each laser intensity,

dozens of single particle blinking trajectories were analyzed. The results are summarized

in the table 3.1.

Page 45: Development of Conjugated Polymer Nanoparticles for

38

Table 3.1. Blinking parameters of 20% PCBM doped PFBT CPNs, determined at 50 Hz. Excitation Intensity

“On”/”off” contrast

“On” duration (s)

Duty cycle Photons detected per “on” event

Number of cycles before photobleach

200 W/cm2 10.1±4.3 8.2±5.9 0.28±0.10 5.2±2.8×104 22.6±8.8

400 W/cm2 13.9±6.2 4.9±3.5 0.16±0.07 3.8±2.0×104 19.9±6.9

800 W/cm2 15.3±6.5 3.4±1.6 0.05±0.04 3.5±2.2×104 20.7±7.4

1600 W/cm2 16.7±5.9 2.1±1.2 0.008±0.01 3.0±1.8×104 15.9±9.0

According to the table 3.1, the duty cycle of the blinking CPNs (fraction of time in the

“on” state) depends on excitation intensity. At higher excitation intensity, the increased

polaron generation rate and equilibrium population lead both to fewer “on” events per unit

time and reduced “on” state durations (Fig. 3.3 d). If we assume that the polaron population

fluctuations in CPNs follow Poisson statistics, the duty cycle 𝐷 can be written as a function

of the equilibrium polaron population in CPNs,

𝐷 = 1 −∑𝑛𝑒𝑞𝑖 𝑒−𝑛𝑒𝑞

𝑖!

𝑛𝑞

𝑖=0

,(3.1)

where 𝑛𝑞 is the number of polarons required to totally quench the fluorescence of a CPN

and 𝑛𝑒𝑞 is the equilibrium polaron population inside the CPN. Based on the previously

estimated polaron quenching efficiency (~10%) in undoped PFBT CPNs, we assume that

𝑛𝑞 = 10. According to the equation 3.1 and duty cycles determined in the table 3.1, we

estimated that the equilibrium polaron populations in 20% PCBM doped PFBT CPNs

under 200W/cm2, 400W/cm2, 800W/cm2, 1600W/cm2 excitation intensities to be 12, 14,

17, 21 respectively, assuming 𝑛𝑞 = 10. The larger polaron population at high excitation

Page 46: Development of Conjugated Polymer Nanoparticles for

39

intensities nearly completely suppresses the polymer fluorescence--only low probability

large fluctuations in the polaron population lead to occasional, rare fluorescence bursts.

Assuming the localization precision is shot noise limited, the theoretical localization

precision or uncertainty is given by 𝜎 = 𝑠 √𝑁⁄ , where s is the STD of the PSF and N is the

number of detected photons.70 For 20% PCBM doped PFTB CPNs, under low laser

excitation intensity (200W/cm2) and 1 Hz framerate, an average of 6.3×103 photons were

detected per frame during “on” state, yielding a theoretical per frame localization precision

of ~1.7 nm. The total number of photons detected per switching event was calculated by

integrating over all the frames during an “on” event. Under 200W/cm2, the average “on”

duration is ~8 s, an average of 5.2×104 photons were detected per switching event (1-2

orders of magnitude higher than dye molecules), resulting in an expected theoretical

localization precision of ~0.6 nm, about 4 times better than the typical resolution obtained

by localization of photoswitching dye molecules.90 The expected resolution is confirmed

by frame-by-frame centroid analysis of single burst events.

It should be noted that the polaron generation/recombination dynamics are highly

dependent on the local structure. At the polymer/PCBM interface, charge separation is

likely to be instantaneous whereas in the regions far away from the interface, charge

generation is much slower. The (proposed) mechanism of the polydot photoswitch is

described as follows. Upon excitation, reversible photoinduced charge transfer of electrons

from the polymer to the PCBM dopant quickly establishes a dynamically fluctuating

steady-state population of hole polarons on the polymer. The density of hole polarons is

Page 47: Development of Conjugated Polymer Nanoparticles for

40

sufficient that there are multiple overlapping quenching spheres over nearly the entire

nanoparticle volume, rendering the particle essentially nonfluorescent. We roughly

estimate that 10-30 polarons per particle is sufficient to achieve near total fluorescence

quenching. However, due to the reversible nature of the electron transfer, occasionally the

local population of hole polarons within a region of the particle drops below a threshold

amount for a period of time, such that a burst of fluorescence is emitted (additional details

and evidence supporting this mechanism, including model calculations, were reported).24

Additional experiments are required in order to obtain a complete physical picture of

charge generation/recombination processes in such a complex system.

Figure 3.4. Proposed photoswitching mechanism in PCBM-doped CPN photoswitches.

3.1.3 Superresolution Imaging of E. coli

As a proof of concept, photoswitching polydots were adsorbed on Escherichia coli (E.

coli). E. coli stocks were frozen in glycerol at -80 °C. Prior to each experiment, E. coli was

streaked on a LB-Kan-Amp plate, single colonies were selected for growth in low-

fluorescence media (1X Gibco M9 minimal media, 1% glucose, 100 ug/mL leucine, 25

Page 48: Development of Conjugated Polymer Nanoparticles for

41

mg/L ferric citrate, 50 ug/mL thiamine), supplemented with kanamycin and ampicillin,

overnight at 37 °C (approximately 16-20 hours).

Figure 3.5. (a) An AFM image of unlabeled E. coli, (b, c, d) Fluorescence microscopy images of

(b) unlabeled E. coli, (c) E. coli labeled with PFBT CPNs, (d) E. coli labeled with 40% PCBM

doped PFBT CPNs, imaged under 800W/cm2 excitation, 10 Hz framerates.

The bacteria were then dispersed on a glass coverslip and washed with phosphate-buffered

saline (PBS) three times. The bacteria fixation was carried out by submerging E. coli under

ice cold methanol for 10 min. The E. coli were washed with PBS buffer for three times

after fixation. The fixed bacteria were rinsed with a solution of a cationic surfactant, (1-

hexadecyl) trimethylammonium bromide (0.01M), to introduce positive charges to the

bacteria surface, then submerged under 200µL of nanoparticle suspension (~1 nM) for 30

Page 49: Development of Conjugated Polymer Nanoparticles for

42

mins and washed with PBS buffer afterwards. It was observed that PCBM doped PFBT

CPNs (with a zeta potential around -35mV) were efficiently adsorbed on the positively

charged bacteria surface (hundreds of CPNs per E. coli after 30 mins). In contrast, poor

labeling efficiency was observed for E. coli that had not been treated with the cationic

surfactant (0 to a few CPNs per E. coli after 30 mins), likely due to repulsion between

CPNs and the naturally negatively charged bacteria surfaces. The E. coli were imaged

under 800W/cm2 excitation, N2 atmosphere using 10 Hz framerates. E. coli labeled with

20% or 40% PCBM doped PFBT CPNs exhibited pronounced blinking (Fig. 3.6 a). While

the background autofluorescence from E. coli is typically dim as compared to the

brightness of 20% and 40% PCBM doped PFBT CPNs (Fig. 3.5 b, d), background

subtraction was performed to ensure optimal localization accuracy. The background

fluorescence was determined from frames with no CPN switched “on”. ~100 frames of the

background fluorescence were averaged and subtracted from each frame prior to analysis.

The position of each CPN was determined using the single particle localization method and

the positions obtained, were then used to reconstruct the shape of the E. coli, which is

shown by the scatter plot in Fig. 3.6 c. Analysis of the scatter plot shows that features well

below the diffraction limit were resolved. As shown in the Fig. 3.6 c subplot, the two

clusters correspond to the localized positions of two nearby CPNs, which are clearly

separated. The 33.6 nm distance between the centers of the clusters is highlighted with a

red dashed line. These results suggest that nanoparticles separated by ~30 nm are readily

resolved. From the standard deviation (𝜎𝑥 + 𝜎𝑦)/2 of the two clusters, we determined the

per frame localization uncertainty of the two CPNs to be 5.95 nm (left) and 6.82 nm (right),

Page 50: Development of Conjugated Polymer Nanoparticles for

43

which are consistent with the brightness of 20% PCBM doped PFTB CPNs under the

imaging conditions.

Figure 3.6. (a) A sequence of fluorescence microscopy images showing the blinking behavior

of an E. coli labeled with 20% PCBM doped PFBT CPNs. (b) A 3D fluorescence intensity map

shows a few switched “on” CPNs on an E. coli, which are indicated by the red arrows. The

inserted plot shows a 2D view of the same image with localized positions of switched “on”

CPNs indicated by red crosses. (c) A scatter plot shows the superresolution image of an E. coli

constructed from the localized positions of blinking CPNs. The subplot shows the zoomed in

view of two clusters in the scatter plot, indicating CPNs separated by ~30 nm are clearly

resolved.

Page 51: Development of Conjugated Polymer Nanoparticles for

44

3.2 Attempts to improve the photoswitching CPNs: PCBM Doped MEH-

CNPPV Nanoparticles

Given the exceptional brightness and photostability of CPNs,19-22 there has been

considerable interest in developing photoswitchable nanoparticle based on CPNs.23, 86 The

unparalleled brightness of photoswitching polydots and their flexible optical and other

properties could yield improvements in demanding super resolution imaging applications

such as 2D or 3D super resolution localization-based microscopy of complex samples. The

higher brightness of photoswitching polydots should help overcome autofluorescence and

scattering in complex 3D samples that can reduce signal to noise ratio. However, designing

photoswitchable CPNs have been proven to be challenging. Previous efforts in modifying

CPNs for superresolution imaging typically involved incorporating photochromic dyes into

CPNs.

In section 3.1 of this document, we demonstrated that PCBM doped PFBT nanoparticles

exhibit spontaneous photoswitching behavior with unprecedented brightness. But there is

still a lot to be improved for these nanoparticles, to further develop the photoswitching

polydots and enable application to a broad range of imaging applications. In particular, we

wanted to achieve a more even distribution of PCBM in CPNs to reduce considerable

particle-to-particle variations in brightness, duty cycles, etc.

In this direction aggregation studies were performed with nanoparticles made of

different conjugated polymers including CNPPV and MEH-CNPPV, the DLS and AFM

data showed that MEH-CNPPV suspensions have better colloidal stability than PFBT,

Page 52: Development of Conjugated Polymer Nanoparticles for

45

yielding a narrower range of the nanoparticle sizes. Also, we already have stablished that

the addition of PCBM did not change the size distribution for PCBM doped PFBT

nanoparticles, so it was decided to make PCBM doped MEH-CNPPV polydots to study

their photoswitching behavior and wether or not we could a narrower size distribution of

the photoswitching polydots.

3.2.1 Nanoparticle Preparation and Characterization

Nanoparticles of the conjugated polymer MEH-CNPPV doped with various percentages

of PCBM were prepared using the nano-precipitation method described previously.71, 74, 91

20 ppm solutions of MEH-CNPPV and PCBM in THF were prepared and filtered through

a 0.45 μm membrane filter (Millipore). The solutions of MEH-CNPPV and PCBM were

mixed in various ratios to form precursors of varying dopant percentages (1%, 3%, 10%),

2 mL of the precursor solution was rapidly injected into 8 mL of water under mild

sonication. Then the THF solvent was removed by partial vacuum evaporation. After

evaporation, the sample was filtered through a 100 nm membrane filter (Millipore) to

remove aggregates. The size distribution of the nanoparticles was determined using AFM

and DLS. No significant difference in particle size was observed throughout the doping

range, as shown in Fig. 3.7. The typical particle sizes determined from AFM and DLS are

11.1±1.5 nm, 13.1±2.2 nm in diameter, respectively.

Page 53: Development of Conjugated Polymer Nanoparticles for

46

Figure 3.7. (left) Number weighted particle size distributions of PCBM doped MEH-CNPPV CPNs at various PCBM doping percentages, determined from DLS measurements. (right) Normalized fluorescence spectra (excited at 470 nm) of PCBM doped PFBT CPNs at various PCBM doping percentages.

It is worth noting that the fluorescence was almost completely quenched at 10% PCBM

doping ratio (Figure 3.7), the fluorescence quantum yield dropped by 95% as compared to

undoped MEH-CNPPV CPNs, so the doping ratios were adjusted in comparison to the (5-

40% doping ratio used in the PCBM doped PFBT experiment).

To further analyze the data a Stern-Volmer plot was made plotting F0/F vs [Q]

(molecular ratio between the quencher and the polymer in moles).92 The equation that

represents the Stern-Volmer linear relationship was determined by least squares linear

fitting and is giving by:

𝐹0𝐹= 1 + 5.532[𝑄],𝑅2 = 0.9957(3.2)

The value 5.532 is better known as KD, and 1/ KD (0.18), represents the molecular

fraction of the quencher needed to quench the initial fluorescence by 50%.

Page 54: Development of Conjugated Polymer Nanoparticles for

47

3.2.2 Single Particle Fluorescence

The single particle fluorescence study of PCBM doped MEH-CNPPV CPNs was carried

out with an Olympus IX71 inverted microscope with a 100X 1.3 NA fluorescence

objective. An illumination laser is fiber coupled to the microscope for epi illumination. For

some experiments, a manual XY micrometer stage (Thorlabs) is used to manipulate the

sample. The CPNs were dispersed on a glass coverslip and excited with a 445 nm laser.

Sequences of fluorescence microscopy images were acquired at different framerates (10Hz

and 20Hz) using a sCMOS camera, for around 180s.

Figure 3.8. Fluorescence microscopy images of (a) 1% PCBM doped MEH-CNPPV CPNs (b) 10% PCBM doped MEH-CNPPV CPNs, images acquired at 400 W/cm2, 10 Hz framerate.

a b

Page 55: Development of Conjugated Polymer Nanoparticles for

48

Figure 3.9. (a) The trajectory of a MEH-CNPPV CPN, showing characteristic exponential photobleaching, acquired at 400 W/cm2, 20 Hz framerate. (b) The trajectory of a 1% PCBM doped MEH-CNPPV CPN, showing some blinking behavior acquired at 200 W/cm2, 10 Hz framerate. (c) The trajectory of a 1% PCBM doped MEH-CNPPV CPN, showing initial blinking behavior followed by photobleaching, acquired at 400 W/cm2, 20 Hz framerate. (d) The trajectory of a 10% PCBM doped MEH-CNPPV CPN, showing initial blinking behavior followed by photobleaching, acquired at 200 W/cm2, 10 Hz framerate.

In figure 3.9 (a) we can see the florescence decay curve for an undoped MEH-CNPPV

nanoparticle, in contrast when a percentage of the dopant is added to the nanoparticle (Fig,

3.9 b, c, d) the photobleaching behavior is modified and does not follow single or double

exponential models, and some blinking behavior is present, analogous the PCBM doped

PFBT nanoparticle already described. It is worth noting that at low PCBM doping level

(1%) most of the nanoparticles exhibit fluctuations, however, do not completely turn “off”

Page 56: Development of Conjugated Polymer Nanoparticles for

49

during the experiment. When the PCBM percentage is higher (10%) we observed

transitions between a dark state and multiple “on” states in the single particle fluorescence.

The death number was calculated by using the equation:

𝑁 =𝑇𝑜𝑡𝑎𝑙𝑐𝑜𝑢𝑛𝑡𝑠 ∗ 𝑔𝑎𝑖𝑛

𝑄𝑢𝑎𝑛𝑡𝑢𝑚𝑒𝑓𝑓.∗ 𝑐𝑜𝑙𝑙𝑒𝑐𝑡𝑖𝑜𝑛𝑒𝑓𝑓(3.3)

The gain factor used in the calculations was 0.6 electrons/count, the quantum efficiency

of the camera is 0.57 electrons/photon and the collection efficiency 0.03. For undoped

MEH-CNPPV nanoparticles the average death number was 7.8x107 photons and the

average intensity over the first 20 frames (1s) was 6.7x103 counts, the single exponential

time constant determined from the exponential fitting was 14.3 s in average for a collection

of CPNs. For 1% PCBM doped MEH-CNPPV CPNs the death number ranged from

1.6x107 for lower intensity to 2.8x107 photons in higher intensity nanoparticles and the

average intensity over the first 20 frames (1s) was 1.1x103 counts. In the case of 10%

PCBM doped MEH-CNPPV CPNs the death number ranged from 1.3x107 for lower

intensity to 2.2x107 photons in higher intensity nanoparticles and the average intensity over

the first 20 frames (1s) was 500 counts.

In summary, we discovered blinking behavior for the PCBM doped MEH-CNPPV

CPNs which poses these nanoparticles as a candidate for a successful photoswitching

polydot with promising potential in superresolution imaging, given the narrower size

distributions and the higher colloidal stability in comparison with PCBM doped PFBT

CPNs, also the death numbers data show that the PCBM doped MEH-CNPPV

Page 57: Development of Conjugated Polymer Nanoparticles for

50

nanoparticles are brighter in the “on” state than conventional dyes used for superresolution

imaging. However, the single particle fluorescence data obtained for the PCBM doped

MEH-CNPPV photoswitches was very complex and goes beyond the scope of this thesis,

for future work it is evident and necessary to improve the current methods to analyze these

sets of complex data, developing better scripts for analysis and apply segmented analysis

of the nanoparticle trajectories to achieve better understanding of how the position

trajectories varied over time. It is also important to approach this data analysis issue from

simulation analysis of the polaron dynamics, understanding the autocorrelation data and

the polaron recombination and generation rates,87 respectively.

3.3 Electric Field Modulation of CPN’s fluorescence.

Based on the demonstration that PCBM doped PFBT nanoparticles exhibit spontaneous

photoswitching behavior with unprecedented brightness, other novel ideas were considered

to explore new possible applications including electrical switching for sensing membranes

potential or even the possibility of attaching nanoparticles to neurons and being able to

report when a neuron “fires”, thus changing the membrane potential.93

There is some information published about electric fields and conjugated polymers

fluorescence, 45, 94, 95 however, there is very limited information about electric field

modulation of the photoluminescence in conjugated polymer nanoparticles. Therefore,

based on our novel fluorescent photoswitch idea, we decided to test the response of our

nanoparticles to electric fields in the single molecule level, starting with a simple use of a

voltage trying to induce photoswitching of PFBT nanoparticles.

Page 58: Development of Conjugated Polymer Nanoparticles for

51

The PFBT nanoparticles for this study were prepared used the same nano-precipitation

method described previously in sections 2.2 and 3.1.1.71, 74, 75 The size distribution of the

nanoparticles was determined using AFM and DLS. The typical PFBT particle sizes

determined from AFM and DLS are 14.4±2.8 nm, 16.6±3.6 nm in diameter, respectively.

In this work, the samples were prepared on Indium tin oxide (ITO) coated glass

coverslips as an electrode attached to a copper wire using silver paint. Those coverslips

were cleaned with the following steps: soap water, DI water, and plasma cleaning. Before

the addition of nanoparticles, a dielectric material polystyrene (1.5% in Toluene) was spin-

coated on top of the ITO coverslip, followed by deposition of nanoparticles on top of the

PS film by drop casting, using 200 times dilution from the nanoparticles vial previously

prepared (~0.16ppm in polymer). It is worth noting that the experiments were performed

with and without dielectric (PS), and better results were observed with PS coated film.

Once the nanoparticles were deposited on the coverslip and placed on top of the

inverted microscope objective, a drop (~100 uL) of 0.01 M NaCl solution was placed in

the middle of the coverslip and a platinum wire was held inside the electrolyte solution.

These experiments were performed initially using DI water only, but there was no electric

field modulation observed. After adding the electrolyte, the system was connected to a

RIGOL wave form generator, and an oscilloscope in parallel to measure the voltage. Square

wave forms were used in most of the experiments, with different frequencies and

amplitudes of the electric field. The best results were obtained by using a -2V – 0V square

wave and pulses of 1-5 s of duration.

Page 59: Development of Conjugated Polymer Nanoparticles for

52

The single particle fluorescence study of PCBM CPNs was carried out with an inverted

microscope and the excitation laser at 445 nm. Sequences of fluorescence microscopy

images were acquired at 10Hz using a sCMOS camera, for over 120 s under a 100 W/cm2

excitation.

Figure 3.10. (a) Fluorescence intensity trajectories of PCBM CPNs modulated by -2V-0V

square wave pulses acquired at 10 Hz framerate, and excitation power of 100 W/cm2. (b) Data

zoomed for the first 30 s, showing the 1s square pulses on top of the fluorescence intensity

graph.

Page 60: Development of Conjugated Polymer Nanoparticles for

53

The results from the experiments shown in figure 3.10 show that there is an evident

effect on the photoluminescence of the PFBT nanoparticles under the influence of an

electric field, the intensity of the fluorescence goes up and down when turning the electric

field on instead of following the regular photobleaching exponential decay observed when

there is no external electric field applied in the system. These preliminar results are very

promising and lead us to believe that new possible applications including electrical

switching for sensing membranes potential or even the possibility of attaching

nanoparticles to neurons is possible, however it is necessary to pursue more elaborate

experiments to understand the mechanisms of charge transport in order to be able to

develop nanoparticles or photoswitches that can be sensible to electric fields at the cell

level.

3.4 DNA wrapped polydots development.

We recently demonstrated the first polydots or CPNs with excellent switching

characteristics for localization microscopy.24 Continuing with the overall research plan to

further develop the photoswitching polydots to enable application to a broad range of

imaging applications, we decided to pursue novel approaches for polydot sensors based on

DNA-wrapped particles. Based on reports of carbon SWNT wrapped with DNA and RNA

prepared by sonication,96 it was decided to attempt this with polydots to produce DNA-

wrapped polydots (i.e., DNA adsorbed on the particle surface). The initial efforts of this

initiative are shown in this thesis, focused on development of protocols for wrapping

polydots in DNA oligomers, and the discovery that the hybrid DNA polydots have the

potential to act as sensors for neurotransmitters such as dopamine.

Page 61: Development of Conjugated Polymer Nanoparticles for

54

Nanoparticles of MEH-CNPPV were prepared and characterized using the procedures

previously described on section 3.2.1 of this document. It is worth noting that MEH-

CNPPV was selected for this study due to its better colloidal stability as shown in previous

results. The typical MEH-CNPPV particle sizes determined from AFM and DLS are

11.1±1.5 nm, 13.1±2.2 nm in diameter, respectively. Several buffers were made to test the

best option to keep the pH above 7.4 and at the same time to have a low degree of

aggregation for the MEH-CNPPV nanoparticles. The best option was selected to be TE

(Tris-EDTA) 1mM buffer 7.8 pH. All dilutions were performed using this buffer.

The MEH-CNPPV suspension was diluted to a final concentration of 1.5 ppm of

polymer. This concentration was determined by absorbance measurements, where 67% of

the material was recovered after sample preparation and 93% after filtration, yielding a

total of 63% recovery (~37% loss). Two DNA oligomers samples PGEX3 and ss(GA)15

(purchased from Integrated DNA Technologies) were diluted to a 0.6 ppm concentration

and mixed under sonication (4 minutes) with the MEH-CNPPV nanoparticle suspension.

At this point absorbance and fluorescence spectra were taken and no changes were

observed due to the presence of DNA. Finally, dopamine was added to the mixture to a

final concentration of 20uM dopamine, 0.6 ppm MEHCNPV and 0.24 ppm DNA, no

changes in the absorbance were observed and the fluorescence data is shown as follow in

figure 3.11.

Page 62: Development of Conjugated Polymer Nanoparticles for

55

Figure 3.11. Normalized fluorescence spectra for the samples and the control experiment using

an excitation wavelength of 445 nm.

It is worth nothing that the addition of dopamine to the nanoparticles suspension does

not affect the fluorescence, but when mixed with DNA the fluorescence peak is enhanced

by a factor of 6.3 for the ss(GA)15 DNA and 4.8 for PGEX3’ commercial sequence

respectively as shown in figure 3.11. This experiment result is very interesting as we

demonstrated that the addition of dopamine to DNA wrapped MEH-CNPPV nanoparticles

enhances the fluorescence significantly and opens the possibility of doing superresolution

imaging of cellular dopamine efflux and to resolve where on the cell surface dopamine is

released and how cell morphology affects the location of release sites,97 among other

possible applications.

Page 63: Development of Conjugated Polymer Nanoparticles for

56

CHAPTER FOUR

GENERAL CONCLUSIONS AND FUTURE DIRECTIONS

4.1 Conclusions and future directions

In summary, we have developed a new class of fluorescent nanoparticles for

superresolution imaging with significantly improved spatial resolution. PCBM doped

PFBT CPNs establish large polaron populations inside the nanoparticles that nearly

completely suppress the polymer fluorescence. However, fluctuations in the number of

charge carriers lead to occasional bursts of fluorescence. For 20% PCBM doped PFBT

CPNs, we determined that around 3-5×104 photons were detected during the fluorescence

burst, which results in a localization precision of ~0.6 nm, about 4 times better than typical

resolution obtained by localization of dye molecules.

Since polaron generation is a photo-driven process, we showed that the blinking duty

cycle of PCBM doped PFBT CPNs can be controlled by excitation intensity as well as by

PCBM fraction. Finally, we demonstrated superresolution microscopy of E. coli using 20%

PCBM doped PFBT CPNs. At 10 Hz framerates, we obtained the precise shape of the

bacteria with a localization precision of ~5 nm. These results suggest that PCBM doped

PFBT CPNs represent a novel class of promising superresolution probes with

unprecedented localization precision and tunable spontaneous photoswitching (blinking)

properties, which provide clear advantages for imaging of various biological systems.

Page 64: Development of Conjugated Polymer Nanoparticles for

57

Other efforts to improve these photoswitches were attempted and it was encountered

that PCBM doped MEH-CNPPV nanoparticles also exhibit spontaneous blinking, with the

advantage of having a better colloidal stability and narrower distribution sizes, which is

important for better measurements on biological systems. For future directions in order to

improve this PCBM doped MEH-CNPPV photoswitch it is necessary to perform more

single particle fluorescence measurements to probe the relevant photochemistry and

photophysics. Also blinking kinetics experiments as a function of excitation intensity,

determining the effect of various particle compositions and surface functionalities, and

blending and doping. This will help to understand how the various components are

interacting, and whether undesired processes are occurring that could interfere with the

photoswitching process. It is important to analyze the data to treat the nanoparticles as a

diverse population, collecting and interpreting results representing the overall distribution

of particle properties and particle-to-particle variability.

Finally, we developed protocols for wrapping polydots in DNA oligomers, and

discovered that the hybrid DNA polydots can act as sensors for neurotransmitters such as

dopamine, resulting on a very promising, novel approach to chemical sensing in polydots,

which could yield a broad range of useful sensors for chemical microscopy. We found that

the resulting nanoparticles exhibited a roughly six-fold enhancement in fluorescence in the

presence of biologically relevant concentrations of a neurotransmitter. This could provide

the basis for a “turn-on” sensor for neurotransmitters that can be used in imaging of neuron

dynamics. Additionally, we can combine in future work the sensing with photoswitching,

for super resolution chemical microscopy.

Page 65: Development of Conjugated Polymer Nanoparticles for

58

REFERENCES

(1) Heeger, A. J., Nobel Lecture: Semiconducting and Metallic Polymers: the Fourth Generation of Polymeric Materials. Reviews of Modern Physics 2001, 73, 681. (2) Huang, B.; Bates, M.; Zhuang, X. W. Super-Resolution Fluorescence Microscopy. Annual Review of Biochemistry 2009, 78, 993-1016. (3) Teshima, Y.; Saito, M.; Mikie, T.; Komeyama, K. Osaka. I., Bithiazole Dicarboxylate Ester: An Easily Accessible Electron-Deficient Building Unit for π-Conjugated Polymers Enabling Electron Transport. Macromolecules 2021, 54 (7), 3489-3497. (4) Mitschke, U.; Bauerle, P., The electroluminescence of organic materials. Journal of Materials Chemistry 2000, 10, 1471-1507. (5) Bagwe, R. P.; Yang, C. Y.; Hilliard, L. R.; Tan, W. H. Optimization of dye-doped silica nanoparticles prepared using a reverse microemulsion method. Langmuir 2004, 20, 8336-8342. (6) Barbara, P. F.; Gesquiere, A. J.; Park, S. J.; Lee, Y. J. Single-molecule spectroscopy of conjugated polymers. Accounts of Chemical Research 2005, 38, 602-610. (7) Hide, F.; DiazGarcia, M. A.; Schwartz, B. J.; Heeger, A. J. New developments in the photonic applications of conjugated polymers. Accounts of Chemical Research 1997, 30, 430-436. (8) Gunes, S.; Neugebauer, H.; Sariciftci, N. S. Conjugated polymer-based organic solar cells. Chemical Reviews 2007, 107, 1324-1338. (9) Burroughes, J. H.; Bradley, D. D. C.; Brown, A. R.; Marks, R. N.; Mackay, K.; Friend, R. H.; Burn, P. L.; Holmes, A. B. Light-Emitting-Diodes Based on Conjugated Polymers. Nature 1990, 347, 539-541. (10) Samuel, I. D. W.; Rumbles, G.; Collison, C. J.; Friend, R. H.; Moratti, S. C.; Holmes, A. B. Picosecond time-resolved photoluminescence of PPV derivatives. Synthetic Metals 1997, 84, 497-500. (11) Bunz, U. H. F., Poly(aryleneethynylene)s: Syntheses, Properties, Structures, and Applications. Chemical Reviews 2000, 1000, 1605-1644. (12) Kloppenburg, L.; Jones, D.; Bunz, U. H. F., High Molecular Weight Poly(p-phenyleneethynylene)s by Alkyne Metathesis Utilizing "Instant" Catalysts: A Synthetic Study. Macromolecules 1999, 32, (13), 4194-4203. (13) Mullen, K.; Scherf, U., Novel Conjugated Polymers. Synthetic Metals 1993, 55, (2-3), 739-752. (14) Heeger, A. J.; Kivelson, S.; Schrieffer, J. R.; Su, W. P., Solitons in Conducting Polymers. Reviews of Modern Physics 1988, 60, (3), 781-850. (15) Su, W. P.; Schrieffer, J. R.; Heeger, A. J., Solitons in Polyacetylene. Physical Review Letters 1979, 42, (25), 1698-1701.

Page 66: Development of Conjugated Polymer Nanoparticles for

59

(16) Schwartz, B. J., Conjugated Polymers as Molecular Materials. Annu. Rev. Phys. Chem. 2003, 54, 141. (17) Wu, C. F.; Chiu, D. T. Highly Fluorescent Semiconducting Polymer Dots for Biology and Medicine. Angewandte Chemie-International Edition 2013, 52, 3086-3109. (18) Ye, F.; Sun, W.; Zhang, Y.; Wu, C.; Zhang, X.; Yu, J.; Rong, Y.; Zhang, M.; Chiu, D. T., Single-Chain Semiconducting Polymer Dots. Langmuir 2014, 31 (1), 499-505. (19) Tian, Z.; Yu, J.; Wu, C.; Szymanski, C.; McNeill, J., Amplified Energy Transfer in Conjugated Polymer Nanoparticle Tags and Sensors. Nanoscale 2010, 2 (10), 1999-2011. (20) Jiang, Y.; McNeill, J., Light-harvesting and Amplified Energy Transfer in Conjugated Polymer Nanoparticles. Chemical Reviews 2017, 117 (2), 838–859. (21) Wu, C. F.; Szymanski, C.; Cain, Z.; McNeill, J. D. Conjugated Polymer Dots for Multiphoton Fluorescence Imaging. Journal of the American Chemical Society 2007, 129, 12904-12905. (22) Wu, C. F.; Bull, B.; Christensen, K.; McNeill, J. Ratiometric Single-Nanoparticle Oxygen Sensors for Biological Imaging. Angewandte Chemie-International Edition 2009, 48, 2741-2745. (23) Harbron, E. J.; Davis, C. M.; Campbell, J. K.; Allred, R. M.; Kovary, M. T.; Economou, N. J. Photochromic Dye-Doped Conjugated Polymer Nanoparticles: Photomodulated Emission and Nanoenvironmental Characterization. Journal of Physical Chemistry C 2009, 113, 13707-13714. (24) Jiang, Y.; Novoa, M.; Nongnual, T.; Powell, R.; Bruce, T. F.; McNeill, J. D. Improved Superresolution Imaging Using Telegraph Noise in Organic Semiconductor Nanoparticles. Nano Letters 2017, 17 (6), 3896–3901. (25) Yu, J. B.; Wu, C. F.; Sahu, S. P.; Fernando, L. P.; Szymanski, C.; McNeill, J. D. Nanoscale 3D Tracking with Conjugated Polymer Nanoparticles. Journal of the American Chemical Society 2009, 131, 18410-18414. (26) Grimland, J. L.; Wu, C. F.; Ramoutar, R. R.; Brumaghim, J. L.; McNeill, J.

Photosensitizer-doped conjugated polymer nanoparticles with high cross-sections for one-

and two-photon excitation. Nanoscale 2011, 3, 1451-1455.

(27) Betzig, E.; Patterson, G. H.; Sougrat, R.; Lindwasser, O. W.; Olenych, S.; Bonifacino,

J. S.; Davidson, M. W.; Lippincott-Schwartz, J.; Hess, H. F., Imaging Intracellular

Fluorescent Proteins at Nanometer Resolution. Science 2006, 313 (5793), 1642-1645.

(28) Rust, M. J.; Bates, M.; Zhuang, X., Sub-Diffraction-Limit Imaging by Stochastic

Optical Reconstruction Microscopy (STORM). Nature Methods 2006, 3 (10), 793-796.

(29) Wu, C.; McNeill, J., Swelling-Controlled Polymer Phase and Fluorescence Properties

of Polyfluorene Nanoparticles. Langmuir 2008, 24 (11), 5855-5861.

(30) Hu, Z.; Tenery, D.; Bonner, M. S.; Gesquiere, A. J., Correlation between

Spectroscopic and Morphological Properties of Composite P3HT/PCBM Nanoparticles

Studied by Single Particle Spectroscopy. Journal of Luminiscence 2010, 130 (5), 771-780.

(31) Labastide, J. A.; Baghgar, M.; Dujovne, I.; Venkatraman, B. H.; Ramsdell, D. C.;

Venkataraman, D.; Barnes, M. D., Time-and Polarization-Resolved Photoluminescence of

Page 67: Development of Conjugated Polymer Nanoparticles for

60

Individual Semicrystalline Polythiophene (P3HT) Nanoparticles. Journal of Physical

Chemistry Letters 2011, 2 (17), 2089-2093.

(32) Yang, Z.; Huck, W. T.; Clarke, S. M.; Tajbakhsh, A. R.; Terentjev, E. M., Shape-Memory Nanoparticles from Inherently Non-Spherical Polymer Colloids. Nature Materials 2005, 4 (6), 486-490. (33) Kasha, M.; Rawls, H. R.; El-Bayoumi, M. A. The Exciton Model in Molecular Spectroscopy. Pure and Applied Chemistry 1965, 11, 371-392. (34) Bardeen, C. J. The Structure and Dynamics of Molecular Excitons. Annual Review of Physical Chemistry, Vol 65 2014, 65, 127-148. (35) Haugeneder, A.; Neges, M.; Kallinger, C.; Spirkl, W.; Lemmer, U.; Feldmann, J.; Scherf, U.; Harth, E.; Gugel, A.; Mullen, K. Exciton diffusion and dissociation in conjugated polymer fullerene blends and heterostructures. Physical Review B 1999, 59, 15346-15351. (36) Chen, J. W.; Cao, Y. Development of Novel Conjugated Donor Polymers for High-Efficiency Bulk-Heterojunction Photovoltaic Devices. Accounts of Chemical Research 2009, 42, 1709-1718. (37) Shaw, P. E.; Ruseckas, A.; Samuel, I. D. W. Exciton diffusion measurements in poly(3-hexylthiophene). Advanced Materials 2008, 20, 3516. (38) Hennebicq, E.; Pourtois, G.; Scholes, G. D.; Herz, L. M.; Russell, D. M.; Silva, C.; Setayesh, S.; Grimsdale, A. C.; Mullen, K.; Bredas, J. L.; Beljonne, D. Exciton Migration in Rigid-Rod Conjugated Polymers: An Improved Forster Model. Journal of the American Chemical Society 2005, 127, 4744-4762. (39) Dias, F. B.; Knaapila, M.; Monkman, A. P.; Burrows, H. D. Fast and slow time regimes of fluorescence quenching in conjugated polyfluorene-fluorenone random copolymers: The role of exciton hopping and dexter transfer along the polymer backbone. Macromolecules 2006, 39, 1598-1606. (40) Yu, G.; Heeger, A. J. Charge Separation and Photovoltaic Conversion in Polymer Composites with Internal Donor-Acceptor Heterojunctions. Journal of Applied Physics 1995, 78, 4510-4515. (41) Hooley, E. N.; Tilley, A. J.; White, J. M.; Ghiggino, K. P.; Bell, T. D. M. Energy transfer in PPV-based conjugated polymers: a defocused widefield fluorescence microscopy study. Physical Chemistry Chemical Physics 2014, 16, 7108-7114. (42) Chen, R. Apparent Stretched-Exponential Luminescence Decay in Crystalline Solids. Journal of Luminescence 2003, 102, 510-518. (43) Yu, J.; Song, N. W.; McNeill, J. D.; Barbara, P. F., Efficient exciton quenching by hole polarons in the conjugated polymer MEH-PPV. Israel Journal of Chemistry 2004, 44, (1-3), 127-132. (44) Salem, L., The Molecular Orbital Theory of Conjugated Systems. W. A. Benjamin, Inc.: New York, 1966. (45) McNeill, J. D.; O'Connor, D. B.; Adams, D. M.; Kämmer, S. B.; Barbara, P. F., Field-induced Photoluminescence Modulation of MEH-PPV under Near Field Optical Excitation. Journal of Physical Chemistry B 2001, 105, 76.

Page 68: Development of Conjugated Polymer Nanoparticles for

61

(46) Deussen, M.; Scheidler, M.; Bassler, H., Electric Field Induced Photoluminescence Quenching in Thin Film Light Emitting Diodes Based on poly(phenyl-p-phenylene vinylene). Synthethic Metals 1995, 73, 123. (47) Arkhipov, V. I.; Bassler, H.; Deussen, M.; Gobel, E. O.; Lemmer, U.; Mahrt, R. F. Field-induced dissociation of optical excitations in conjugated polymers. Journal of Non-Crystalline Solids 1996, 198, 661-664. (48) McNeill, J. D.; Barbara, P. F. NSOM investigation of carrier generation, recombination, and drift in a conjugated polymer. Journal of Physical Chemistry B 2002, 106, 4632-4639. (49) Yu, J. B.; Wu, C. F.; Tian, Z. Y.; McNeill, J. Tracking of Single Charge Carriers in a Conjugated Polymer Nanoparticle. Nano Letters 2012, 12, 1300-1306. 28. (50) Lakowicz, J. R., Principles of Fluorescence Spectroscopy. 3rd ed.; Springer Press+Business Media, LLC: New York, 2006. (51) Magde, D.; Wong, R.; Seybold, P. G. Fluorescence quantum yields and their relation to lifetimes of rhodamine 6G and fluorescein in nine solvents: Improved absolute standards for quantum yields. Photochemistry and Photobiology 2002, 75, 327-334. (52) Croce, R.; Muller, M. G.; Caffarri, S.; Bassi, R.; Holzwarth, A. R., Energy transfer pathways in the minor antenna complex CP29 of photosystem II: A femtosecond study of carotenoid to chlorophyll transfer on mutant and WT complexes. Biophysical Journal 2003, 84, (4), 2517-2532. (53) Bokinsky, G.; Rueda, D.; Misra, V. K.; Rhodes, M. M.; Gordus, A.; Babcock, H. P.; Walter, N. G.; Zhuang, X. W., Single-molecule transition-state analysis of RNA folding. Proceedings of the National Academy of Sciences of the United States of America 2003, 100, (16), 9302-9307. (54) Chen, J.; Shi, C.; Kang, X.; Shen, X.; Lao, X.; Zheng, H. Recent advances in fluorescence resonance energy transfer-based probes in nucleic acid diagnosis. Analytical Methods, 2020, 12, 884-893. (55) Stryer, L.; Haugland, R. P., Energy Transfer: A Spectroscopic Ruler. Proc. Natl. Acad. Sci. USA 1967, 58, 719. (56) Ambrose, W. P.; Moerner, W. E., Fluorescence Spectroscropy and Spectral Diffusion of Single Impurity Molecules in a Crystal. Nature 1991, 349, 225. (57) Betzig, E.; Grubb, S. G.; Chichester, R. J.; Digiovanni, D. J.; Weiner, J. S., Fiber Laser Probe For Near-Field Scanning Optical Microscopy. Applied Physics Letter 1993, 63, (26), 3550-3552. (58) Thiele, J. C.; Helmerich D. A.; Oleksiievets, N.; Tsukanov, R.; Butkevich, E.; Sauer, M., Confocal Fluorescence-Lifetime Single-Molecule Localization Microscopy. ACS Nano 2020, 14, (10), 14190–14200. (59) Oleksiievets, N.; Thiele, J. C.; Weber, A.; Gregor, I.; Nevskyi, O.; Isbaner, S.; Tsukanov, R.; Enderlein. J., Wide-Field Fluorescence Lifetime Imaging of Single Molecules, The Journal of Physical Chemistry A 2020, 124, (17), 3494-3500. (60) Plakhotnik, T.; Donley, E. A.; Wild, U. P. Single-molecule spectroscopy. Annual Review of Physical Chemistry 1997, 48, 181-212. (61) Moerner, W. E. A dozen years of single-molecule spectroscopy in physics, chemistry, and biophysics. Journal of Physical Chemistry B 2002, 106, 910-927.

Page 69: Development of Conjugated Polymer Nanoparticles for

62

(62) Hu, D. H.; Yu, J.; Barbara, P. F. Single-molecule spectroscopy of the conjugated polymer MEH-PPV. Journal of the American Chemical Society 1999, 121, 6936-6937. (63) VandenBout, D. A.; Yip, W. T.; Hu, D. H.; Fu, D. K.; Swager, T. M.; Barbara, P. F. Discrete intensity jumps and intramolecular electronic energy transfer in the spectroscopy of single conjugated polymer molecules. Science 1997, 277, 1074-1077. (64) Jiang, Y.; McNeill. J., Superresolution mapping of energy landscape for single charge carriers in plastic semiconductors. Nature Communications 2018, 9, (1), 4314. (65) Jiang, Y.; Hu, Q.; Chen, H.; Zhang, J.; Chiu, D.; McNeill, J., Dual‐Mode Superresolution Imaging Using Charge Transfer Dynamics in Semiconducting Polymer Dots. Angewandte Chemie 2020, 132, (37), 16307-16314. (66) Yildiz, A.; Tomishige, M.; Vale, R. D.; Selvin, P. R. Kinesin walks hand-over-hand. Science 2004, 303, 676-678. (67) Schuler, B.; Lipman, E. A.; Eaton, W. A. Probing the free-energy surface for protein folding with single-molecule fluorescence spectroscopy. Nature 2002, 419, 743-747. (68) Lin, H. Z.; Camacho, R.; Tian, Y. X.; Kaiser, T. E.; Wurthner, F.; Scheblykin, I. G. Collective Fluorescence Blinking in Linear J-Aggregates Assisted by Long-Distance Exciton Migration. Nano Letters 2010, 10, 620-626. (69) E. Abbe Beiträge zur theorie des mikroskops und der mikroskopischen wahrnehmung. Archiv für mikroskopische Anatomie 1873, 9, 413-418. (70) Thompson, R. E.; Larson, D. R.; Webb, W. W. Precise nanometer localization analysis for individual fluorescent probes. Biophysical Journal 2002, 82, 2775-2783. (71) Groff, L. D.; Wang, X.; McNeill, J. D., Measurement of Exciton Transport in

Conjugated Polymer Nanoparticles. J. Phys. Chem. C 2013, 117, (48), 25748-25755.

(72) Jiang, Y.; Nongnual, T.; Groff, L.; McNeill J., Nanoscopy of Single Charge Carrier

Jumps in a Conjugated Polymer Nanoparticle. The Journal of Physical Chemistry C 2018,

122, (2) , 1376-1383.

(73) Wu, C. F.; Peng, H. S.; Jiang, Y. F.; McNeill, J. Energy transfer mediated fluorescence from blended conjugated polymer nanoparticles. The Journal of Physical Chemistry B 2006, 110, 14148-14154. (74) Wu, C. F.; Zheng, Y. L.; Szymanski, C.; McNeill, J. Energy transfer in a nanoscale multichromophoric system: Fluorescent dye-doped conjugated polymer nanoparticles. The Journal of Physical Chemistry C 2008, 112, 1772-1781. (75) Kurokawa, N.; Yoshikawa, H.; Hirota, N.; Hyodo, K.; Masuhara, H. Size-dependent spectroscopic properties and thermochromic behavior in poly(substituted thiophene) nanoparticles. ChemPhysChem 2004, 5, 1609-1615. (76) Eaton P.; P., W. Atomic Force Microscopy. Oxford University Press Inc.: New York, 2010; first edition. (77) Berne, B. J.; Pecora, R. Dynamic Light Scattering: With Applications to Chemistry, Biology, and Physics. Courier Corporation: 2000. (78) Weber, G.; Teale, F. W. J. Determination of the Absolute Quantum Yield of Fluorescent Solutions. Trans. Faraday Soc. 1957, 53, 646-655.

Page 70: Development of Conjugated Polymer Nanoparticles for

63

(79) Fery-Forgues, S.; Lavabre, D. Are Fluorescence Quantum Yields So Tricky to Measure? A Demonstration using Familiar Stationery Products. Journal of Chemical Education 1999, 76, 1260-1264. (80) Press, W. H. Numerical Recipes 3rd Edition: The Art of Scientific Computing.

Cambridge university press: 2007.

(81) Klar, T. A.; Hell, S. W. Subdiffraction Resolution in Far-Field Fluorescence

Microscopy. Opt. Lett. 1999, 24 (14), 954-956.

(82) Subach, F. V.; Patterson, G. H.; Manley, S.; Gillette, J. M.; Lippincott-Schwartz, J.; Verkhusha, V. V. Photoactivatable Mcherry for High-Resolution Two-Color Fluorescence Microscopy. Nat. Methods 2009, 6 (2), 153-159. (83) Bates, M.; Blosser, T. R.; Zhuang, X. Short-Range Spectroscopic Ruler Based on a Single-Molecule Optical Switch. Phys. Rev. Lett. 2005, 94 (10), 108101. (84) Fölling, J.; Belov, V.; Kunetsky, R.; Medda, R.; Schönle, A.; Egner, A.; Eggeling, C.; Bossi, M.; Hell, S. W. Photochromic Rhodamines Provide Nanoscopy with Optical Sectioning. Angew. Chem. Int. Ed. 2007, 46 (33), 6266-6270. (85) Hu, D.; Tian, Z.; Wu, W.; Wan, W.; Li, A. D. Photoswitchable Nanoparticles Enable

High-Resolution Cell Imaging: PULSAR Microscopy. J. Am. Chem. Soc. 2008, 130 (46),

15279-15281.

(86) Chan, Y.-H.; Gallina, M. E.; Zhang, X.; Wu, I.-C.; Jin, Y.; Sun, W.; Chiu, D. T.

Reversible Photoswitching of Spiropyran-Conjugated Semiconducting Polymer Dots.

Anal. Chem. 2012, 84 (21), 9431-9438.

(87) Yu, J.; Hu, D.; Barbara, P. F. Unmasking Electronic Energy Transfer of Conjugated

Polymers by Suppression of O2 Quenching. Science 2000, 289 (5483), 1327-1330.

(88) Gesquiere, A. J.; Park, S.-J.; Barbara, P. F. Hole-Induced Quenching of Triplet and

Singlet Excitons in Conjugated Polymers. J. Am. Chem. Soc. 2005, 127 (26), 9556-9560.

(89) Ralls, K.; Skocpol, W.; Jackel, L.; Howard, R.; Fetter, L.; Epworth, R.; Tennant, D.

Discrete Resistance Switching in Submicrometer Silicon Inversion Layers: Individual

Interface Traps and Low-Frequency (1/F?) Noise. Phys. Rev. Lett. 1984, 52 (3), 228.

(90) Dempsey, G. T.; Vaughan, J. C.; Chen, K. H.; Bates, M.; Zhuang, X. Evaluation of

Fluorophores for Optimal Performance in Localization-Based Super-Resolution Imaging.

Nat. Methods 2011, 8 (12), 1027-1036.

(91) Wang, X.; Groff, L. C.; McNeill, J. D. Photoactivation and Saturated Emission in

Blended Conjugated Polymer Nanoparticles. Langmuir 2013, 29 (45), 13925-13931.

(92) Fraiji, L.; Hayes, D.; Werner, T. Static and dynamic fluorescence quenching experiments for the physical chemistry laboratory. Journal of Chemical Education 1992, 69 (5), 424. (93) Ludwig, A.; Serna, P.; Morgenstein, L.; Yang, G.; Bar-Elli, O.; Ortiz, G.; Miller, E.; Oron, D.; Grupi, A. Development of Lipid-Coated Semiconductor Nanosensors for Recording of Membrane Potential in Neurons. ACS Photonics 2020, 7 (5), 1141-1152. (94) Hania, R.; Thomsson, D.; Scheblykin, I. Host Matrix Dependent Fluorescence Intensity Modulation by an Electric Field in Single Conjugated Polymer Chains. The Journal of Physical Chemistry B 2006, 110, 25895-25900.

Page 71: Development of Conjugated Polymer Nanoparticles for

64

(95) Limori, T.; Awasthi, K.; Chiou, C.; Diau, E.; Ohta, N. Fluorescence enhancement

induced by quadratic electric-field effects on singlet exciton dynamics in poly(3-

hexylthiophene) dispersed in poly(methyl methacrylate). Phys. Chem. Chem. Phys., 2019,

21, 5695-5704

(96) Kruss, S.; Landry, M.; Vander, E.; Lima, B.; Reuel, N.; Zhang, J.; Nelson, J.; Mu, B.;

Hilmer, A.; Strano, M. Neurotransmitter Detection Using Corona Phase Molecular

Recognition on Fluorescent Single-Walled Carbon Nanotube Sensors. Journal of American

Chemical Society 2017, 136, 713.

(97) Polo, E.; Kruss, S. Nanosensors for neurotransmitters. Anal. Bioanal. Chem. 2016,

408 (11), 2727–2741.