You are on page 1of 149

NOVEL SPECTROSCOPY

TECHNIQUES FOR FUNCTIONAL


IMAGING OF BACTERIAL BIOFILMS
AT MICROSCOPIC LEVEL

ANDREAS KARAMPATZAKIS
(M. Sc.)

A THESIS SUBMITTED FOR THE


JOINT DEGREE OF DOCTOR OF PHILOSOPHY

NUS GRADUATE SCHOOL FOR INTEGRATIVE


SCIENCES AND ENGINEERING
NATIONAL UNIVERSITY OF SINGAPORE

DEPARTMENT OF PHYSICS
IMPERIAL COLLEGE LONDON

2017

Supervisors:
Professor Thorsten Wohland, National University of Singapore
Professor Peter Török, Imperial College London

Examiners:
Associate Professor Sanjay Swarup, National University of Singapore
Dr Darryl Overby, Imperial College London
Associate Professor Kimberly Kline, Nanyang Technological University
Declaration

I hereby declare that this thesis is my original work and it has been written by

me in its entirety. I have duly acknowledged all the sources of information which

have been used in the thesis.

This thesis has also not been submitted for any degree in any university previ-

ously.

Andreas Karampatzakis

August 2017

iii
iv
Copyright Declaration

The copyright of this thesis rests with the author and is made available under

a Creative Commons Attribution Non-Commercial No Derivatives licence. Re-

searchers are free to copy, distribute or transmit the thesis on the condition that

they attribute it, that they do not use it for commercial purposes and that they do

not alter, transform or build upon it. For any reuse or redistribution, researchers

must make clear to others the licence terms of this work.

Results of this thesis have been partially published in:

- A. Karampatzakis, J. Sankaran, K. Kandaswamy, S. A. Rice, Y. Cohen, T. Wohland “Mea-

surement of oxygen concentrations in bacterial biofilms using transient state monitor-

ing by single plane illumination microscopy” Biomed. Phys. Eng. Express 3 035020

(2017).

- A. Karampatzakis, C. Z. Song, L. P. Allsopp, A. Filloux, S. A. Rice, Y. Cohen, T. Wohland,

P. Török “Probing the internal micromechanical properties of Pseudomonas aeruginosa

biofilms by Brillouin imaging” npj Biofilms and Microbiomes, 3, 1–7, (2017).

v
vi
Acknowledgements

The long journey that has led to the writing of this thesis would not have been

possible without the help of many persons that I was fortunate enough to meet

along the way.

First and foremost, I would like to express my sincere thanks to my super-

visors Thorsten Wohland (NUS) and Peter Török (ICL) for giving me the oppor-

tunity to work on this project, and for their continuous support throughout this

project. Most importantly, I would like to thank them for their trust in me and

my work despite the difficulties along the way.

Many thanks to my thesis advisor committee members Nanguang Chen (NUS)

and Shu Sin Chng (NUS) for bringing their expertise to the discussions of my re-

sults. I would also like to thank professors Yehuda Cohen (SCELSE, NTU), Scott

Rice (SCELSE, NTU) and Alain Filloux (ICL) for accepting me in their labs where

most of the biological work for this thesis was done.

My appreciation goes to all my colleagues at CBIS, SCELSE, and ICL, and es-

pecially to Jagadish Sankaran, Cheng–Ze Song, Luke Allsop, Xue–Wen Ng, Emilio

Sanchez, Anand Singh, Kumaravel Kandaswamy, Sarala Neomi Tantirimudalige,

Pei–Ju Wung, Giuseppe Antonacci and Guillaume Lepert.

I would also like to express my gratitude to NUS Graduate School for Inte-

grative Sciences and Engineering for funding my studies and expenses via the

NGS Scholarship, and for being extremely helpful and accommodating.

Last but not least, I would like to thank my mother, and my wife, Charis Sim,

for their love, patience, understanding and support throughout these years.
vii
viii
Contents

Summary xiii

List of Tables xv

List of Figures xviii

List of Abbreviations and Symbols xix

1 General Introduction 1

1.1 Historical perspective . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Photophysics and Photochemistry 8

2.1 Light scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1.2 Brillouin scattering . . . . . . . . . . . . . . . . . . . . . . . . 10

2.1.3 Brillouin modulus and mechanical properties . . . . . . . . 12

2.2 Luminescence phenomena . . . . . . . . . . . . . . . . . . . . . . . 15

2.2.1 Energy states . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.2.2 Singlet vs Triplet states . . . . . . . . . . . . . . . . . . . . . 18

2.2.3 Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.2.4 Quenching by Oxygen . . . . . . . . . . . . . . . . . . . . . . 21

ix
CONTENTS

3 Biofilms 25

3.1 The biofilm mode of life . . . . . . . . . . . . . . . . . . . . . . . . . 25

3.2 EPS and mechanical properties of biofilms . . . . . . . . . . . . . . 29

3.3 The role of oxygen in biofilms . . . . . . . . . . . . . . . . . . . . . . 32

4 Brillouin Microscopy for Mechanical Measurements 35

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

4.1.1 Confocal microscopy . . . . . . . . . . . . . . . . . . . . . . . 35

4.1.2 Brillouin spectroscopy . . . . . . . . . . . . . . . . . . . . . . 37

4.1.3 The conventional use of term ’stiffness’ . . . . . . . . . . . . 40

4.2 Material and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . 42

4.2.1 Experimental setup . . . . . . . . . . . . . . . . . . . . . . . . 42

4.2.2 Sample preparation . . . . . . . . . . . . . . . . . . . . . . . 45

4.2.3 Acquisition and analysis . . . . . . . . . . . . . . . . . . . . . 47

4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4.3.1 Measurements in hydrogel beads . . . . . . . . . . . . . . . 51

4.3.2 Characterisation of the stiffness of P. aeruginosa biofilms . 51

4.3.3 Brillouin shift measured at increasing depths . . . . . . . . 53

4.3.4 Temporal changes of stiffness . . . . . . . . . . . . . . . . . . 55

4.3.5 Effect of flow rate . . . . . . . . . . . . . . . . . . . . . . . . . 58

4.4 Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

5 SPIM – TRAST for O2 Measurements 65

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

5.1.1 Single Plane Illumination Microscopy (SPIM) . . . . . . . . 65

5.1.2 Transient States (TRAST) spectroscopy . . . . . . . . . . . . 67

5.2 Material and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . 70

5.2.1 Simplified energy system . . . . . . . . . . . . . . . . . . . . 70

x
CONTENTS

5.2.2 TRAST monitoring theory . . . . . . . . . . . . . . . . . . . . 71

5.2.3 Derivation of TRAST contrast . . . . . . . . . . . . . . . . . . 73

5.2.4 TRAST contrast as a measure of [O2 ] . . . . . . . . . . . . . . 75

5.2.5 Experimental setup . . . . . . . . . . . . . . . . . . . . . . . . 76

5.2.6 Sample preparation . . . . . . . . . . . . . . . . . . . . . . . 78

5.2.7 Acquisition and analysis . . . . . . . . . . . . . . . . . . . . . 81

5.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

5.3.1 Measurements in bulk solutions . . . . . . . . . . . . . . . . 83

5.3.2 Measurements in heterogeneous samples . . . . . . . . . . 85

5.3.3 Measurements in bacterial biofilms . . . . . . . . . . . . . . 86

5.4 Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

5.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

6 Afterword 96

6.1 Overall conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

6.2 Limitations and outlook . . . . . . . . . . . . . . . . . . . . . . . . . 98

Appendices

A Typical TRAST illumination scheme A-1

B Protocol for TRAST Contrast vs Oxygen calibration B-1

xi
xii
Summary

This thesis deals with the development and application of spectroscopy meth-

ods in the study of bacterial biofilms at the microscopic level. More specifically,

with a combination of Brillouin microscopy with fluorescence imaging, we were

able to probe the stiffness in the interior of growing biofilms in a non–contact

manner, where we found distinct profiles indicating colonies in different stages

of their lifecycle. With the implementation of Transient State (TRAST) imaging

on a Single Plane Illumination Microscope (SPIM) we were able to measure the

triplet state populations and develop a protocol that can quantify oxygen con-

centrations. When biofilms were measured, steep oxygen gradients were found.

The report is structured as follows: Chapter 1 provides a historical overview

of microscopy and spectroscopy methods, and sets the scope of the thesis. Chap-

ter 2 presents the theoretical background on light scattering mechanisms that is

essential for interpreting Brillouin imaging results, as well as on fluorescence

and quenching phenomena which are key concepts for understanding TRAST

imaging. Chapter 3 introduces and discusses the concept of biofilms as a mode

of bacterial life. In particular, the distinct structural morphology of biofilms is

highlighted, together with the importance of oxygen in their lifecycle. Chap-

ters 4 and 5 contain all the information for the development of the experimental

methods, together with the results and discussions for Confocal Brillouin mi-

croscopy and SPIM – TRAST, respectively. Lastly, Chapter 6 contains the overall

conclusions of this project and an outlook.

xiii
xiv
List of Tables

2.1 Types of scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.2 Typical lifetimes of electronic transitions . . . . . . . . . . . . . . . 18

5.1 Characteristics of Eosin Y . . . . . . . . . . . . . . . . . . . . . . . . 78

5.2 Triplet relaxation and intersystem crossing times for Eosin Y in air–

saturated and deaerated solutions . . . . . . . . . . . . . . . . . . . 83

A.1 Lookup table for TRAST image acquisition settings . . . . . . . . . A-2

xv
xvi
List of Figures

2.1 Typical scattering spectrum. . . . . . . . . . . . . . . . . . . . . . . 10

2.2 Schematic of momentum conservation in Brillouin scattering . . 11

2.3 Types of material deformations based on the distribution of forces. 13

2.4 A typical Jablonski diagram. . . . . . . . . . . . . . . . . . . . . . . . 16

2.5 Absorption and emission spectra of Eosin Y. . . . . . . . . . . . . . 20

3.1 P. aeruginosa bacterial cells forming a colony . . . . . . . . . . . . 27

3.2 Schematic illustrating the late stages of a biofilm life cycle. . . . . 31

4.1 Diagram of a confocal microscope. . . . . . . . . . . . . . . . . . . . 36

4.2 Schematic of the confocal Brillouin microscope. . . . . . . . . . . . 42

4.3 Schematic of our continuous–culture biofilm flow cell setup . . . 46

4.4 Brillouin spectrum of water, used for calibration. . . . . . . . . . . 48

4.5 A typical Brillouin spectrum from a point inside a biofilm . . . . . 50

4.6 Brillouin imaging in hydrogel beads. . . . . . . . . . . . . . . . . . . 52

4.7 Brillouin, widefield and fluorescence images of typical colonies. . 53

4.8 Brillouin shift (average and standard deviations within ROIs) mea-

sured in colonies of various sizes. . . . . . . . . . . . . . . . . . . . 54

4.9 Brillouin shift measured in increasing depths in a biofilm colony. 54

4.10 Brillouin shift measured along a cross-section of a large colony at

various depths. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

xvii
LIST OF FIGURES

4.11 Brillouin shift (average and standard deviations within ROIs) mea-

sured in colonies at various times post–inoculation. . . . . . . . . 57

4.12 Brillouin shift measured along a cross–section of a colony at two

consecutive days. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

5.1 SPIM – TRAST schematic. . . . . . . . . . . . . . . . . . . . . . . . . 67

5.2 Simplified, three–state Jablonski diagram. . . . . . . . . . . . . . . 70

5.3 The effect of triplet relaxation and intersystem crossing time on

the appearance of TRAST curves. . . . . . . . . . . . . . . . . . . . . 73

5.4 TRAST contrast for a number of triplet relaxation times. . . . . . . 75

5.5 TRAST contrast plotted against oxygen concentrations. . . . . . . 76

5.6 Mounting of samples and optical sectioning schematics. . . . . . . 81

5.7 Assessment of bleaching recovery . . . . . . . . . . . . . . . . . . . 82

5.8 TRAST measurements in bulk solutions of Eosin Y. . . . . . . . . . 84

5.9 TRAST contrast measured in samples of varying oxygen content. . 84

5.10 TRAST measurements in hydrogel beads. . . . . . . . . . . . . . . . 86

5.11 TRAST measurements in P. aeruginosa biofilms. . . . . . . . . . . . 88

5.12 TRAST contrast image and oxygen concentration contours inside

a colony. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

B.1 Parallel measurements with an optical sensor and SPIM. . . . . . . B-2

xviii
List of Abbreviations and Symbols

List of Abbreviations

A. A. Alginate Acid

AFM Atomic Force Microscopy

BE Beam Expanders

BS Beam Splitter

CARS Coherent anti–Stokes Raman spectroscopy

C Cylindrical lens

CCD Charge–Coupled Device

CW Continuous Wave

DB Doublet lens

DICM Differential–Interference Contrast Microscopy

DM Dichroic Mirror

EPS Extracellular Polymeric Substances

e–DNA extracellular DNA

FCS Fluorescence Correlation Spectroscopy

FCCS Fluorescence Cross–Correlation Spectroscopy

FEP Fluorinated Ethylene Propylene

xix
ABBREVIATIONS AND SYMBOLS

FLIM Fluorescence Lifetime Imaging Microscopy

FRAP Fluorescence Recovery After Photobleaching

FRET Förster (or Fluorescence) Resonance Energy Transfer

FSR Free Spectral Range

FxP Frequency per Pixel

HOMO Highest Occupied Molecular Orbital

LB Luria–Bertani medium (Lysogeny Broth)

LUMO Lowest Unoccupied Molecular Orbital

M Mirror

NA Numerical Aperture

OPFOS Orthogonal–Plane Fluorescence Optical Sectioning

PCS Photon Correlation Spectroscopy

RLS Raman correlation spectroscopy

ROI Region Of Interest

S Spherical lens

SNR Signal–to–Noise Ratio

SPIM Selective (or Single) Plane Illumination Microscopy

TRAST imaging Transient State imaging

TIRF Total Internal Reflection Fluorescence

VIPA Virtually Imaged Phase Array

WD Working Distance

xx
ABBREVIATIONS AND SYMBOLS

List of Symbols

[Q] Concentration of quencher Q

α Duty cycle of pulse train

A Surface cross–section

γ Scaling factor that depends on the concentration of fluorophores and

the detection efficiency of the camera

Γ Concentration of fluorophores

δ Optical path difference

∆v B Spectral linewidth of Brillouin peaks

η Viscosity

θ Scattering angle

θi Angle between the prism and the beam at the common path interfer-

ometer

λ0 Wavelength of incident light

ξ Conversion factor between the emitted intensity and the digital read-

out of the camera

ρ Density

σexc Excitation cross–section

σ(λ) Wavelength–dependent molecular absorption cross–section

τ01 Excitation lifetime

τ10 Relaxation time of the first excited state

τT Triplet relaxation time

τT 0 Triplet relaxation time in absence of quencher

xxi
ABBREVIATIONS AND SYMBOLS

τf Fluorescence lifetime

τf 0 Fluorescence lifetime in absence of quencher

τi sc Intersystem crossing time

Bn Position of Brillouin peaks on the dispersion axis (n = 1, 2)

c Speed of light in vacuum

d Particle size

dg Length of glass prism of the common path interferometer

D Diffusion coefficient

E Young’s modulus

f Focal length

F Force

fQ Quenching efficiency

G Shear modulus

h Planck’s constant (≈ 6.63 × 10−34 J s)

I exc Excitation intensity

If Fluorescence intensity

If 0 Fluorescence intensity in absence of quencher

ICW Fluorescence intensity taken with 100% duty cycle

I shor t Fluorescence intensity taken with 1% duty cycle

K Bulk modulus

k0 Diffusion–controlled bimolecular quenching constant

k 0n Fluorescence excitation rate (n = 1, 2...)

k2 Triplet state population rate after the onset of excitation

xxii
ABBREVIATIONS AND SYMBOLS

kB Boltzmann’s constant (≈ 8.6 × 10−5 eV/K)

kf Fluorescence rate

ki Wave vector of incident photons

ki c Internal conversion rate

k i sc Intersystem crossing rate

k nr Rate of non–radiative processes

kp Wave vector of phonons

k ph Phosphorescence rate

kq Bimolecular quenching constant

ks Wave vector of scattered photons

kT Triplet, non–radiative, relaxation rate

M Longitudinal modulus

M0 Storage (or Brillouin) modulus

M 00 Loss modulus

NA Avogadro’s number (6.022 × 1023 mol−1 )

n Refractive index

P0 Population of ground electronic state

P1 Population of first excited electronic state

P l aser Excitation power

PT Population of triplet state


eq
PT Relative triplet state population at equilibrium

qf Quantum efficiency of a fluorophore

qT Triplet quantum yield

R Molecular radius

xxiii
ABBREVIATIONS AND SYMBOLS

Sn Singlet electronic energy levels (n = 0, 1, 2...)

T Period of pulse train

T ac q Total acquisition time

Ti l l Total illumination time

Tn Triplet electronic energy levels (n = 0, 1, 2...)

v Light frequency

va Frequency of absorbed photons

vi Frequency of incident photons

vf Frequency of fluorescence photons

vp Frequency of phonons

v ph Frequency of phosphorescence photons

vs Frequency of scattered photons

V Acoustic velocity

Vn Vibrational energy levels (n = 0, 1, 2...)

vB Brillouin frequency shift

w Pulse width

xxiv
Chapter 1

General Introduction

1.1 Historical perspective

It is believed that the first compound microscopes were conceived and pro-

duced in the 1590s by two Dutch spectacle makers, Zacharias Jansen and his

father, Hans, who experimented with placing multiple lenses in a tube. The first

documented scientific observations on microscopy, however, came from Robert

Hooke and his book Micrographia, published in January 1665 [1]. Amongst oth-

ers, in this book Hooke coined the term cell to describe the basic unit of life,

inspired by the resemblance of the microstructures he saw in plant samples to

the arrangement of cells in a monastery. Micrographia went on to become one

of the first scientific best sellers and is today recognised as a milestone in the

evolution of microscopy [2].

Another name associated with the birth of microscopy is that of Antonie

van Leeuwenhoek, who is also considered the father of microbiology [3]. van

Leeuwenhoek was not a scientist by training, and it is believed that his interest in

microscopy was sparked from reading Hooke’s Micrographia. He built “single–

lens microscopes” which would today be classified as magnifying glasses. His

microscopes were simpler in design compared to other instruments of his time

1
CHAPTER 1. GENERAL INTRODUCTION

(compound microscopes were already in use). However, due to his great skill

in lens grinding and his attention to lighting conditions, he was able to achieve

magnification levels exceeding 200 times and he produced clearer and brighter

images than any of his colleagues. This allowed him to make one of his most

important observations, what he called animacules (little animals), that were

moving in water. His letter to the Royal Society, date 9 October 1676, where he

reports these findings is believed to be one of the first observations on living

bacteria ever recorded [4]. His findings were later reproduced and confirmed by

Hooke.

In one of his most famous experiments, van Leeuwenhoek looked at the

plaque formed in his own teeth and compared his findings with samples taken

from his wife, daughter as well as from two older men who reportedly never

brushed their teeth. What he saw “with great wonder” was “very prettily a-

moving” [...] “animacules” of different sizes and shapes. More specifically, in

the case of the older men, he reports “an unbelievably great company of liv-

ing animalcules, a–swimming more nimbly than any I had ever seen up to this

time. The biggest sort ... bent their body into curves in going forwards ... More-

over, the other animalcules were in such enormous numbers, that all the water

... seemed to be alive”. With this, van Leeuwenhoek had effectively made the

first observation of living bacterial biofilms1 in the human body.

On the theoretical front, the physical limits in discernible objects were in-

vestigated in the 19th century. In 1835, George Airy studied the diffraction pat-

tern created by light passing through an illuminated circular aperture [5], which

is now referred to as “Airy disc” and considered a fundamental concept in mi-

croscopy. A few years later, in 1873, Ernst Abbe in a landmark paper reported

that “the smallest resolvable distance between two points using a conventional

microscope may never be smaller than half the wavelength of the imaging light”
1
An introduction and historical perspective of biofilms can be found in Chapter 3.

2
CHAPTER 1. GENERAL INTRODUCTION

[6], which has become the well known Abbe’s resolution criterion.

Around the same period, the first observations of fluorescence are docu-

mented. In 1845, Sir Frederik William Herschel noted that a quinine solution,

exhibits a “vivid and beautiful celestial blue color” when illuminated and ob-

served under certain incidences of sunlight [7]. In 1852, George Stokes described

Herschel’s observation in greater detail and set the theoretical foundation for

fluorescence [8].

It did not take long till the first fluorescent dye, fluorescein, was synthesized

by Adolf von Bayer in 1871 [9], and the first fluorescence microscopes were de-

veloped. In 1904, August Köhler used ultraviolet light to observe luminescence

in a microscope built at Zeiss Optical Works in Jena, Germany, and in the coming

years researchers started using fluorescent dyes to enhance autofluorescence of

cells and tissues [10].

A number of techniques that are heavily used today were introduced in the

first half of the twentieth century. Most notably, Frits Zernike’s work on Phase

Contrast Microscopy in the 1930s [11], Wilhelm Josef Schmidt and Shinya In-

oué’s work on Polarizing Microscopy [12, 13], and Francis Smith and Georges

Nomarski’s work on Differential–Interference Contrast Microscopy (DICM) in

the 1950s [14, 15] have all revolutionised the way we look at cells. Around the

same time came the invention of the Confocal Microscope by Marvin Minsky

which led to an improvement of resolving power of fluorescence [16, 17]. The

first confocal microscopes were produced in late 1980s [18], however, scanning

confocal microscopes as those routinely used today only appeared in the ’90s,

after advances in laser systems, electronics etc. allowed their production [19].

Other developments in the last decades of the twentieth century include

the application of the phenomenon of Total Internal Reflection in Fluorescence

(TIRF) microscopy, as demonstrated by Axelrod in 1981 who visualized cellu-

3
CHAPTER 1. GENERAL INTRODUCTION

lar focal adhesions on a glass slide [20]. Additionally, the concept of illumina-

tion perpendicular to detection – which was introduced by Siedentopf and Zsig-

mondy in 1902 as ultramicroscopy [21] – was revisited by Voie et al. who built

an Orthogonal–Plane Fluorescence Optical Sectioning (OPFOS) microscope in

1993 [22]. The latter has led to the development of what is commonly known

today as Selective (or Single) Plane Illumination Microscopy (SPIM) – a term

introduced by Stelzer and colleagues in 2004 [23] – a valuable technique that

allows high spatio–temporal resolution, minimal photo–toxicity and thus more

measurements per sample [24].

Certainly, the use of microscopes goes beyond imaging; first, a number of

fluorescence spectroscopy methods have been developed in the last decades,

often referred to as F–Techniques [25]. For instance, Fluorescence Correlation

Spectroscopy (FCS), which was developed independently by Watt Webb and

Rudolf Rigler during the early 1970s [26, 27], measures fluctuations in the flu-

orescence signal to quantify chemical and biological reactions, diffusion and

flow. The breakthrough of the technique was its implementation in a confo-

cal arrangement by Rigler and coworkers in 1990s, which allowed tracking at

single molecule level [28]. An extension of this method is Fluorescence Cross–

Correlation Spectroscopy (FCCS) that uses two distinct fluorophores. The method

was introduced by Ricka and Binkert in 1989 [29] and first applied in biological

systems by Schwille in 1997 [30]. Other techniques include Fluorescence Re-

covery After Photobleaching (FRAP), developed to study dynamics of proteins

within and between cellular membranes [31], Förster (or Fluorescence) Reso-

nance Energy Transfer (FRET) imaging, which can measure distance and detect

molecular interactions in a number of dynamic systems, such as during protein

conformational changes [32] and Fluorescence Lifetime Imaging Microscopy

4
CHAPTER 1. GENERAL INTRODUCTION

(FLIM) that can be used to detect physicochemical changes in the environment

of the fluorophore [33].

Many non–fluorescence spectroscopic techniques have also been implemen-

ted in microscopic setups allowing the extraction of functional, chemical and

structural information at optical spatial resolutions. For example, Raman cor-

relation spectroscopy (RLS) [34] and coherent anti–Stokes Raman spectroscopy

(CARS) [35] are used to identify molecules and study chemical bonding, and

Photon Correlation Spectroscopy (PCS) is used to determine particle sizes in

suspensions [36].

1.2 Scope

Biofilms are structured bacterial communities in which cells are held together

within a matrix formed by self–secreted polymeric compounds, termed the Ex-

tracellular Polymeric Substances (EPS) [37]. Through cell–to–cell signalling, known

as quorum sensing, bacterial cells are known to coordinate the biofilm forma-

tion, and collaborate in a way that benefits the entire population [38, 39]. This

unique characteristic enables bacteria to survive unpredictable environmental

stress, such as temperature or pH changes, desiccation or pressure, thus mak-

ing them collectively very resistant to attacks that would otherwise destroy the

same cells in their planktonic state. Biofilms have a significant impact on hu-

man health being the leading cause of bacterial infections [40]. They are re-

sponsible for contaminations in the food and dairy industry [41, 42] and on sur-

faces of catheters [43] and medical devices [44], as well as for the degradation of

drinking water quality [45] and biocorrosion [46].

Biofilm colonies are still largely treated as uniform populations similar to

planktonic cells. However, this approach is incorrect, as studies have shown

that a number of chemical and physical gradients exist within the colonies [37].

5
CHAPTER 1. GENERAL INTRODUCTION

The formation of gradients is stabilized by the immobilization of bacterial cells

within the matrix, and leads to the establishment of sub–populations of cells

that exist in distinct physiological states. As a consequence of this heterogene-

ity, studies of the genetic and physiological pathways that control biofilm devel-

opment, e.g., transcriptomics, metabolomics and proteomics, are almost inde-

cipherable due to the effect of averaging changes across the biofilm for cells in

different physiological states.

It is therefore essential to be able to identify and measure the aforemen-

tioned gradients so that the status of the cells can be defined along the differ-

ent stages of biofilm development, as well as in different structural parts of the

biofilm such as near the outer surface versus deep in the micro–colony core. In

this thesis, two novel spectroscopic techniques, namely Brillouin Imaging and

Transient State monitoring (TRAST) are presented, further developed and ap-

plied on the study of bacterial biofilms.

Brillouin spectroscopy measures the frequency shift of light scattered by in-

herent thermal acoustic waves, which is associated with the stiffness of the me-

dium. Recently, Brillouin spectroscopy has been combined with optical mi-

croscopy, allowing one to perform high–resolution mechanical imaging in a non–

invasive manner [47]. Here, we apply confocal Brillouin microscopy to study

the stiffness of living Pseudomonas aeruginosa biofilms along different stages of

their life cycle, as they grow under constant flow inside a flow cell.

TRAST monitoring was introduced by Tor Sandén, Gustav Persson and Jerker

Widengren [48, 49] as a method to measure kinetics of transient photo–induced

states without the need of fast–scale time–resolved measurements. In TRAST

imaging, the kinetic rates of transient states can be calculated by fitting on a

electronic state model the time–averaged fluorescence intensity of samples illu-

minated with pulse trains of different characteristics. The extracted triplet relax-

6
CHAPTER 1. GENERAL INTRODUCTION

ation time can then be related to concentrations of quenching agents according

to the Stern–Volmer equation. Here, we implement TRAST imaging on a SPIM

set–up and present an experimental protocol that can measure concentrations

of molecular oxygen inside biofilm colonies.

Precise knowledge of the molecular concentration of oxygen and the me-

chanical properties of biofilms can improve our understanding of the processes

that take place at microscopic levels. For example, oxygen is a key metabolite

that is believed to be responsible for the cueing of dispersal, during which, dras-

tic changes of the structural properties of colonies are expected to take place.

Additionally, oxygen becomes a limiting factor in deeper parts the biofilm, from

where dispersal has also been found to originate. Therefore, we hypothesise that

the depletion of oxygen and the onset of dispersal could be correlated and col-

localised, and we propose the development of methods that can provide us with

quantitative information at optical resolutions. Findings can potentially be of

help in the assessment of drug treatments as well as for strategising methods to

inhibit the growth or promote the removal of biofilms.

7
Chapter 2

Photophysics and Photochemistry

This chapter provides a theoretical background of light–matter interactions that

is necessary for the understanding of the spectroscopy techniques presented in

Chapters 4 and 5, and for the interpretation of the results.

2.1 Light scattering

2.1.1 Overview

Scattering is termed the change of direction of light propagation, resulting from

its incidence and interaction with matter. It can be classified as inelastic or elas-

tic, depending on whether there is energy transfer during the light–matter inter-

action or not.

Other classification methods are based on the nature, size and geometry of

the scatterer: Rayleigh scattering refers to scattering by particles whose size (d )

is much smaller than the wavelength of the incident light (λ0 ), Mie scattering

refers to scattering by spherical particles of any size, and Tyndall scattering is

an extended model of Mie scattering that includes particles of any shape. Bril-

louin scattering – which is of main interest in this thesis and will be described in

8
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

detail later on – occurs from the interaction of photons with acoustic phonons

that exist in matter. Lastly, Raman scattering refers to the interaction of light

with optical phonons that arise from intra–molecular vibrations and rotations

with energies higher than acoustic phonons. Table 2.1 summarises the above; a

comprehensive review of the different scattering types can be found at ref. [50],

and an in–depth discussion about Brillouin scattering and its connection with

mechanical properties of material can be found at ref. [51].

Type Scatterer Spectral shift


Rayleigh Particles (d << λ0 ) N/A
Mie Spherical particles (any d ) N/A
Tyndall Particles of any shape (any d ) N/A

Brillouin Acoustic phonons (density fluctuations) 5 – 20 GHz


Raman Optical phonons (vibrational or rotational modes) > 1 THz

Table 2.1: Types of scattering. d is the particle size and λ0 the wavelength of
incident light. Spectral shift corresponds to the amount of energy ex-
change during the inelastic scattering processes.

Information about the structure and dynamics of a material being probed by

light can be extracted by studying its scattering spectra. A typical spectrum of

the scattered light, where the frequency of scattered light is plotted on x–axis,

is shown in Figure 2.1. The strongest component corresponds to the elastic

(Rayleigh) scattering, giving rise to the peak located at v i that is the frequency of

the incident light. Additionally, peaks in higher (anti–Stokes) and lower (Stokes)

frequencies are typically observed. These correspond to the energy gain or loss,

respectively, of the photons as a result of scattering. Such scattering phenom-

ena where energy exchange takes place are referred to as inelastic scattering,

and include Brillouin and Raman scattering.

9
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

Rayleigh

Stokes anti-Stokes

Brillouin Brillouin
Raman Raman

vi frequency

Figure 2.1: Typical scattering spectrum showing the central, elastic, Rayleigh peak to-
gether with a Brillouin and a Raman peak on each side. The two inelastic
side bands are referred to as the Stokes and the anti–Stokes peaks associ-
ated with a lower and higher frequency shift, respectively.

2.1.2 Brillouin scattering

Brillouin scattering is an inelastic scattering process arising from the interaction

of light with density fluctuations – termed acoustic phonons or acoustic lattice

vibrations – that propagate inside a medium. The scattered photons experience

frequency shift by an amount equal to the frequency of the scattering sound

wave [52]. The phenomenon was first reported in 1922 by Lèon Brillouin [53].

For low photon densities, Brillouin scattering is termed as spontaneous and the

scattered light is proportional to the intensity of the incident beam. When in-

cident powers exceed certain thresholds, optical fields substantially contribute

to the phonon population leading to a stimulated excitation of the medium.

Stimulated Brillouin scattering is a non–linear effect commonly studied in the

context of optical fibres. It becomes more prominent for longer optical fibres,

and poses limitations in optical fibre communications since it creates signifi-

cant back–scattering of the beam [54]. This thesis only deals with spontaneous

Brillouin scattering.

10
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

ks
(a) (b)
θ
kp
kp ki
ks
ki

kp
ki
ks

(c)

Figure 2.2: Schematic of momentum conservation in Brillouin scattering. (a) The gen-
eral case. (b, c) Stokes and anti–Stokes scattering, respectively, from a back–
scattering geometry such as the one used in the experiments of this thesis.

From a classical physics point of view, a travelling acoustic phonon creates

a periodic change of density and hence of the local dielectric constant and re-

fractive index in a medium. To an incident beam of light, the medium acts as

a scattering lattice, which is moving at the speed of sound. For that matter, the

scattered light experiences a positive or negative shift in frequency, depending

on the direction of the phonon movement, due to the Doppler effect.

From a quantum mechanical point of view, an incident photon (wave vector

k i , frequency v i , Figure 2.2) interacts with a phonon and is annihilated. There is

instantaneous generation of a scattered photon (wave vector k s and frequency

v s , travelling at an angle θ) and an annihilation or creation of a phonon (wave

vector k p , frequency v p ). The scattered photon will have either gained or lost

energy equal to the energy of the phonon that was annihilated or created, re-

spectively, referred to as Stokes and anti–Stokes processes, which are responsi-

ble for the symmetry of the spectra (Figure 2.1).

Equations (2.1–2.2) show the energy and momentum conservation in Bril-

louin scattering, respectively. The minus symbol corresponds to the Stokes scat-

11
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

tering, where the material absorbs the photon, creates a phonon and inelasti-

cally emits a photon with a lower energy than that of the absorbed photon. The

plus symbol corresponds to the anti–Stokes scattering, in which the incoming

photon absorbs a phonon (by phonon annihilation), and a photon with a higher

energy than that of absorbed photon is emitted.

hv s = h(v i ± v p ) (2.1)

k s = ki ± k p (2.2)

The absolute frequency shift of the Brillouin peaks can be calculated by:

2n
vB = V sin(θ/2) (2.3)
λ0

where V is the acoustic velocity, n the refractive index, λ0 the wavelength of the

incident light and θ the scattering angle. It can readily be seen that for θ = 0◦

there is no shift in frequency, whilst, for θ = 180◦ , which is the case in back–

scattering set–ups such as the one used in this thesis, the shift is maximised so

that:
2n
v B max = V (2.4)
λ0

2.1.3 Brillouin modulus and mechanical properties

In material sciences, mechanical properties of materials refer to their response

to applied forces. While a number of measures can be used to quantify the be-

haviour or a sample under stress (force per unit area) depending on the distri-

bution of forces, all of them are based on the same principle, that is, expressing

the ratio of stress over strain (ratio of deformation over initial length).

12
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

F
A

(a) (b) (c)

Figure 2.3: Types of material deformations based on the distribution of forces. (a) Axial
stress (F /A), applied perpendicularly to the area cross–section (A), resulting
in deformation (∆L). (b) Shear stress (F /A) applied coplanarly to the area
cross–section (A), resulting in shear deformation (∆x). (c) Force (F ) applied
uniformly to the material leading to a change in volume.

In an isotropic material, Young’s modulus, E , is defined as the ratio between

a stress that acts to change the length of a body and the fractional change in

length caused by this force, as shown in Figure 2.3a. The shear modulus, G, is

defined as the ratio between shear stress (stress applied coplanarly to the area

cross section) and shear strain, and describes a situation where all the strains

perpendicular to the stress direction are zero, that is, the volume of the mate-

rial remains unchanged while alterations in the shape occur (Figure 2.3b). The

bulk modulus, K , refers to the situation where force is applied uniformly to the

material and all the strains normal to the stress direction are equal; it is com-

monly used for liquids, and is defined as the ratio of applied stress, or pressure,

to volume change (Figure 2.3c).

A parameter more closely related to our study is the longitudinal modulus,

M . It is defined as the ratio of axial stress to axial strain, that is stress and strain

applied only along the longitudinal axis of the material. It differs from Young’s

modulus only in that there is no shear strain produced. The stress that a sam-

ple experiences under longitudinal acoustic waves in a backscattering Brillouin

microscope, such as the one used in our study, fall under this category.

Biological samples exhibit a viscoelastic behaviour, that is they respond to

stress by a fast deformation followed by a return to the original state – unless

strained beyond their breaking point – with a time delay which corresponds to

13
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

the energy that is lost during the process. The mathematical formulation of the

complex longitudinal modulus of such a medium can be derived from the dif-

ferential equation describing the propagation of a longitudinal acoustic wave in

a medium [55] and is expressed as:

∆v B
µ ¶
M = M 0 + i M 00 = ρV 2 + i ρV 2 (2.5)
vB

where ρ is the density and ∆v B is the spectral linewidth of the Brillouin peaks.

Further, M 0 and M 00 are termed storage – or Brillouin – and loss moduli, respec-

tively.

Recalling equation (2.3), it can be seen that the Brillouin modulus is con-

nected to the experimentally measured Brillouin frequency shift, u B , via:

¶2
vB λ
µ
0
M =ρ . (2.6)
2n

It should be noted, however, that this relation is valid in general for isotropic me-

dia; in anisotropic matter such as biological samples where direction–dependent

hypersound velocities exist and moduli are expressed as tensors, it can only be

specifically applied to a particular directions. Therefore, even though the lon-

gitudinal modulus is generally related to the bulk and shear moduli by M =

K +4/3G, a direct translation of the measured quantity to commonly–used mod-

uli is not possible. Caution is thus needed when interpreting Brillouin spectra,

as it is later discussed in section 4.4.

14
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

2.2 Luminescence phenomena

Luminescence is generally termed the phenomenon of light emission by a sub-

stance that has not been heated. There are a number of different types of lu-

minescence; for example, chemiluminescence and electrochemiluminescence

is the emission of light resulting from a chemical or biochemical reaction, re-

spectively, bioluminescence results from biochemical reactions in a living or-

ganism, crystalloluminescence is emitted during crystallization, tribolumines-

cence is the result of breaking bonds inside a material when it is scratched or

rubbed, etc.

In this chapter we will be focusing on photoluminescence that is light emis-

sion from a material after absorption of photons. Photoluminescence can fur-

ther be separated into fluorescence and phosphorescence; in both processes,

light is produced during the electronic transition from higher to lower electronic

states, and the energy of the emitted light is characteristic of the transition en-

ergy. The main way that fluorescence differs from phosphorescence, however,

is that the electronic energy transition that is responsible for fluorescence does

not involve a change in electron spin. This is not a trivial difference, and, as will

be discussed later in detail, has a great impact in the transition dynamics. The

theory presented in this chapter is largely based on the textbook of J. R. Lakowicz

[56], which is an excellent reference point for fluorescence.

2.2.1 Energy states

As quantum mechanical systems, molecules of fluorophores can only take on

certain discrete values of energy. These electronic states and the transitions be-

tween them are usually depicted using Jablonski diagrams, which are line dia-

grams that group the states vertically by energy and horizontally by spin multi-

15
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

plicity. A typical Jablonski diagram is shown in Figure 2.4 where radiative transi-

tions are shown by solid straight arrows and non–radiative transitions by dashed

arrows.

vibrational relaxation, kic


LUMO S2
HOMO
internal
conversion, kic

inte
r
LUMO S1 cro system
ssin
g
HOMO
k
isc T1 LUMO
HOMO
fluorescence, kf
non-radiative
relaxation knr

triplet relaxation
excitation k0n

non-radiative

rescence, kph
phospho-
hva hvf
hvph

kT
V3
V2
LUMO S0 V1
HOMO V0

Figure 2.4: A typical Jablonski diagram showing the singlet ground state (S 0 ), two ex-
cited states (S 1 and S 2 ) and the lowest energetic triplet state (T1 ). For each
electronic state, four vibrational states (V0 – V3 ) are shown. Photon absorp-
tion of energy hv a can lead to excitation to state S n , at a rate of k 0n , where
n > 0 depends on v a as per equation (2.8). De–excitation mechanisms in-
clude vibrational relaxation and internal conversion (k i c ), fluorescence (k f ),
non–radiative processes (k nr ) and intersystem crossing (k i sc ), which leads
to build–up of T1 . T1 is subsequently depopulated by phosphorescence
(k ph ) or by non–radiative triplet relaxation processes (k T ). The arrows in
the boxes indicate the spin directions at each state.

Typically, fluorophores exist in the ground electronic state (S 0 ). Further, for

each discrete energetic state, the molecule may adapt several vibrational modes,

which are represented by the thin vertical lines in Figure 2.4. In room tem-

perature (T = 300 K) the thermal energy of molecules is E = k B T ≈ 0.026 eV,

where k B ≈ 8.6 × 10−5 eV/K is the Boltzmann constant. The energy separation

between vibrational modes is typically ∆E ≈ 0.19 eV, therefore, when consider-

ing the Boltzmann distribution that gives the probability of particles in a system

to be in a particular state:

E Vn − E Vn−1
µ ¶
P (Vn ) ∝ exp − . (2.7)
kB T

16
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

One can see that fewer than 1 in every 1,000 molecules will be at the first ex-

cited vibrational state V1 . Further, when considering the separation of electronic

states, which is in the order of a few eV, the probability of a molecule to be at S 1

in room temperature becomes vanishingly small. Therefore it is safe to assume

that all excitation events occur from S 0V0 .

The S 0 state is characterised by a fully occupied Highest Occupied Molecular

Orbital (HOMO) and an empty Lowest Unoccupied Molecular Orbital (LUMO).

When a system absorbs a photon, it gains energy and can enter an excited state,

S n , where n = 1, 2, 3... The exact state that the system will be excited to depends

on the wavelength (or energy) of the incident photon, hv a , which has to be equal

to the energy difference between the states (∆E ) according to:

∆E = hv = hc/λ (2.8)

where h is Planck’s constant, v the frequency, λ the wavelength and c the speed

of light in vacuum. The molecule can now be at different vibrational states of the

excited electronic state, however, fast vibrational relaxation (k i c ) quickly leads

them down to the lowest vibrational level of the S 1 state. From there, one way for

the system to relax to S 0 is by fluorescence, that is by emission of a photon (k f ).

Usually, the system returns to an excited vibrational state of S 0 , and internal

conversion events follow leading to complete de–excitation. The energy of the

emitted photon (hv f ) is lower than that of the absorbed one, and its exact value

can again be calculated using equation 2.8. Fluorescence is a random process

and can be described by a decaying exponential for the S 1 evolution:

µ ¶
−t
S 1 (t ) = S 0 exp (2.9)
τf

17
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

where τ f is the fluorescence lifetime. For commonly used fluorescent dyes, τ f

is in the order of ns (Table 2.2).

Other than fluorescence – which is a radiative transition – there exist a num-

ber of non–radiative transitions (k nr ). These include the relaxation of each ex-

cited state to its lowest vibrational level, which happens in the ps scale and is

termed vibrational relaxation, the internal conversion that occurs when a vibra-

tional state of an electronically excited state couples to a vibrational state of a

lower electronic state and intersystem crossing (k i sc ). The latter is the transition

to a state with a different spin multiplicity, for example from S 1 to T1 . T1 can

subsequently be depopulated by phosphorescence (k ph ) with the emission of a

photon (hv ph ), or by non–radiative triplet relaxation processes (k T ). Transitions

between states of different spin multiplicity are slower (Table 2.2) since they in-

volve a spin flip, as will be further discussed in the next section. It should also

be noted that it is always v a < v f < v ph as a result of the respective energy loses

before each emission.

Transition Time scale (s)


Absorption 10−15
Vibrational Relaxation 10−12 − 10−10
Internal Conversion 10−11 − 10−9
Fluorescence 10−10 − 10−7
Intersystem crossing 10−10 − 10−6
Phosphorescence 10−4 − 102

Table 2.2: Typical lifetimes of electronic transitions [56].

2.2.2 Singlet vs Triplet states

Singlet and triplet states differ in the electron configuration of their valence

bands – that is the highest energy levels where electrons can be found – as de-

picted by the rectangles beside each energy state in Fig 2.4. The multiplicity of

18
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

each state is defined by M = 2S +1, where S is the total spin number i.e. the sum

of all individual spins.

Singlet states (S n ) have the property that the electron spins are paired in such

a way that for each electron with an “up” spin (s = +1/2) there is an electron

with a “down” spin (s = −1/2); note that in the excited singlet states, the spin of

the excited electron remains paired with the ground state electron. These states

have therefore S = 0 and M = 1 (hence the name “singlets”). On the other hand,

triplet state molecules have an unpaired spin configuration that gives S = 1 and

M = 3.

The transitions between singlet and triplet states are said to be “quantum–

mechanically forbidden”, meaning that they have much lower probabilities to

occur compared to transitions within states of the same spin multiplicity. They

occur via intersystem crossing (k i sc ) which is a non–radiative process. The prob-

ability of intersystem crossing increases when the vibrational levels of the two

excited states overlap, since little or no energy must be gained or lost for the

transition to happen, and it is higher in molecules with large spin–orbit cou-

pling. Such molecules can experience high triple state build–up at relatively

lower irradiances, and are therefore suitable for TRAST monitoring.

Since another spin flip is required for the relaxation of triplet states, they are

much longer lived in comparison to the singlet excited states. Typical relaxation

times τT range from µs to s, versus singlet relaxation times which are in the ns

range. This stability is exploited in TRAST monitoring [48]. Fluorophores re-

laxing back to the ground state originating from a triplet state may emit weakly

(phosphorescence) or not at all, which is why triplet states are often called dark

states.

19
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

2.2.3 Spectra

Figure 2.5 shows the absorption and emission spectra of Eosin Y, a dye used

extensively in this thesis. It can be seen that the emission spectrum is shifted

to longer wavelengths than the absorption. This shift corresponds to the energy

dissipation due to internal conversion that occurs between the events of absorp-

tion and emission (Figure 2.4). The spectral shift between the maxima of ab-

sorption and emission is termed the Stokes shift – in honour of Sir George Stokes

who first observed this phenomenon – and is the key concept behind spec-

tral separation that is essential in most fluorescence–based techniques. From

a quantum mechanical perspective, each energy state can be described by a

wavefunction. The absorption maximum represents the most probable tran-

sition, that is the excitation from the lowest vibrational mode of the ground

electronic state, S 0V0 , to that vibrational level of S 1 , S 1Vn , whose wavefunction

overlaps the most significantly with S 0V0 . This is the so–called Franck–Condon

principle, a rule that states that during an electronic transition, the likelihood

for a change from one vibrational level to another will be proportional to the

square of the overlap integral of the initial and final vibrational wavefunctions.

Emission
COO- Absorption

Br Br
Intensity (AU)

O- O O
Br Br

200 300 400 500 600 700 800


Wavelength (nm)

Figure 2.5: Absorption and emission spectra of Eosin Y dissolved in basic ethanol (data
from ref. [57]). Inset: Chemical structure of Eosin Y.

20
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

It can also be seen that the spectra are continuous, which might appear to

be in disagreement with the image painted by the Jablonski diagrams that show

distinct, quantised energy levels. This apparent discrepancy, however, can read-

ily be explained when considering the fact that organic dyes are large molecules

and thus have plenty of vibrational modes between which transitions can occur

(Figure 2.5:inset). Additionally, the individual line spectra are usually not seen

because of thermal (or Doppler) broadening of the spectra at room tempera-

tures where experiments typically take place.

Another two interesting and important characteristics of fluorescent spectra

are that the emission spectra are typically independent of the excitation wave-

length, and that the fluorescence spectra are approximately the mirror image of

the absorption spectrum. The former results from the fact that all fluorescent

relaxation events originate from the lowest vibrational level of S 1 . This allows

us to acquire the full fluorescence spectrum even when a source of different

wavelength than that of maximum absorption is used. The latter reflects the

fact that the vibrational modes are approximately equally spaced at the differ-

ent energetic states, and that Franck–Condon principle is applied both for the

absorption and the emission.

2.2.4 Quenching by Oxygen

Fluorescence quenching refers to any process that decreases the fluorescence

intensity of a sample. Quenching phenomena can be classified as static or dy-

namic (also referred to as collisional). In static quenching, the fluorophore and

quencher form non–fluorescent complexes while the fluorophore is in the ground

state. In contrast, dynamic quenching refers to energy transfer that occurs while

the fluorophore is in the excited state, and thus creates a new, non–radiative

pathway for de–excitation. Therefore, dynamic quenching affects the lifetimes

21
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

whereas static does not. This thesis focuses on dynamic quenching events, for

which to take place, the fluorophore and the quencher must come in contact.

Therefore, the quencher must be able to diffuse to the fluorophore during the

lifetime of the excited state.

The relationship between the concentration of a quencher and the fluores-

cence intensity is described by the Stern–Volmer equation:

If 0
= 1 + k q τ f 0 [Q] (2.10)
If

where I f 0 is the intensity without and I f with a quencher, k q is the bimolecu-

lar quenching constant, τ f 0 the fluorescence lifetime of the fluorophore in the

absence of quencher, and [Q] is the concentration of the quencher.

The Stern–Volmer equation (2.10) can be derived as follows (after [56]). The

fluorescence intensity at absence and presence of the quencher (I f 0 and I f , re-

spectively) are proportional to the population of the excited state S 1 at absence

and presence of the quencher, respectively, that is I f 0 ∝ S 1abs and I f ∝ S 1pr s .

When considering continuous excitation, f (t ) ∝ k 01 , a population equilibrium

is established: d S 1 /d t = 0. Therefore, we have:

d S 1abs
= f (t ) − (k f + k nr )S 1abs = 0 (2.11)
dt

in the absence, and,

d S 1pr s
= f (t ) − (k f + k nr + k q [Q])S 1pr s = 0 (2.12)
dt

in the presence of the quencher. Combining equations (2.11–2.12), it follows

that:
S 1abs (k f + k nr + k q [Q]) k q [Q]
= = 1+ (2.13)
S 1pr s (k f + k nr ) k f + k nr

22
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

which can be rewritten in the form of (2.10), given that the rate of decay at

the absence of quencher is k f + k nr = τ−1


10 . The bimolecular quenching con-

stant is a measure of the accessibility of the fluorophores to the quencher and is

given by k q = f Q k 0 where f Q is the quenching efficiency and k 0 is the diffusion–

controlled bimolecular quenching constant.

One of the most potent quenchers in nature is molecular oxygen. It has a

unique property that its ground state is a triplet state while its lowest available

excited state is a singlet state. The exact mechanism of quenching by oxygen

remains a subject of debate. Reports have suggested that it occurs by mixed

mechanisms that include intersystem crossing, charge transfer, and electron

exchange [56]. The most probable explanation, however, is that when a fluo-

rophore which is excited in a singlet state encounters a triplet oxygen molecule,

a change of its excited singlet state to an excited triplet state occurs. Since the

triplet states are also quenched by oxygen, and are long–lived, they are likely to

be quenched again by the same agent, driving the molecules to the ground state

via a non–radiative path.

For oxygen, it is f Q ≈ 1 and we can relate the bimolecular quenching con-

stant to the diffusion coefficient via the Smoluchowski equation:

k q = 4π(R f + R q )(D f + D q )N A (2.14)

where the sum of R f and R q (the molecular radii of the fluorophore and quencher,

respectively) defines the collision radius, D f and D q are the diffusion coeffi-

cients of the fluorophore and quencher, respectively, and N A is Avogadro’s num-

ber. The diffusion coefficient can be calculated from Stokes–Einstein equation:

D = k B T /6πηR (2.15)

23
CHAPTER 2. PHOTOPHYSICS AND PHOTOCHEMISTRY

where k B is Boltzmann’s constant, η is the solution viscosity and R is the molec-

ular radius.

Dynamic quenching is an additional rate process that depopulates the ex-

cited state, therefore, the lifetime in the presence of quencher becomes τ =

(k f + k nr + k q [Q])−1 . It can readily be shown that an equivalent decrease in flu-

orescence intensity and lifetime is expected to take place. Considering the first

half of equation (2.13), we have:

S 1abs τ−1
= (2.16)
S 1pr s τ−1
0

where τ0 is the lifetime in the absence of quencher. Equation (2.16) can take the

final form of:


If 0 τf 0
= . (2.17)
If τf

Apart from excited singlet states, triplet states can also be quenched by oxy-

gen. In fact, when considering the diffusion coefficient of O2 in H2 O (≈ 2 ×

10−6 cm2 s−1 ) and the usual concentrations of dissolved O2 in air–saturated H2 O

(≈ 30 µM) one can see that, with a typical value of k q ≈ 1010 L mol−1 s−1 [58],

we have k q [O 2 ] ≈ 3 × 106 s−1 . This time scale is orders of magnitude slower than

the fluorescence lifetime (which is a few ns). We can therefore assume that oxy-

gen is a much more competitive quencher of the slower triplet states instead,

and it is possible to re–write the Stern–Volmer equation (2.10) in terms of triplet

lifetimes as:
τT 0
= 1 + k q τT 0 [O 2 ] (2.18)
τT

where τT 0 and τT are the triplet lifetimes in the absence and presence of oxygen,

respectively.

This relationship between the triplet relaxation time and molecular oxygen

is the key concept behind the development of our protocol for quantification of

oxygen concentrations, as will be discussed in Chapter 5.


24
Chapter 3

Biofilms

The present chapter aims to introduce the reader to the biofilm concept. A his-

toric overview is given, followed by discussions on the importance of two fac-

tors, namely oxygen concentration distribution and mechanical properties, on

the growth and lifecycle of the biofilms.

3.1 The biofilm mode of life

Biofilms are structured bacterial communities in which cells are held together

within a matrix formed by self–secreted polymeric compounds, termed the Ex-

tracellular Polymeric Substances (EPS). They are considered to be one of the old-

est, most successful and widespread form of life on Earth [37]. The first recorded

microscopic observation of such ensembles dates back to the 17th century when

van Leewenhoeck experimented with plaque samples taken from human teeth

[4], however, it was only in 1978 when the term biofilm was coined by Bill Coster-

ton [59], who is considered the founding father of the field [60].

Biofilms have a deleterious impact on many domains of life as they can affect

both human health and man–made systems. It is estimated that up to 80% of all

bacterial infections are biofilm related [40], including cystic fibrosis [61], endo-

25
CHAPTER 3. BIOFILMS

carditis [62], urinary tract infections [63], periodontitis [43], and other chronic

infections [64]. Additionally, many medical devices, especially catheters [63],

prosthetic heart valves, cardiac pacemakers, skull implants, cerebrospinal fluid

shunts and orthopedic devices, have been found to be easily colonized by biofilms,

leading to device–associated infections and urgent need for replacement [65, 66,

44]. Other sectors affected are the food and dairy industry [67], where biofilms

can cause food spoilage potentially leading to outbreaks of foodborne pathogens

[41], water distribution systems and plumbing [45, 68], metalworking, structural

and civil engineering where mechanical blockages, impedance of heat transfer

processes and increase in the corrosion rate of surfaces have been documented

[42]. On the other hand, biofilms can also serve beneficial purposes, and are

currently used for the treatment of drinking water, purification of wastewater

and detoxification of hazardous waste [69, 70].

Remarkably, it has been found that the genetic expression of bacteria which

are organised in a biofilm formation can differ from that of the same type bac-

teria that exist in a planktonic state [71, 72]. This is a key concept in under-

standing why remedies developed against planktonic bacteria often fail to work

effectively against biofilms [73, 74], and highlights the importance of studying

the dynamics of formed biofilms rather than individual bacteria.

Another unique feature of biofilms is the ability of their member cells to

communicate with each other and regulate the expression of specific genes once

thresholds in cell densities are reached to promote phenotypes that are the most

beneficial for the entire biofilm. This process is known as quorum sensing [38,

75], and it it because ot it that biofilms are able to survive environmental stress,

such as temperature or pH changes, desiccation or pressure which would oth-

erwise destroy these cells in their planktonic state [39].

26
CHAPTER 3. BIOFILMS

(a) (b)

Figure 3.1: Widefield images of Pseudomonas aeruginosa bacterial cells forming a


colony on a coverslip. (a) Image taken 24 hours and (b) 90 hours post–
inoculations. Scalebar: 2 µm.

Biofilm formation is a dynamic process during which cells change and adapt

the regulation of motility and the production of different EPS components that

affect the physico–chemical properties of the cell surfaces, hence determining

the adhesiveness of the cells [76, 77, 78]. In general, the life cycle of biofilms

is considered to begin when free–floating, or planktonic, bacteria encounter

and attach to a surface (Figure 3.1a). EPS production starts, and the emerg-

ing biofilm community develops a complex, three–dimensional structure [79].

During this stage of maturation, biofilms can take many forms, including pillar–

and mushroom–like microcolonies (Figure 3.1b, top–view), filamentous stream-

ers as well as undifferentiated layers of cells [80]. In the latest of stages of the

biofilm lifecycle, propagation is done through detachment of clumps of cells,

or by release of individual cells. This is known as dispersal, a process that in-

volves bacteria swimming away from the interior portions of colonies, leaving

behind hollow structures (Figure 3.2). The single or clusters of escaped cells are

then able to attach on a surface and create new colonies, or merge with existing

colonies downstream [75, 81].

The dispersal can be triggered in different ways by a number of metabolites.

For example, it has been found that biofilms undergo dispersal in response to a

27
CHAPTER 3. BIOFILMS

sudden increase in nitric oxide [82], oxygen depletion [83], as well as in response

to either a sudden decrease or increase in carbon substrate availability [84]. At

the dispersal stage, detachment of either single motile cells, or sloughing off of

large aggregates of cells happens [85, 81]; this leaves behind empty shell–like

colonies [86, 87] while at the same time, the single cells and multicellular aggre-

gates that dispersed can go on to initiate new biofilms downstream [81].

A number of factors affect the extend of biofilm formation. Biofilms can be

developed under diverse nutrient conditions, from abundant to almost non–

existent. They are more dense in nutrient–rich environments which promote

the transition of bacterial cells from planktonic to biofilm state [70]. Similarly,

pH plays a role in the formation of EPS. For the majority of bacteria, the optimal

pH value is around 7, yet it varies between species [88]. Lastly, the surface of at-

tachment greatly influences the proliferation of biofilms. For example, surface

roughness at microscale levels enhances the adhesion of bacteria by providing

more surface area for cell attachment in the initial phases, and reduces the shear

force experienced by bacterial cells that grow in the presence of liquids flowing

at high flow rates [88]. Other factors such as charge, hydrophobicity and elastic-

ity of the surface are also influential in microbial attachment [89].

In nature and pathology, biofilms can comprise cells from up to several hun-

dred different species [231, 90]. Undoubtedly, in such systems, the interspecies

interactions become important and are expected to affect the biofilm structure

and functionalities. However, given the complexity and the exponential increase

in possible combinations as the species number increases, most biofilm studies

are traditionally carried out in single–species models.

Likewise, in this thesis, biofilm colonies of Pseudomonas aeruginosa are cul-

tured and studied. P. aeruginosa is a common Gram–negative, opportunistic,

rod-shaped (length ≈ 1.5 µm), biofilm–forming pathogen bacterium that can

28
CHAPTER 3. BIOFILMS

thrive both in aerobic and hypoxic conditions. P. aeruginosa can grow as a biofilm

of considerable thickness (up to hundreds of µm) and can secrete large quanti-

ties of polysaccharides which surround the cells by forming a glycocalyx [91].

P. aeruginosa biofilms may exhibit up to 100 times greater resistance to biocides

than the corresponding planktonic cells [92]. It is one of the most widely studied

systems as it is considered a model organism for investigating biofilms [85].

3.2 EPS and mechanical properties of biofilms

Biofilms have been metaphorically called “cities of microbes” [93]. EPS, which

accounts for 50 to 90% of their total organic matter [44], is comprised of polysac-

charides (long polymeric chains that usually represent the largest component of

EPS [94]), various proteins, glycoproteins, glycolipids, and in some cases, extra-

cellular DNA (e–DNA) [95]. At the most basic level, EPS can be thought of as the

key component for maintaining the biofilm architecture and cellular organisa-

tion [96, 97, 98]. Getting back to the analogy of a city, it is not surprising that EPS

has already been termed the “house of the biofilm cells” [95].

The role of EPS is neither static, nor limited to providing structural support

to cells. It determines the porosity, density and water content [96, 37, 97]. The

matrix is also responsible for fortifying the biofilm against desiccation and en-

vironmental stresses [99, 100], providing antimicrobial resistance and tolerance

[39] and making biofilms robust towards chemical perturbations [101].

Another key function of EPS is to facilitate resource capturing by passive

sorption, which gives biofilms a phenotypic advantage over the planktonic mode

of life [39]. This process involves a continuous exchange of nutrients, gases and

other molecules between the environment and biofilms, leading to the estab-

lishment of chemical gradients that facilitate the transport of solutes, nutrients

and small molecules (e.g. oxygen) in and out, as well as within the biofilm [102,

29
CHAPTER 3. BIOFILMS

81, 39]. These gradients have been observed in biofilms regardless of their age

and size [103], and become more prominent in multilayered biofilms. For ex-

ample, in aerobic biofilms, depletion of oxygen in the lower layers of the biofilm

leads to stratification of organisms according to local oxygen availability. [37]

Similarly, nutrient availability gradients can lead to different growth states in

different parts of the biofilm, as starvation of organisms in the lower layers oc-

curs due to the nutrient consumption by organisms in the upper layers [37].

In EPS, different polysaccharides are responsible for different processes; for

instance, adhesion has been linked to the polysaccharides Pel and Psl [79], and

structural stability and cohesion to the polysaccharides of algae, because of their

binding affinity to Ca2+ or Sr2+ ions that form cross–linked polymer networks

[104]. These three exopolysaccharides are produced by P. aeruginosa bacte-

rial cells, and have been identified within the matrix of formed P. aeruginosa

biofilms [80].

The production rate of each polysaccharides can play an important role in

the morphology and the mechanical properties of the colonies [79], and, re-

markably, it has been found that the matrix changes its properties through-

out the lifecycle of the biofilms in ways that benefit the microbial population

[80, 85, 105]. For example, in the early stages of P. aeruginosa biofilm develop-

ment, Psl is anchored on the cell surface in a helical pattern promoting cell–cell

interactions to facilitate the assembly of a matrix that holds bacterial cells in the

biofilm and on the surface [85]. During the maturation stage, Psl has been ob-

served to localise in the periphery of the colonies, facilitating the formation of

matrix cavities that prepare the biofilm colony for seeding dispersal [85, 84] and

eventually leading to the formation of hollow colonies [86, 87].

Being linked with such a large number of functionalities, it is not surpris-

ing that the mechanical properties of the EPS have been a subject of interest for

30
CHAPTER 3. BIOFILMS

Cells EPS

i ii iii

Figure 3.2: Schematic illustrating the late stages of the lifecycle of a biofilm, starting
from a compact colony (i), evolving to a mature colony that facilitates seed-
ing dispersal (ii), and finally to a post–dispersal colony (iii).

many years. In general, biofilms are considered as viscoelastic materials, mean-

ing that they undergo both reversible elastic responses as well as irreversible

deformation, depending on the forces acting on the EPS matrix [106, 101]. The

environmental conditions under which biofilms grow also a role; more specifi-

cally, biofilms grown under higher shear stress are more strongly attached and

cohesively stronger than those grown under lower shear stress [106].

A number of techniques at different scales have been employed, such as

Atomic Force Microscopy (AFM) [107, 105], magnetic tweezers [108], tensile test-

ing using micro–cantilevers [109], microindentation [110], small–scale bulk rheom-

etry [98], microbeads AFM [111], single particle tracking [112], video particle

tracking [80] and others [113, 94, 100]. However, the aforementioned meth-

ods have some limitations: they either require direct interactions between the

sample and a probe; they are able to measure only in discreet positions and

only near surfaces; or they are suitable only for bulk measurements. There-

fore, their applicability in imaging thick, living, three–dimensional samples is

limited, even more so when specimens are grown under continuous flow [105].

Hence, there is a great need of developing a method that can assess stiffness in

living samples non–invasively, in real time and in three dimensions. With that in

mind, we have here applied confocal Brillouin microscopy, a technique capable

31
CHAPTER 3. BIOFILMS

of probing the mechanical properties of specimens by measuring the frequency

shift of light due to thermal fluctuations [47, 114], to investigate the stiffness of

P. aeruginosa biofilms along different stages of their lifecycle.

3.3 The role of oxygen in biofilms

Oxygen is required for respiration in aerobic organisms, such as P. aeruginosa,

and, being the major electron acceptor during cellular respiration, it plays an

essential role in the generation of energy that is necessary for cell maintenance

and growth [115, 116]. The distribution of oxygen within biofilms can be het-

erogenous and may result in bacterial cells exhibiting distinct physiological states

depending on their local oxygen availability [39].

During the initial stages of the biofilm life cycle, the base of the biofilm is

usually aerated. However, as the biofilm develops into larger aggregates, respi-

ratory activity rapidly consumes oxygen, which then becomes depleted on the

interior of the biofilm [117, 119, 118] as the consumption of oxygen by aerobic

organisms in the higher layers of the biofilm is faster than the rate of diffusion

[39]. This dynamic behaviour leads to the formation of oxygen gradients, which

can be relatively stable, or fluctuate and move vertically over time with time

scales of seconds, minutes or hours [115]. Oxygen limitation significantly affects

metabolic activities. For some bacteria, oxygen depletion has been identified as

an environmental cue for seeded dispersal [87, 120, 83], which often occurs from

the centre of the biofilm, in agreement with the understanding that oxygen be-

comes limiting in the interior of biofilms. Lack of oxygen has also been linked

with changed physiology, where some organisms may grow fermentatively, oth-

ers may switch to alternative electron acceptors and some bacteria may enter a

stationary phase and show little cellular activity [44].

32
CHAPTER 3. BIOFILMS

Traditionally, measurements of oxygen have been carried out using Clark

type electrodes [121]. The main limitation of using electrodes, however, is that

they consume oxygen while measuring [122]. Additionally, measurements with

microsensors are invasive; their tips – which can be up to 10 µm in diameter

[119] – can obstruct the local flow field and alter the micromechanical proper-

ties of the sample leading to artefacts [123].

Another approach is the use of optical methods, which do not consume oxy-

gen and are able to image oxygen concentrations over large areas [124]. Such

techniques are based on the principle of dynamic quenching of luminescence

of an indicator in the presence of oxygen [125, 56, 126]. The simplest applica-

tion comes in the form of planar optodes [119], which have been used to obtain

oxygen profiles of various types of biofilms, such as tap water biofilms [126],

biofilms formed by an activated sludge from a waste water treatment plant [119]

and biofilms of P. putida [128]. One of the shortcomings of optodes is that they

often lead to erroneous measurements due to the heterogeneity in the excitation

light field. To overcome this, ratiometric [129] or lifetime [130] measurements

often need to be employed. Another important limitation is that such devices

can only measure close to the surface of the sensor, limiting their applicability

in thick samples such as biofilms.

Hence, there is a need for developing tools to measure oxygen dynamics

in three–dimensions without depth limitations in order to map the profile of

an entire biofilm structure. In this study, we have used TRAST monitoring by

time–modulated illumination, which is a technique that can measure the rel-

ative populations of fluorophores in singlet and triplet (transient) states, and

their transition rates [48, 49]. This is done by measuring and integrating the

fluorescence emission intensity while exciting the sample with trains of pulses

of different characteristics. We have also developed a protocol that allows faster

33
CHAPTER 3. BIOFILMS

acquisition and can provide absolute values of local oxygen concentrations, which

was validated in different microenvironments.

The method was here combined with SPIM, an imaging approach commonly

used in thick biological samples as it allows for unparalleled optical slicing and

decreased phototoxicity [131]. Using this approach we were able to identify ar-

eas of different oxygen concentrations inside biofilm colonies of P. aeruginosa.

34
Chapter 4

Confocal Brillouin Microscopy for

Mechanical Measurements

This chapter introduces the application of Brillouin Spectroscopy on a Confocal

Microscope for measuring the stiffness of living biofilm colonies. Detailed in-

formation about the methods and techniques are presented, followed by a dis-

cussion of results from measurements on hydrogel gels and PA biofilms. Results

of this chapter have been partially published in ref. [132].

4.1 Introduction

4.1.1 Confocal microscopy

The concept of confocal microscopy was patented in 1957 by Minsky [16, 17],

who wanted to overcome the limitations of existing wide–field microscopes in

order to image dense, unstained neural networks by optical sectioning. The first

usage of the term “confocal”, however, came from Colin J. R. Sheppard and his

group, who published a theoretical analysis of confocal and laser–scanning mi-

35
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

detector

pinhole

dichroic mirror
Source

objective lens

focal plane
sample

Figure 4.1: Schematic outlining the key components and operating principles of a con-
focal microscope.

croscopes in 1977, and later developed microscopes with epi–laser–illumination,

stage scanning and photomultiplier tubes as detectors [18].

In a confocal microscope, light is emitted by a point source and is directed to

the sample. A pinhole1 is placed in front of the detector at an optically conjugate

plane with the sample and rejects out–of–focus light rays, which would lead to

image degradation (Figure 4.1). Confocal microscopy is able to produce thin

(0.5 to 1.5 µm) optical sections through thick fluorescent specimens.

Confocal microscopy is one of the most popular techniques currently used

in life sciences and cell biology due to the achievable image quality and the rel-

ative ease of implementation and operation [133].

The main drawback of a confocal microscope is the sequential, point by

point (usually by a scanning galvanic mirror) image acquisition, which results

in rather slow image acquisition times. Another consideration is the possible

increased photobleaching due to the continuous scanning. Those issues have

however been addressed by the development of spinning disk confocal microsco-

py, which allows fast sample scanning that decreases the total acquisition time
1
In certain applications, the pinhole is replaced by a single mode fibre.

36
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

and hence photobleaching [134], and by multi–photon microscopy that reduces

overall photobleaching by limiting it to the focal plane, even though bleaching

and higher–order photon interactions within the focal volume remain an issue

[135].

4.1.2 Brillouin spectroscopy

The theoretical background of spontaneous Brillouin scattering has been intro-

duced in Section 2.1.2; in short, when light interacts with density fluctuations

–termed acoustic phonons– inside a medium, there is an energy exchange that

causes a frequency shift in the scattered light.

Brillouin spectra yield information about the phonon’s properties, which can

be associated with the viscoelastic characteristics of the medium [114]. First,

information about the acoustic velocity and the viscoelastic coefficients of the

material can be inferred from the magnitude of the frequency shifts. Second,

the spectral line width can be related to the lifetime of the photon–phonon in-

teraction, which gives information about the viscous properties, and, last, polar-

ization of the scattered light can reveal information about the local anisotropic

properties of the medium. This thesis deals with measuring the magnitude of

spectral shifts for probing the local stiffness of the samples.

Brillouin scattering was first observed experimentally by Gross [136]. The

first experiments had been limited to compressed gases and fluids, however,

with the advent of laser sources in the 1960s, Brillouin spectroscopy started to

find a wider range of applications in solids, liquids and gases [137]. Since then,

the technique has been extensively used in material physics and environmental

sensing to study the viscoelastic and high–frequency viscous behavior of crys-

talline and amorphous solids [138, 139, 140, 141], polymers [142], as well as to

investigate the behaviour of elastic constants as a function of pressure [143] and

temperature [144].
37
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

Despite its rapid adaptation in those fields, Brillouin spectroscopy was not

equally well–embraced by biomedical imaging for reasons that will be discussed

later. The first such application is dated to 1980, when Vaughan and Randall

measured the elastic properties of the lens and cornea of the eye [55]. Even

though this study showcased that Brillouin spectroscopy is compatible with bi-

ological matter, the technique suffered from very long acquisition times and low

Signal–to–Noise Ratio (SNR). The major challenge in Brillouin scattering has al-

ways been isolating the Brillouin signal, which is on the scale of a difference of

a few pm, and orders of magnitude weaker than the elastically scattered light,

which makes conventional spectrometers unable to resolve it. This is partic-

ularly problematic in biological samples where significant elastic scattering is

commonly expected. Due to these limiting factors, early research was focused

only on ex vivo measurements on fixed samples that can remain unchanged for

long periods of time, such as collagen fibres [145, 146, 147] and bone [148].

Traditionally, high–finesse scanning Fabry–Perrot interferometers or angle–

dispersive etalons were used. These are optical devices that consist of a trans-

parent plate with two reflecting surfaces, or two parallel highly reflecting mir-

rors. As light travels through the paired flats, it is multiply reflected and pro-

duces multiple transmitted rays which are collected and focused to produce an

interference pattern. However, due to their inherent limitation in transmission

efficiency, etalons are too slow to provide a satisfactory SNR in most biological

samples [149]. In 1970, Sandercock showed that this shortcoming could be sig-

nificantly improved by a multipass arrangement [150]. Such designs have since

been used to a number of studies with biological interest, such as to observe

Brillouin scattering from single muscle fibres [151], to investigate thermal de-

naturation of lysozyme [152], and to determine the complete elastic moduli of

spider silks [153].

38
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

Up to this point, Brillouin spectroscopy had been limited only to acquisition

of spectra in point measurements. In 2005, however, Koski and Yarger intro-

duced the concept of Brillouin microscopy, which refers to the pixel–wise ac-

quisition of spectra while the sample is being scanned. In their first paper, they

presented a Brillouin image of a solid–liquid interface at a spatial resolution of

20 µm [47]. It should be noted, however, that the acquisition time was still too

high (10 s per pixel) to be applicable in dynamic biological measurements.

An important advancement came about three years later, in 2008, with the

implementation of Virtually Imaged Phase Arrays (VIPAs) in Brillouin micro-

scopy [154]. A VIPA is a solid Fabry–Perot etalon with a highly reflective front

surface, except for a narrow window through which the beam enters, while the

back surface is partially reflective. These devices were originally developed in

1996 [155]. In VIPA arrangements, the input beam is focused by a cylindrical

lens and enters the etalon at an angle through the aforementioned window. The

beam makes multiple internal trips by reflecting on the surfaces. An array of

output beams with increasing phase delays is produced. These beams interfere

at the output and components of different frequencies are emitted at different

angles. VIPAs can achieve large angular dispersion and low light loss leading to

faster data acquisition speeds as compared to conventional Fabry–Perot etalons.

Indeed, by using a VIPA, imaging of a mouse eye in situ became feasible with ac-

quisition time of only 500 ms [154].

Nonetheless, because the signal strength of the Brillouin scattered light is

small compared to the elastic signal, a single VIPA often does not allow clear

separation of the Brillouin peak from the background. A solution to improve the

suppression of the non–specific background that in turn permits clearer differ-

entiation of the Brillouin peak is cascading a number of sequential spectrome-

ters. For example, cascading two and three VIPAs in a cross–axis configuration

has shown a 55 dB and 80 dB SNR improvement, respectively [156].

39
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

A highly efficient Charge–Coupled Device (CCD) camera is commonly used

in conjunction to facilitate data acquisition, however, even in two–stage VIPA

systems, discrimination against strong light scattering can be difficult in thick,

opaque samples [157]. Therefore, additional methods for background clean-

ing have been developed and employed. These methods include molecular ab-

sorption cells (that can offer up to 50 dB suppression) [157, 158], coupling of

a Michelson interferometer [160], multi–pass Fabry–Perot interferometer [159],

tunable etalon–based notch filters [161] and common path interferometric fil-

tering [162].

All these improvements, together with background studies that showed the

applicability of high Numerical Aperture (NA) objective lenses [163], have al-

lowed Brillouin spectroscopy to find a number of applications in the life sci-

ences and medicine, both in tissue–level, such as for in situ and in vivo biome-

chanical imaging of the eye tissues [154, 156, 52, 164, 162], screening for in-

creased total protein in cerebrospinal fluid (CSF) during meningitis [165], study-

ing the viscoelasticity of amyloid plaques in transgenic mouse brain [166], and

quantification of plaque stiffness in arteries [167], and in single–cell or subcel-

lular level [168, 169, 170, 171].

4.1.3 The conventional use of term ’stiffness’

As discussed earlier in Section 2.1.3, the measured frequency shifts are con-

nected to the speed of sound in the material, and, in turn to the longitudi-

nal modulus as well as to other mechanical moduli. However, it is currently

not possible to convert the measured frequency shifts to absolute values that

are traditionally used in biomechanical measurements (as for example Young’s

modulus) for a number of reasons: First, equation (2.3) only holds for isotropic

materials, while, in anisotropic media, M is a tensor, that is, there exists a sep-

40
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

arate modulus in each direction, corresponding to direction–dependent hyper-

sound velocities. Second, the refractive index and material density are not al-

ways known in a pixel–wise manner, and local differences might exist. Third,

the relationship between macro– and micro–mechanical quantities is, in most

cases, complicated and very much dependent upon the medium being inves-

tigated, and, last, Brillouin microscopy probes a material at ultrafast GHz fre-

quencies, such that the viscoelastic behaviour may drastically differ from the

behaviour exhibited when the same material is probed by other methods that

operate in much slower scales, from seconds to minutes.

Unsurprisingly, results acquired by Brillouin imaging studies differ by sev-

eral orders of magnitude from those acquired by rheology, AFM or other es-

tablished techniques [156]. Some researchers have tried to correlate Brillouin

modulus with Young’s modulus, based on empirical observations [156, 168],

something that has led a number of studies to report Brillouin microscopy re-

sults using Pa scale [172, 170]. A more recent study attempts to shed some light

on what exactly Brillouin microscopy measures [173]. By investigating the fre-

quency shifts in differently prepared hydrogels, measured at different degrees

of swelling, the authors of the latter study suggest that the Brillouin modulus

is not directly related to the Young’s modulus, but rather to the polymer solid

fraction that is intrinsically coupled to Young’s modulus due to the mechanics

of swelling.

It is evident that, to date, there is a lack of a proper theoretical model and

fundamental understanding of the connection between Brillouin shift and elas-

tic modulii in biological material, however, it is the consensus that the Brillouin

shift is indicative of the stiffness of the probed material. Therefore, in this the-

sis, the raw measured Brillouin shift values (in GHz) are used, and, for simplic-

ity, materials with a larger Brillouin frequency shift will be referred to as “stiffer”,

similar to other studies in the literature [167, 169, 162].

41
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

4.2 Material and Methods

4.2.1 Experimental setup

Confocal microscope for Brillouin and fluorescence imaging The Brillouin

fluorescence confocal microscope was constructed using a commercial Olym-

pus IX–71 microscope platform. The laser source for Brillouin imaging is a Con-

tinuous Wave (CW) diode–pumped solid state laser (Cobolt 05–01 Jive) operat-

ing at 561 nm with a maximum power of 300 mW. The fluorescence excitation

laser is a diode laser (Cobolt 06–01 488) operating at 488 nm with maximum

power of 60 mW.

water cuvette
M filter

sample DM2 BS to filter


BK7
(b)

sCMOS
EF 2-VIPA spectrometer
DB

DB S2
Mask 2

BE
VIPA 2
Pinhole S1

Mask 1
Detector DM1 C2

488 nm VIPA 1
C1

(a) 561 nm (c)

Figure 4.2: (a) Confocal arrangement for Brillouin and fluorescence imaging. DM:
Dichroic mirror, BS: beam splitter cube, BE: beam expansion, M: mirror, EF:
emission filter, DB: achromatic doublet lens. (b) Schematic of the common–
path interferometric filter using BK7 glass slab. (c) Crossed-axis VIPA spec-
trometer. C1, C2: cylindrical lenses, S1, S2: spherical lenses, DB: achromatic
doublet pair.

Two laser beams were delivered to the microscope using optical fibres (kine-

FLEX), combined by a long pass dichroic mirror (DM1) (Chroma T550lpxr) and

then expanded by 6× beam expanders (BE) to fill the entrance pupil of the ob-

jective. The beams were split by a beam splitter (BS) (Thorlabs BS019) with a

power ratio of 30:70 (R:T). The transmitted beams were reflected by a mirror

42
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

(M) to a 10X microscope objective, which focuses the beams into the water cu-

vette that is used for calibration purposes. The reflected beams were directed to

the oil–immersion microscope objective lens (Nikon UPLFLN / 100X, NA = 1.3)

for the Brillouin and fluorescence imaging (Figure 4.2a).

Light collected from the sample was first spectrally separated by the dichroic

mirror DM2 (Chroma ZT488/561TPC), which transmits the Brillouin and Rayleigh

signals and reflects the fluorescence signal. Approximately seventy percent of

the Brillouin and Rayleigh signals were transmitted by the beam splitter (BS) and

coupled into the fibre to be delivered to the common path interferometric filter.

The fluorescence signal reflected by DM2 was further filtered by an emission

filter (Chroma ET525/50) (EM). An achromatic doublet lens ( f = 75 mm) (DB)

was used to focus light on a motorised pinhole (Thorlabs MPH16) that couples

the signal to a multi–mode fibre delivering the signal to the detector. A photon

counting detector (Hamamatsu H10682–110) together with a data acquisition

unit (NI USB–6351) were used to measure the fluorescence signal strength.

Common path interferometer A wavefront division interferometer was used

to suppress the elastically–scattered light, as presented in ref. [162]. More specif-

ically, light delivered from the microscope through the fibre was collimated to

achieve a beam diameter of 4 mm. A glass slab was aligned so that exactly half

of the beam was transmitted through the slab while the other half propagated

above the glass (Figure 4.2b). The optical path difference between the two parts

of the beam (δ) that depends on the angle between the prism and the beam (θi )

is given by:
θi2
à !
δ = d g (n − 1) 1 + , (4.1)
2n

where d g and n are the length and refractive index of the glass slab, respectively.

43
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

Destructive interference occurs when the two halves of the beam overlap

upon and couple into a single–mode fibre. The exact wavelength at which de-

structive interference occurs can be adjusted by titling the prism to vary the

angle θi , and therefore the interferometer can be tuned to suppress only the

Rayleigh scattered light. The transmission function is a cosine function with

a Free Spectral Range (FSR) of c/δ, where c is the speed of light in vacuum.

A d g = 30 mm long BK7 glass prism (n = 1.51) was used in our experiment to

achieve a FSR of 20 GHz. A maximum extinction of 40 dB can be achieved with

our 1 MHz linewidth laser source.

VIPA spectrometer A two stage, crossed–axis VIPA spectrometer, as shown in

Figure 4.2c, was used to extract the Brillouin frequency shifts in our experiments

[156, 169]. Both VIPA 1 and VIPA 2 were custom–built with a FSR of 39.5±0.5 GHz

(LightMachinery Inc, Ontario).

Light from the interferometric filter output fibre was collimated to produce

0.35 mm beam waist and focused by a 100 mm cylindrical lens (C1) onto the first

VIPA entrance window. A second 100 mm cylindrical lens (C2) was used to form

the spectrum at its focal plane. A mask (Mask 1) is used to shape the spectrum.

A 100 mm spherical lens (S1) was used to focus the light further onto the second

VIPA whose axis was orthogonal to the first VIPA. Another 100 mm spherical lens

(S2) forms the spectrum at its focal plane where a second mask (Mask 2) was

used to shape the spectrum. A pair of 100 mm achromatic doublets (AC508–

100–A–ML, Thorlabs) with separation of 10 mm between them (DB) was used to

magnify and image the spectrum onto a sCMOS camera (Andor Neo 5.5 sCMOS)

(sCMOS).

44
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

4.2.2 Sample preparation

Hydrogel beads

Calcium alginate hydrogels were prepared similar to the protocol described in

ref. [174]. In specific, sodium alginate (Acros Organics, U.K.) solutions (concen-

trations of 1% and 3%) in distilled water (dH2 O) were prepared and autoclaved.

Using a pipette, droplets of 10 – 20 µL of the prepared sodium alginate solution

were gently dropped into the wells of a 6 well plate (Nunclon Flat, Thermoscien-

tific) containing filter–sterilised CaCl2 (100 mM and 400 mM). The cross linking

was almost instantaneous. The drops were left to gelate and settle for 30 min

in CaCl2 , before they were washed twice with sterile dH2 O. All imaging exper-

iments were performed the same day. Beads were kept in sterile dH2 O before

imaging.

Biofilms

Bacterial strains and growth conditions P. aeruginosa strain PAO1 were grown

in Luria–Bertani (LB) medium at either room temperature or 37o C with the ad-

dition of agar, as required. GFP–tagged P. aeruginosa PAO1 was engineered fol-

lowing the procedure outlined in ref. [175, 176].

Flow cell Biofilm assay Biofilms formation under continuous flow was per-

formed as described in established protocols [177, 178], as shown in Figure 4.3a.

In specific, a flow cell system consisting of a 2 L medium bottle, Fluorinated

Ethylene Propylene (FEP) tubing (30 mm long, 2.1 mm ID and 2.3 mm OD) (Mas-

terflex, Germany), a bubble trap (Technical University of Denmark, Department

of Systems Biology), a flow cell (made by gluing – using Silicone glue (3M Super

Silicone Sealant Clear) – a 50 × 24 mm glass coverslip on a pre–fabricated poly-

carbonate base on which three channels were milled, each 1 mm deep × 4 mm

45
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

wide × 40 mm long) (Technical University of Denmark, Department of Systems

Biology) as shown in Figure 4.3b, and a 2 L waste bottle was assembled and used

in all experiments. Technical drawings and a protocol for constructing the flow

cell can be found at ref. [177]. A six roller peristaltic pump (Langer Instruments,

U.S.A.) was used to pump the medium through the biofilm flow cell at a flow

rate of 6 or 20 ml / h, for the experiments with standard and high flow, respec-

tively, giving rise to flow speeds of 0.042 and 0.14 cm s−1 , respectively. A stage

was specifically manufactured at Imperial College London for the flow cell to be

placed under the microscope (Figure 4.3c).


(a) Air filter

Microscope

Bubble
Peristaltic trap Flow cell Waste bottle
Medium bottle pump

40 mm

4 mm

(b) (c)

Figure 4.3: The continuous–culture biofilm flow cell setup used in the confocal Bril-
louin microscopy experiments of this thesis. (a) Schematic drawing (not to
scale). (b) Close–up photograph of flow cell. (c) The flow cell on the confocal
microscope.

The FEP tubings were autoclaved and primed with minimal medium ([15

mM (NH4 )2 SO4 , 34 mM Na2 HPO4 , 22 mM KH2 PO4 , 58 mM NaCl, 1 mM MgCl2 ,

0.1 mM CaCl2 and 0.001 mM FeCl3 ], supplemented with 2 g/L of Glucose plus

2 g/L of Casamino acids) before the start of every experiment. GFP–tagged P.

46
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

aeruginosa were grown overnight in 40 g/L LB medium (Difco) in a shaking in-

cubator at 37o C and 200 rpm. The culture was diluted to OD600 ≈ 2.0, in minimal

medium and 300 µL was injected into the FEP tubing near the entry port of each

of the channels of the flow cell. The inoculum was allowed to attach in the flow

cell for 1 h without flow, at room temperature. The flow of the minimal medium

was then resumed throughout imaging.

All protocols and risk assessments for developing biofilms in a flow cell were

approved by the Safety Department of Imperial College London.

4.2.3 Acquisition and analysis

Scanning and data acquisition was controlled by in–house developed Labview

(National Instruments, Austin TX, USA) software and data was analysed using

Matlab (The MathWorks Inc., Natick, MA, USA). Typical exposure times varied

between 500 – 900 ms for the Brillouin and 10 – 100 ms for the fluorescence

imaging. A 50 µm pinhole was used for the fluorescence imaging, which corre-

sponds to ≈ 2.25 Airy units2 .

Subsequently, as the sample was scanned, both the Stokes and anti–Stokes

Brillouin peaks were captured. Their positions (determined by fitting a Lorentzian

function on the spectrum) were used to calculate the frequency shift at every

point of the sample, as:

FSR − FxP × (B 2 − B 1 )
uB = (4.2)
2

where FxP is a coefficient that converts the spatial position along the dispersal

axis (in pixels) to frequency (in Hz), B 2 and B 1 are the spatial position of the

Stokes and anti–Stokes Brillouin peaks along the dispersion axis (Figure 4.4a).
2
d Airy = 1.22λ M/NA = 22.2 µm where λ ≈ 510 nm, M = 1 (magnification), NA = D/2 f = 0.28,
D is the 6× expanded beam (original radius 0.35 mm) and f = 75 mm the focal distance of the
achromatic doublet.

47
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

(a)
R
150
B1

Intensity (a.u.)
100

B2
50
R

0
10 20 30 40 50 60 70 80 90 100
(b) Number of Pixels

Figure 4.4: (a) Brillouin spectrum of water, measured by two-stage VIPA spectrome-
ter and interferometer. The Stokes and anti-Stokes Brillouin peaks (B) and
Rayleigh peaks (R) are shown. (b) The intensity profile along the dispersion
axis is fitted with Lorentzian curves to extract the positions of the Brillouin
peaks.

Brillouin microscope calibration At the start of every experiment the micro-

scope was calibrated using a distilled water sample placed in a cuvette. A typi-

cal acquired spectrum is shown in Figure 4.4, where the Brillouin and Rayleigh

peaks are marked with B and R, respectively. The Stokes and anti–Stokes Bril-

louin peaks are located with the software and their positions are used to define

the slope dispersion axis. The theoretical Brillouin frequency shift of water is

known to be 7.04 GHz for a 561 nm laser wavelength. Therefore, from the mea-

sured distance between the two peaks, a coefficient that relates frequency shift

and peak position in spatial domain FxP (Frequency per Pixel, in GHz) can be

calculated as:
FSR − 2 × 7.04
FxP = (4.3)
B2 − B1

where FSR is the free spectral range of the VIPA etalon, and B1 , B2 are the posi-

tions of the Stokes and anti-Stokes Brillouin peaks along the dispersion axis.

Zero–padding The Discrete Fourier Transform (DFT) converts a finite sequen-

ce of N samples of a recorded signal to an equally–sized sequence of samples in

the frequency domain. Zero–padding is commonly used in signal processing to

improve speed and precision of the results; it involves appending one or more

48
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

zeros in one domain and results in an increased sampling rate in the other do-

main [179].

A signal – here the spectrum image acquired by our camera – of N samples

can be expressed with an inverse DFT as:

NX
−1
f (m ? δx) = F (n ? δu) exp [i 2πmn/N ] (4.4)
n=0

where δx is the sampling distance in image space, δu is the sampling in Fourier

domain, f (m ? δx) is the spectrum image and F (n ? δu) is the discrete Fourier

transform of the spectrum image. It follows that:

1
δx = (4.5)
(N ? δu)

Padding M zeros at the borders of the transformed image in the Fourier do-

main allows us to obtain a denser sampling grid when applying the inverse DFT,

which improves the localisation of peaks when fitting the spectrum image with

Lorentzian curves to localise the Brillouin peaks (Figure 4.5).

This process does not improve the resolution as it does not produce new

information. However, it increases the number of points in the spectrum, as

they are directly proportional to the number of points used in the DFT. That is

essentially an interpolation in the image domain.

49
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

R 190 (b)
180
B1 170 Data points
Intensity (a.u.)

160 Lorentzian Fitting

150

140

130
B2 120

R 110

100

0 10 20 30 40 50 60 70 80 90 100
(a)
Pixel number along dispersion axis

20
(c)
19

18
Data points
17
Intensity (a.u.)

Lorentzian Fitting
16

15

14

13

12

11
0 50 100 150 200 250

Pixel number along dispersion axis

Figure 4.5: (a) Brillouin spectrum measured at a point inside a living biofilm. The Stokes
and anti–Stokes Brillouin peaks (B) and Rayleigh peaks (R) are shown. The
dispersion axis is marked by the red line. (b, c) Intensity profiles along the
dispersion axis, and Lorentzian fits, shown before (b) and after (c) zero-
padding.

50
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

4.3 Results

4.3.1 Measurements in hydrogel beads

In order to assess the applicability of the method in bacterial biofilms, we first

measured the Brillouin shift in cross–linked Alginate Acid (A. A.) – calcium chlo-

ride (CaCl2 ) hydrogel beads (Figure 4.6). Such hydrogels are commonly used as

model synthetic matrices, since alginates, which are abundant in biofilms, are

known to affect the mechanical properties of EPS in bacterial colonies [174, 180,

181].

Typical optical and Brillouin cross–sections of such a hydrogel bead (100

mM CaCl2 and 2% A. A.) can be seen in Figure 4.6c,d. The Brillouin shift mea-

sured in the region of interest (ROI) that is marked by the dashed line was 7.28 ±

0.03 GHz. Next, measurements were performed in beads prepared with a range

of biologically–relevant A. A. and CaCl2 concentrations. Figure 4.6e shows the

means of 10 technical replicate measurements that were performed at 3 differ-

ent points within each bead (total N = 30). The error bars represent the standard

deviations. It can be seen that the Brillouin shift was positively correlated to the

concentration of the A. A. and CaCl2 . An increase of A. A. concentration from

1% to 3% led to an increase in Brillouin shift from 7.02 ± 0.09 GHz to 7.51 ± 0.06

GHz when gelated in a 100 mM CaCl2 bath, and from 7.41 ± 0.07 GHz to 7.59 ±

0.05 GHz in 400 mM CaCl2 . All measurements were performed with the beads

immersed in sterile water.

4.3.2 Characterisation of the stiffness of P. aeruginosa biofilms

Previous studies have shown that P. aeruginosa biofilms form colonies that ex-

pand into pillar–like and mushroom–shaped colonies when grown and devel-

51
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

7.8
(a) (b) 7.4
7.0
6.6
GHz
7.8

7.6

Brillouin shift (GHz)


7.4

7.2

(c) (d) 7.0

6.8

6.6

1% A.A. 3% A.A. 1% A.A. 3% A.A.


(e)
GHz
100 mM CaCl2 400 mM CaCl2
7.10 7.15 7.20 7.25 7.30 7.35

Figure 4.6: (a) Photograph of a typical Alginate Acid (A. A.) – calcium chloride (CaCl2 )
hydrogel bead. (b) Low–resolution Brillouin image of such a bead (pixel size:
30 µm). (c) Wide field image and (d) Brillouin image taken at the interface
of a hydrogel bead (prepared by cross–linking of 100 mM CaCl2 and 2% A.
A.) and water. (e) Brillouin shifts measured in hydrogel beads of different
compositions. The error bars denote the standard deviations of 30 measure-
ments: ten technical replicates at 3 different points within each bead. Scale
bar on (b – d): 200 µm. Note that all measurements were performed with
the beads immersed in dH2 O to avoid dehydration, however, for illustration
purposes only, panel (a) shows a bead resting on a petri dish outside water.

oped in flow cells similar to the ones used in our experiments [178, 105]. Here,

biofilm colonies (or colonies) are defined the visually observable convex struc-

tures that rise above the surrounding biomass, characterised by a circumference

and height. A total of 24 such colonies of different sizes (average diameters rang-

ing from 25 µm to 95 µm) were imaged at various depths and at different time

points along their life cycle.

In all images, gradients of Brillouin shifts within the biofilms were observed.

In particular, two different patterns were seen: 21 out of the 24 imaged colonies

tended to have increasing Brillouin shifts and biomass towards their centres, as

shown by the Brillouin and fluorescence images, respectively (Figure 4.7:Panel I,

example). In the remaining 3 colonies, the interior appeared to be less stiff than

the periphery, or possibly hollow (Figure 4.7:Panel II, example).

In order to investigate a possible relationship between colony size and stiff-

52
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

I GHz
7.6

7.4

7.2

7.0
10 μm (b) 10 μm (c) 10 μm
(a)
6.8

II GHz
7.6

7.4

7.2

7.0
(a) 20 μm (b) 20 μm (c) 20 μm
(c)
6.8

Figure 4.7: (a) Widefield, (b) Brillouin, and (c) Fluorescence images. Panel I: Compact
colony taken 60 h post inoculation at a depth of 24 µm inside a 42 µm thick
biofilm. Panel II: Dispersed colony taken 100 h post inoculation at a depth
of 15 µm inside a 35 µm thick biofilm.

ness, we calculated the mean Brillouin shift of each imaged colony by averaging

the results within a region of interest (ROI) that encloses the whole colony. The

results are plotted against the mean diameter in Figure 4.8, where no correlation

between colony size and average Brillouin shift was found (Pearson correlation

coefficient r = −0.0179 and p = 0.926).

4.3.3 Brillouin shift measured at increasing depths

Next, we investigated the possible correlation between depth and stiffness. The

thickness of the colonies imaged in our study varied from 16 µm to 55 µm. By

convention, we define Z = 0 at the vertex of each colony and therefore z increases

with increasing depth towards the substrate. Figure 4.9 shows such a z–stack

in a 32 µm colony taken in a biofilm 80 h post inoculation. The Brillouin shift

increased towards the centre of the cross–section at each of the different depths,

as well as overall with increasing depth (Figure 4.9b).

53
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

7.8

7.6
Brillouin shift (GHz)

7.4

7.2

7.0

20 30 40 50 60 70 80 90 100
Mean diameter (μm)

Figure 4.8: Brillouin shift in colonies of various sizes. Data points denote the means,
and error bars the standard deviations from all pixels within the ROIs en-
closing the colony. ROIs drawn by visual inspection of the corresponding
widefield images. Biofilms grown under constant flow velocities of either
0.042 or 0.14 cm s−1 (circles and triangles, respectively).

Z = 4 μm Z = 9 μm
10 μm 10 μm 7.7

7.6
Brillouin shift (GHz)

7.5
Z = 14 μm 10 μm
Z = 19 μm 10 μm

7.4

7.3 (b)

(a) GHz 5 10 15 20
7.0 7.2 7.4 7.6 7.8 Depth (μm)

Figure 4.9: (a) Brillouin image cross sections at different depths inside a single colony
of thickness = 32 µm, taken 80 h post insoculation. (b) Mean and standard
deviation of the Brillouin shift at different depths, measured within the ROIs
marked in (a). Data points connected by lines to aid visualisation. Z in-
creases with increasing depth towards the substrate from the vertex of the
colony.

54
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

To extend the study into larger colonies, we had to modify the imaging proto-

col in order to decrease the acquisition time. Therefore, in the following, instead

of acquiring complete two–dimensional Brillouin images at different depths,

we performed repeated measurements at points along a cross section running

through the middle of the colony, at various depths. Figure 4.10a shows the re-

sults from the technical replicates when examining such a large colony (thick-

ness = 38 µm, imaged 72 h post inoculation); it can be seen that there is decrease

in Brillouin shift towards the middle of the cross section when imaging at the

depths of 36 and 25 µm (as indicated by the black and green lines), while there

was an increase in Brillouin shift towards the center of the cross section when

imaging close to the top of the colony (depth = 10 µm, red line). Such a profile

indicates the existence of a shell of higher stiffness that includes a softer core,

or a void (Figure 4.10e). It is worth noting that the Brillouin shift measured at

the points outside of the colony boundaries was higher than that of the medium

around the colonies (≈ 7.05 GHz). This observation can be attributed to swim-

ming bacteria (as observed in previous studies [75]), which can move into the

focal volume during the exposure time of 900 ms, or due to the flocculent na-

ture of the colony walls. Widefield images at different depths of the colony are

shown in Figure 4.10b-d.

4.3.4 Temporal changes of stiffness

All measurements in this study were done between 36 and 120 h post flow cell

inoculation. Earlier than that, the biofilms were too thin to be imaged, as the

Rayleigh background scatter proximal to the coverslip is too strong to suppress.

Figure 4.11 shows the means and standard deviations of the Brillouin frequency

shifts of all imaged colonies, plotted against the time post inoculation. Each

datapoint corresponds to a ROI that encloses the imaged cross–section of the

55
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

7.5 10 μm 10 μm

7.4
0 0
Brillouin shift (GHz)

7.3
(b) (c)
7.2
10 μm (e)
Depth: 10 μm
7.1 Depth:
36 μm 25 μm
25 μm 0
10 μm (a)
7.0 36 μm

-40 -20 0 20 40
(d) - 0 +
Cross section (μm)

Figure 4.10: (a) Brillouin shift measured along a cross-section of a colony (thickness =
38 µm, taken 72 h post inoculation), at various depths. Data points and
error bars represent the mean and standard deviations from 10 technical
replicates, respectively, at each point. The triangles on the top border de-
note the colony boundaries at each depth. (b-d) Widefield images of the
colony at different depths of 10, 25 and 36 µm, respectively. The white
dashed lines (drawn by visual inspection) enclose the in-focus area, which
is considered to define the boundaries of the colony at each depth. (e)
Schematic illustrating the imaged cross–sections. Darker colour indicates
increasing stiffness. Z increases with increasing depth towards the sub-
strate from the vertex of the colony.

colony. No correlation between the average Brillouin shifts within colonies and

time of imaging after flow cell inoculation was found (Pearson correlation coef-

ficient r = 0.234 and p = 0.212).

Next, we imaged a single colony at 48 h post inoculation, fixed the position of

the stage and repeated measurements of the same colony the following day (i.e.

at 72 h post inoculation). The Brillouin shifts were measured both days along a

line cross section at fixed depth at locations inside, near the border and outside

of the colony, and are shown in Figure 4.12a. It can be seen that the difference

in Brillouin shift between the periphery and the interior of the biofilm became

more pronounced at 72 h post inoculation. During this time, the colony had

grown in size, where the cross section increased from 55 to 74 µm (Figure 4.12b)

and the thickness from 33 to 38 µm.

56
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

7.8

7.6
Brillouin shift (GHz)

7.4

7.2

7.0

40 60 80 100 120
Time post-inoculation (hours)

Figure 4.11: Brillouin shift in colonies at various times post–inoculation. Data points
denote the means, and error bars the standard deviations from all pixels
within the ROIs enclosing the colony. ROIs drawn by visual inspection of
the corresponding widefield images. Biofilms grown under constant flow
velocities of either 0.042 or 0.14 cm s−1 (circles and triangles, respectively).

7.5 (b)

7.4
Brillouin shift (GHz)

7.3

7.2 0

7.1
Age:
48 h 10 μm
72 h (a)
7.0
-40 -20 0 20 40
Cross section (μm)

Figure 4.12: (a) Brillouin shifts measured along the cross–section of a colony at two con-
secutive days (at 48 and 72 h post inoculation). Data points and error bars
represent the mean and standard deviations from 10 technical replicates,
respectively, at each point. The triangles on top border denote the colony
boundaries at different days. (b) Widefield image of the colony taken 72 h
post inoculation. The white dashed lines enclose the in-focus area, which
is considered to define the boundaries of the colony at both days.

57
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

4.3.5 Effect of flow rate

During each experiment, the flow rate of the medium was kept constant through-

out the complete lifecycle of the biofilm. However, two different flow velocities

were tested. Most of the experiments (19 out of 24 imaged colonies) were per-

formed at 0.042 cm s−1 , while the remaining at 0.14 cm s−1 . We observed that

colonies which were grown under the faster flow setting were generally smaller

in diameter as well as flatter. However, in terms of stiffness, no effect of the flow

velocity on the average Brillouin shift of colonies was found, as can be seen by

comparison of datapoints that correspond to the two flow settings in Figures

4.8 and 4.11. The measurements done under the high flow rate were limited to

colonies imaged up to 70 h post inoculation, due to the high consumption rate

of the medium.

4.4 Discussions

In this chapter, we used Brillouin microscopy together with fluorescence imag-

ing to image P. aeruginosa biofilms and we investigated the effect of several fac-

tors including the size and thickness of the microcolony and flow rate. All mea-

surements were performed on living, active, samples of biofilms growing in a

flow cell of cross sectional area 4 mm2 under constant flow of medium.

Interpreting our results According to equation (2.3), the measured frequency

shift, u b , is linearly proportional to the average hypersound velocity, V , within

that volume and the refractive index, n. In turn, V is proportional to the ra-

tio of the Brillouin modulus, M , and inversely proportional to ρ, the material

density. Considering the above, it is evident that pixel–to–pixel variations of the

measured u b could arise from changes in either (or from any combination of)

58
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

n, M and ρ. First, let us address possible variations of n. Authors in a previous

study [182] have made the assumption that “since the biofilm has a high water

content, n can be assumed equal to 1.333”. In a later study, this was experimen-

tally confirmed as the refractive index of P. aeruginosa was measured n = 1.33 ±

0.17 and n = 1.348 ± 0.013 by optical density experiments and by using a refrac-

tometer, respectively [183]. Therefore, possible variations in the refractive index

(if any) would be within the detection limits of our spectrometer, as noted in a

previous Brillouin microscopy study [160].

Next, we discuss the dependence of Brillouin shift on local variations of den-

sity. To the best of our knowledge, no study on the density of P. Aeruginosa

biofilms with characteristics comparable to those of our study exists. However,

some estimations can be made on the basis of published results in similar sam-

ples. Using terminal velocity of particles, the density of a wastewater biofilm has

been calculated 1.14 g cm−3 [184]. In a later study, it was calculated that the wet

density of biofilms of mixed filamentous bacteria and bacillus ranges between

1 and 1.1 g cm−3 corresponding to dry densities between 0 and 0.05 g cm−3 , re-

spectively [185]. This is not very far from earlier studies that assummed that

biofilms have a wet density equal to that of water (1 g cm3 ), since they are well

hydrated [186] – it is now known that up to 90% of the wet weight of biofilms

is water[187]. So, even if we were to assume that the density of P. aeruginosa

is significantly higher than that of water, and that the measured changes in u b

result from variations in ρ alone, it would follow that the u b measured inside

biofilms would be smaller to that of water (u b water = 7.04 GHz), which is not

the case in our measurements. Therefore, the observed values of u b are propor-

tional to the Brillouin modulus M , which hence represent a measure of stiffness,

even though it does not seem to be possible to assign an absolute metric to M

without the exact knowledge of the local material density.

59
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

To further support our argument that local differences in u b do not reflect

density heterogeneities, we only found a single study that attempts to make spa-

tial differentiation of density within biofilms, in which the authors measured

1.001 – 1.003 g cm−3 at the top region of biofilms versus 1.01 – 1.02 g cm−3 at the

deeper regions [188]. That is a relative change of ≈1.3% that occurs over the spa-

tial extent of several hundred of micrometres to a few millimetres (their biofilms

were much thicker than those in our study). For reference, the z–stack shown in

our Figure 4.9 shows a 2.6% increase in Brillouin shift over only ≈ 20 µm of depth

gain, which, because it is twice the range of what was observed in ref. [188], is

unlikely to be due to density differences.

Therefore, as we have shown that the observed changes in u b cannot be ex-

plained by changes in density or refractive index, we infer that the profiles of

Brillouin shifts are indicative of the material stiffness; further, all our assump-

tions are based on using the raw Brillouin shifts without attempting to calculate

the hypersound velocity (which would require knowledge of n at each pixel) or

the Brillouin modulus (which would require knowledge of both n and ρ, at each

pixel).

Hydrogels as synthetic model material The applicability of Brillouin micro-

scopy on biofilms was first assessed by imaging cross–linked Alginate Acid (A.A.)

- calcium chloride (CaCl2 ) hydrogel beads, which are commonly used as model

synthetic matrices [174, 180, 181]. We were able to quantify and discriminate

between beads that were prepared by different concentrations of the reagents.

It should be noted, however, that biofilms are much more complex systems

than hydrogel beads, hence, the quantitative results obtained from the measure-

ments in the beads are not – and should not be – used as a calibration look–up

table to infer the composition of biofilms. Nevertheless, the increased degree of

heterogeneity that exists in biological samples does not invalidate our results,

60
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

since only the probed volume (confocal volume < 1 µm3 ) is considered uniform

in our measurements, and not the whole sample.

Characterisation of biofilms In this study, we captured biofilms during var-

ious stages of their development as they were imaged at different times post

inoculation. For most (≈ 88%) of the imaged biofilm colonies, the Brillouin

shift increased towards the centre of the colony, corresponding to growth on

the outer edges of the colonies, with a denser core [189]. These biofilms are

presumably in an earlier developmental stage. Contrastingly, we found some

colonies in which the Brillouin shift decreased towards their centres and within

such colonies, rotating bacteria were frequently observed. A possible explana-

tion is that these colonies are in transition to the dispersal phase [75, 85]. All

of the hollow colonies and colonies with interiors softer than their peripheries

were at least 45 µm in size (mean diameter). This is in agreement with previous

studies reporting that after the colony reaches ≈ 80 µm in diameter, the interior

disperses from a breakout point and leaves a nonmotile, empty colony behind

[190, 191].

Further, by investigated single colonies, we have shown that the Brillouin

shift of compact colonies increased with depth (Figure 4.9), however, no corre-

lation was found between the size of the colony and the measured Brillouin shift

when different colonies were analysed (Figure 4.8). This observation is different

from a previous study [105], however, measurements in our study were focused

on the interior of the colonies and include colonies of different thicknesses that

were imaged at different depths and at different time points after inoculation,

while the other study utilized AFM, which may primarily interrogate only the ex-

terior of the colonies. In agreement with a previous study [105], we did not find

a correlation between stiffness and flow velocity, as the mean Brillouin shifts of

colonies grown under constant flow at two different velocity settings were not

significantly different. Lastly, when we imaged a single colony for two consecu-
61
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

tive days we observed a profile was established where the interior of the colony

became less stiff than its periphery, possibly indicating a transmission to disper-

sal stage [85, 86]. Figure 3.2 (page 31) shows a model of the developmental cycle

of biofilms which our data supports.

Repeatability and sample size information In total, 24 different biofilm colo-

nies, selected randomly by visual inspection, were imaged. Only colonies that

had distinct boundaries, visible by widefield microscopy, were imaged at ran-

dom time points that ranged between 36 and 120 h post inoculation. The di-

ameters ranged from 25 µm to 95 µm and their thicknesses from 16 µm to 55 µm.

Either single two–dimentional cross–sections taken near the half-height of the

colony, or z–stacks at different depths were taken. Data shown in Figure 4.8

and Figure 4.11 consist of the complete pool of imaged biofilm colonies (N =

24), where 19 of them were grown under flow velocity of 0.042 cm s−1 and the

remaining 5 under 0.14 cm s−1 . Data shown in Figure 4.10 and Figure 4.12 rep-

resent the mean and standard deviations from 10 technical replicate measure-

ments at each point of the colony. Data shown in Figure 4.6 consist of 30 mea-

surements on alginate beads: ten repeated measurements taken at three differ-

ent points within each bead of different composition.

Limitations This study was restricted to measurements of colonies that had

distinct boundaries, visible by wide field microscopy. The ROIs used to calculate

the average Brillouin shifts were defined manually by visual inspection. The flat,

undifferentiated portions of the biofilms, or the biofilm substratum were not

investigated. The elastic background scatter near the coverslip was too strong

to suppress, hence only colonies that were at least 20 µm thick were imaged.

For the same reason, measurements very close to the substrate of thick biofilms

were not performed. Further, the exact age of each individual colony cannot

62
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

be precisely determined in our study, as the attachment of different colonies

might have occurred at different time points after the initial cell inoculation. In

terms of characterising the colony size, it remains unclear whether some of the

larger colonies should be treated as single colonies, or as aggregates of smaller

colonies, nonetheless, we believe that our classification is consistent.

4.5 Conclusion

We have demonstrated that Brillouin microscopy is suitable to probe the inter-

nal micromechanical properties of bacterial biofilms. The stiffness of P. aerungi-

nosa biofilm colonies growing in a flow cell was measured in vivo, without dis-

rupting the flow conditions or damaging the samples. To our knowledge, this is

the first study that has revealed internal patterns of stiffness in P. aerunginosa

bacterial biofilms at a µm scale. We have shown that the stiffness patterns dif-

fer between colonies, and areas of reduced stiffness can exist within large, over

45 µm in diameter, colonies. Such observations would not be possible by bulk

rheology methods – which are extensive used in studying the mechanical prop-

erties of biofilms – due to their averaging effect, nor by AFM, as it only probes

the external surfaces [113].

Thus, Brillouin imaging has significant potential in the field of microbiology

and in particular in mechanical studies. Given that it does not require labelling,

it can be applied to almost any biofilm, either formed in the laboratory or natu-

rally occurring. This means that biofilms can be repeatedly measured to obtain

dynamic information on the change in stiffness over time, or as a consequence

of exposure to different treatments or stressors, e.g. dispersal agents. For ex-

ample, it has been found that PA biofilms are able to efficiently recover from

mechanical damage, and that their recovery abilities are not influenced by most

of the common treatment agents [101]. Other studies have found an oscillatory

63
CHAPTER 4. BRILLOUIN MICROSCOPY FOR MECHANICAL MEASUREMENTS

growth and metabolic behaviour [77, 192]. In all cases, it is not clear how this

mechanical remodelling takes place from a biophysical point of view.

We believe that information stemming from Brillouin imaging can offer new

insights towards the qualitative and quantitative understanding of the mechan-

ics at the multicellular bacterial level that drives the biofilm mode of life. Bril-

louin microscopy can complement studies by rheology, fluid flow or molecular

transport. Such a combinatorial approach could be beneficial to the develop-

ment of drugs and drug delivery methods for the removal of biofilms, as it is

currently unclear which biofilm components are the most promising targets for

treatment [101].

Further studies should be done in colonies that undergo dispersion, since

it is at that point where drastic impacts of changes in the mechanical proper-

ties are expected. A number of interesting questions could be answered, for ex-

ample, does dispersion happen only beyond certain stiffness levels, and which

are the sites where dispersion begins. Both natural and laboratory–induced (by

nitric oxide, for example [82]) dispersions could be studied. Lastly, in the fu-

ture, Brillouin microscopy could work in tandem with other techniques, such as

FCS or anisotropy microscopy, that can provide information on local diffusivity,

porosity and viscosity, towards a better understanding of the role of mechanical

properties in signalling and transport of nutrients within biofilm communities

[193], or, with Raman spectroscopy for giving a reasonable chemical and biolog-

ical interpretation of the observed stiffness profiles [195, 194].

64
Chapter 5

SPIM – TRAST for Oxygen

Concentration Measurements

This chapter deals with the application of Transient State (TRAST) monitoring

on Single Plane Illumination Microscopy (SPIM) for measuring oxygen concen-

tration in living biofilms. The theory, concepts and protocols are presented, fol-

lowed by a discussion of findings of this study. Results of this chapter have been

partially published in ref. [196].

5.1 Introduction

5.1.1 Single Plane Illumination Microscopy (SPIM)

SPIM is a versatile tool for imaging complex, three dimensional biological sam-

ples. The unique characteristic of this modality is that it illuminates the sample

by a thin, diffraction–limited laser light sheet that propagates perpendicularly

to the detection axis. This is achieved by the use of two objectives, one for creat-

ing the light–sheet (termed illumination objective) and a second (termed detec-

tion objective) for the imaging of the illuminated plane onto the camera (Figure

65
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

5.1b). The light–sheet is usually created by a combination of a cylindrical lens

and an objective lens, or a fast–scanning laser. The thickness of the light sheet

is determined by the NA of the illumination objective. Typically, light sheets of

width approximately 1 µm are used, which can be created by illumination objec-

tives of low NA (0.2 – 0.5), and imaged with objectives of NA ≈ 1. Thinner light

sheets can either be created with higher NA objectives, which, however, have

smaller working distances possibly limiting their applicability, or by structured

illumination techniques, such as Bessel beams or lattices [197, 198].

The achievable resolution of SPIM is comparable to standard epi–fluorescence

microscopes, and three–dimensional imaging is possible by reconstructing the

projections of multiple views taken while rotating and translating the specimen

between observations. A rather good signal quality can be achieved with SPIM,

since no out–of–focus fluorescence is excited or detected [199]. Additionally, it

is well suited for deep imaging within organisms, and, since tissues are exposed

to only a thin plane of light, photobleaching and phototoxicity are minimal, thus

allowing for long term observations [24].

Initially, SPIM was mostly used in developmental biology due to its ability

to image whole, large samples, such as zebrafish embryos or flies over a large

number of days. Today, however, SPIM is routinely used in various other spec-

troscopic applications beyond imaging, such as for single molecule detection

studies [200], FCS and FCCS [201, 202, 203, 204], lifetime imaging [205] and

FRET [206, 207]. A comprehensive review of the method and a number of its

applications can be found at ref. [131].

Here, an in–house built SPIM system [202] is used and further developed for

TRAST imaging to be implemented. The choice of SPIM as an imaging modality

was driven by its capability to provide the desired optical slicing for thick sam-

ples such as bacterial biofilms, as well as the decreased phototoxicity that makes

SPIM suitable for long term imaging of living samples.

66
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

w T EMCCD
(1)

Iexc (2)
Iem

(3) w
(1) (2) (3)
(a) Pulse width
Det. obj.

(b)
488 nm

Software
function generator Beam expansion Cyl. lens Ill. obj.

Figure 5.1: (a) Schematic of the concept behind TRAST monitoring by modulated illu-
mination: increasing the duty cycle (a = w/T ) of a pulse train while main-
taining a constant total illumination time leads to a decrease of emitted flu-
orescence (Iem ), since molecules have less time to relax to the ground state
before the onset of the following pulse. (b) Diagram of a TRAST–SPIM setup,
not to scale. After [196]. © IOP Publishing. Reproduced with permission. All
rights reserved.

5.1.2 Transient States (TRAST) spectroscopy

TRAST imaging by modulated illumination is a recently developed technique

that allows kinetic information of transient states to be extracted, by measuring

the time–averaged fluorescence response of a sample illuminated with different

pulsed excitation patterns.

Prior to the development of TRAST imaging, triplet–states had been mea-

sured mainly by transient absorption spectroscopy, phosphorescence microscopy

or FCS, all of which have their own limitations. Transient absorption spectroscopy

uses a pump pulse to excite the sample, followed by a delayed probe pulse. By

recording the change in absorption spectrum as a function of time and wave-

length, information regarding the number of decay paths per wavelength as well

as the relaxation times of the decay processes can be revealed. However, this

technique is limited to mainly cuvette experiments, samples of high concentra-

tions (> µM), and requires relative complicated experimental setups including

fast pulsing lasers [208]. Phosphorescence microscopy is largely restricted to

67
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

carefully prepared and deoxygenized samples, since phosphorescence is typi-

cally much weaker than fluorescence, and is very susceptible to oxygen quench-

ing [209]. Lastly, FCS is able to extract triple state information from the fluctua-

tions generated by transitions to and from the dark triplet states of the fluores-

cent molecules, which appear superimposed on those due to diffusion. How-

ever, FCS is limited to measurements in samples of very low concentrations

(typically in the pM – nM range), requires bright dyes, and, since measurements

are time–resolved, fast and sensitive electronics are needed [210]. In contrast,

TRAST uses the time–averaged fluorescence recorded over long periods (ms - s),

circumventing both the need for high time–resolution or sensitive detectors, as

well as limitations in the dye concentrations.

In general, the excitation rate of a fluorophore is given by the product of its

excitation cross–section (σexc ) and the excitation intensity (I exc ) that is k 01 =

σexc I exc . Therefore, by manipulating the excitation intensity one can control

the population of molecules driven to the excited states. Under modulated ex-

citation a triplet–state population is built–up during the pulses, which then de-

cays in–between them. An equilibrium of the populations of energy states is

formed, which depends on the pulse train characteristics. More specifically, as

the duty cycle (i.e. the ratio of pulse width over period, that is α = w/T ) in-

creases, there is an increase in the triplet state population since fluorophores

do not have sufficient time to return to the ground state before the onset of the

next pulse. As fewer molecules populate the ground state at the onset of the

next pulse, less molecules are available to be excited to the singlet states and

fluoresce, leading to weaker recorded signals [48, 211], as illustrated in Figure

5.1a.

This expected decrease in fluorescence as a function of duty cycle has been

analytically derived for rectangular pulses [211]. It is therefore possible to use

68
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

this model for pixel–wise fitting on a series of fluorescence images taken with

increasing duty cycles in order to extract the triplet state dynamics, and more

specifically the intersystem crossing time and the triplet relaxation time. To

date, TRAST imaging has been implemented in a number of microscopic se-

tups. It was first validated on confocal microscopy [48] then on a total internal

reflection microscope [212] and most recently on a SPIM microscope [213].

Due to their long life times – which are in the order of µs - ms (Table 2.2)

– fluorophores in transient states are very likely to interact with their immedi-

ate environment [210]. Thus, triplet kinetics can change considerably due to

changes in accessibility of quencher molecules or micro–viscosities, rendering

TRAST imaging suitable for monitoring environmental changes [48, 49]. As a

proof of concept, it was shown that the changes in triplet–state parameters of

the fluorophore rhodamine 6G due to different concentrations of potassium io-

dide (KI) in aqueous enviroments can be measured with TRAST monitoring [48].

Later, Rhodamine 110 samples were measured in aqueous solutions at different

concentrations of dissolved oxygen [212], and, in a more recent study, the triplet

kinetic rates of a number of different bulk dyes incuding Fluorescein, Eosin Y

and Atto 495 were measured in fully aerated and deoxygenated conditions [213].

To date, not many applications of TRAST imaging of biomedical interest ex-

ist; the first one came in 2010, when Geissbuehler et al. measured the oxygen

consumption during the contraction of a single smooth muscle cell on a wide–

field optical setup [211]. Two years later, TRAST imaging was used to assess

the internalization and recycling of proteins in single living cells by identifying

the relative quantities of a transmembrane receptor based on the differences in

triplet lifetimes of two fluorescent species (autofluorescence versus label) [214].

Lastly, in 2014, TRAST monitoring was used to study patterns of altered oxygen

consumption in cancer versus healthy cells [215].

69
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

5.2 Material and Methods

5.2.1 Simplified energy system

In this thesis, a simplified, three–state energy system which can adequately de-

scribe most fluorophores is used. The Jablonski diagram of such a model is

shown in Figure 5.2. The ground state is considered to be the singlet S 0 . S 1

and T0 are the excited singlet and triplet states, respectively. The more higher–

energy excited states have been omitted, as their contribution is negligible [216].

The system of equations that describes such a three–level energy system is:

kisc=τ isc-1
S1
T0
-1 -1
k01=τ 01
k10=τ 10
k T =τ T-1
S0

Figure 5.2: Jablonski diagram for a simplified, three–state model such as the one con-
sidered in the present study.

 
P1 PT P0
 
 + − τ10 τT τ01 
P
 0  
 

d   
  P P P

 P 1  =  − τ101 − τi sc1 + τ010 (5.1)
 

dt    
   
PT
 
P1 PT
τ − τT
 
i sc

where P 0 , P 1 and P T are the populations of S 0 , S 1 and S T respectively, τ10 is the

lifetime for the singlet state relaxation, τi sc is the intersystem crossing lifetime

from the excited singlet state to the triplet state, τ01 is the excitation lifetime,

and τT is the triplet state lifetime. The solution of the above system is a second

order differential equation involving the four transition rates [48, 211]:

d 2PT τ10 τT d P T τ10 q T τT


µ ¶ µ ¶
1
τ10 τT + τ10 + τT + + 1+ + P T = q T τT (5.2)
dt 2 τ01 dt τ01 τ01 τ01

70
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

where q T = τ10 /τi sc is the triplet quantum yield.

5.2.2 TRAST monitoring theory

Equation (5.2) can be solved for the case of illumination with rectangular pulses

[211], and an expression for the time–averaged fluorescence emission intensity

of a fluorophore that is illuminated by an isodose of light delivered by a train of

pulses can be described by:

³ ³ ´´
eq
1 − exp − Tτ−w
 
τT P T exp(k w) − 1
T 2
I em = γ 1 − ³ ´ × , (5.3)
T −w
1 − exp − τT + k 2 w w

where w denotes the pulse width and T the period of the pulse train. Addition-

ally, k 2 describes the population rate of the triplet state after onset of excitation,
eq
P T is the relative triplet state population at equilibrium for constant excitation,

and γ is a scaling factor that depends on the concentration of fluorophores and

the detection efficiency, defined as:

τ01 + τ10 + q T τT
µ ¶
k2 = − (5.4)
τT (τ01 + τ10 )

eq q T τT
PT = (5.5)
τ01 + τ10 + q T τT

qf
γ = ξΓ (5.6)
τ01 + τ10 + q T τT

where
τi sc
qf = (5.7)
τi sc + τ10

71
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

is the quantum efficiency of the fluorophores, Γ is the concentration of fluo-

rophores and ξ is a conversion factor between the emitted intensity and the dig-

ital readout of the camera.

The time–averaged fluorescence intensity in a TRAST experiment depends

on all the transition rates, as seen on equation (5.3). One of the assumptions

made for the three–state model used is that the excitation time, τ01 , is limited

by the illumination intensity and the absorption cross–section of the molecule.

The excitation rate for a particular dye and experimental settings, is therefore

considered constant and can be calculated by:

hc
τ01 = (5.8)
λP l aser σ(λ)

where h is the Planck constant, c the speed of light, λ the excitation wavelength,

P l aser the excitation power and σ(λ) the wavelength–dependent molecular ab-

sorption cross–section. Using equation (5.8) and for laser powers that match

our experimental setup, τ01 for Eosin Y is calculated to be in the range of a few

ns, in agreement to the previously measured value of 1.55 ns [217].

On the other hand, the triplet relaxation rate and intersystem crossing are

more difficult to define, and, as discussed earlier, depend largely on environ-

mental factors such as the availability of quenchers and the viscosity of the sol-

vent. The triplet relaxation time for Eosin Y has been found to change from ap-

proximately 2 to 30 µs between a completely aerated and deaerated solution,

respectively [212, 213]. Likewise, a range for intersystem crossing times can

be found in literature – from approximately 2 ns to 5 ns [212, 218, 219, 213].

It should be noted that both the measured triplet relaxation and intersystem

crossing times additionally depend on the experimental method.

72
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

1 1
τΤ (μs) τisc (ns)
2 0.8 2
0.8 4

Intensity (AU)
3
Intensity (AU)

8
10 4
0.6 14 0.6
20 5

0.4 0.4

0.2 0.2
(a) (b)
0 0
0 10 20 30 40 50 0 10 20 30 40 50
Pulse width (μs) Pulse width (μs)

Figure 5.3: (a) The effect of triplet relaxation time on TRAST curves. Parameters used
in modelling: T = 50 µs, τ01 = 2.2 ns, τi sc = 2.5 ns, τ10 = 3 ns [213]. (b) The
effect of intersystem crossing time on TRAST curves. Parameters used in
modelling: T = 50 µs, τ01 = 2.2 ns, τT = 3 µs, τ10 = 3 ns [213].

The effect of triplet relaxation and intersystem crossing time on the appear-

ance of TRAST curves is shown in Figure 5.3, where out of the two parameters,

it can clearly be seen that triplet relaxation time plays a much more crucial role

in the intensity of fluorescence under isodose pulsed illumination. The range of

values modelled for the two parameters is experimentally–relevant, and covers

the cases of deaerated to fully aerated aqueous samples. For each plot, all the

other parameters apart from the investigated one were considered to remain

unchanged.

5.2.3 Derivation of TRAST contrast

As can be seen in Figure 5.3a, a change in triplet relaxation time affects mostly

images taken with higher duty cycles, due to the inability of molecules to recover

between light pulses. On the other hand, when very short pulses are used, there

is sufficient time for molecules to relax and hence there is no decrease in average

fluorescence for different triplet relaxation times.

Therefore, it is expected that the greatest differences in fluorescence inten-

sity will be found between images taken with the shortest and longest pulse

73
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

widths. The ratio of fluorescence images acquired with different laser beam

dwell times on a scanning confocal microscope has been successfully used be-

fore to distinguish samples of different triplet populations [49]. Here, in order to

simplify and speed up the collection process, we propose and examine an acqui-

sition protocol that is based on the maximum relative decrease in fluorescence

intensity, that is the relative decrease between images taken with the shortest

(t p =1 µs) and longest (t p =50 µs) pulses, which correspond to duty cycles of 1%

and 100%, respectively. We can therefore define the TRAST contrast as:

I shor t − ICW
CT = (5.9)
I shor t

where I shor t and ICW are the intensities recorded from a train of pulses with the

aforementioned characteristics. The values of ICW can be calculated analyti-

cally by setting w = T in equation (5.3), which is reduced to:

A
I em |w=T ≡ ICW = H (5.10)
B +C τT

where H is a constant factor depending on the concentration of the fluorophores

and acquisition efficiency, and,

A = τi sc

B = τi sc (τ01 τi sc + τ10 τi sc + τ01 τ10 + τ210 ) (5.11)

C = τ210 + τ10 τi sc .

On the other hand, I shor t needs to be numerically solved.

Figure 5.4 shows the solution of equation (5.9) for a range of different triplet

relaxation times. It can be seen that TRAST contrast increases from 64.7% to

94.7% for an increase in triplet relaxation time from 2 µs to 20 µs.

74
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

100
Numerical solution
95
90
85

CT (%)
80
75
70
65
60
2 4 6 8 10 12 14 16 18 20
τT (μs)

Figure 5.4: TRAST contrast, defined as C T = (I shor t − ICW )/I shor t , where I shor t is the
sample fluorescence when illuminated using a low duty cycle (1%), and ICW
when using continuous illumination (total illumination time constant), for
a range of triplet relaxation times.

5.2.4 TRAST contrast as a measure of [O2 ]

The effect of molecular oxygen on the triplet dynamics has been a matter of

interest for many years, since the longer lifetimes associated with triplet states

lead to higher quenching probabilities and therefore offer a good candidate for

studies of the molecular microenvironment. Using FCS, a five–fold decrease in

triplet relaxation lifetime when measuring Rh6G in an oxygen atmosphere ver-

sus a de–aerated one has been observed [210]. Similar behaviour for Eosin Y has

been reported [215] and it has been shown that the significant decrease in triplet

state relaxation rate is mainly due to quenching by oxygen [212].

Here, we are examining the relationship between molecular oxygen concen-

trations and TRAST contrast, in order to be able to quantify local oxygen con-

centrations by applying equation (5.9) that only requires the acquisition of two

fluorescence images. It is possible to re–write equation 2.18 as:

τT 0
τT = (5.12)
1 + k q τT 0 [O 2 ]

75
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

95 95
TT0 kq
12 μs 0.6 10-9 M-1 s-1
90 16 μs 90 0.8 10-9 M-1 s-1
20 μs 10-9 M-1 s-1

CT (%)
CT (%)

30 μs
85 85

80 80
(a) (b)
75 75
0 50 100 150 200 250 0 50 100 150 200 250
O2 (μM) O2 (μM)

Figure 5.5: TRAST contrast plotted against oxygen concentrations and the effect of (a)
triplet relaxation time in absence of quencher and (b) bimolecular quench-
ing constant.

where τT 0 is the triplet lifetime in the absence of the quencher.

By combining equations (5.10) and (5.12), we can find the relationship be-

tween TRAST contrast and absolute values of oxygen concentration. When char-

acteristic values for Eosin Y are considered (τ01 = 2.2 ns, τi sc = 2.5 ns, τ10 = 3

ns [213], k q = 109 M−1 s−1 [220] and τT 0 = 16.5 µs1 ), a linear fit (R2 = 0.99) can

suffiently be used to describe the relationship for values of oxygen concentra-

tion ranging from deaerated to air–saturated solutions (0 – 250 µM [221]). The

resulting curve depends on the bimolecular quenching constant, k q , and the

triplet relaxation time in absence of oxygen, τT 0 , both of which were here treated

as constants [221]. More specifically, k q , changes the slope of the curve while

τT 0 sets the y–intercept, as can be seen in Figure 5.5.

5.2.5 Experimental setup

All experiments were performed using an in–house-built SPIM setup. A 488 nm

laser (OBIS LX, Coherent Inc., Santa Clara, CA, USA) is used and its beam is ex-

panded about 5 times by two sets of achromatic lenses. It then passes through

an achromatic cylindrical lens ( f = 75 mm, Edmund optics, Singapore). The


1
Value from own measurements in deaerated bulk solution

76
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

resulting beam over–illuminates the back focal aperture of the low NA illumina-

tion objective (SLMPlan 20X / 0.25, Olympus, Singapore) to obtain a light sheet

≈ 1.2 µm thick. The illumination objective has a working distance of WD = 21

mm. This provides the necessary space to bring the light sheet into the focal

plane of the detection objective (LUMPLFLN 60x/1.0 W, WD = 2.0 mm, Olympus,

Singapore). The sample mounting unit consists of a custom built sample cham-

ber (Physics mechanical workshop, NUS, Singapore) and motorized linear x–,

y– and z–stages together with a rotation stage (XYZ–linear stages: 3×8MT184–

13DC and rotation stage: 8MR174–1–20, Standa Ltd., Lithuania). The detec-

tion objective was mounted on a piezo flexure objective scanner (P–721 PIFOC,

Physik Instruments, Singapore) to control the objective in nm precision. A flu-

orescence emission filter (BrightLine HC525/50, Semrock Inc., New York) was

mounted behind the detection objective, before a standard tube lens (part no.

LU074700, f = 180 mm, Olympus, Singapore) used to image the sample onto the

camera. An Andor iXon X3 860 EMCCD (Andor Technology, Belfast, UK) cam-

era was used in the experiments. The camera pixel size at the object plane is

400 nm, and n × n pixel binning is possible during or post acquisition.

Modulation of the laser output is digitally controlled using a custom–build

program on LabView (National Instruments, Austin, TX) that can create finite–

length pulse trains of desired characteristics. The pulse sequence is fed from the

computer to both the laser and the camera via a terminal box (NI CB–68LPR, Na-

tional Instruments, Austin, TX) connected to a data acquisition board (NI PCI–

6221, National Instruments, Austin, TX). The camera was triggered by the onset

of the first pulse and the exposure time was equal to the full length of the pulse

train. Pulses of widths as short as 250 ns were validated using a Si switchable

gain detector (PDA36A–EC, Thorlabs) connected to an oscilloscope.

77
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

5.2.6 Sample preparation

Bulk solutions

Eosin Y (Sigma–Aldrich, Singapore) was chosen for our experiments due to its

high triplet–yield that facilitates the build–up of triplet state population at rela-

tively low incident powers [215, 218]. Key characteristics of Eosin Y are shown in

Table 5.1. Stock samples of mM concentration were prepared in ethanol (Sigma

Chemicals, St. Louis, U.S.A.). The samples were further diluted in double-distilled

water (ddH2 O) to final concentration of 2 µM. In the full TRAST monitoring

experiments, bulk solution was placed in heat–sealed sample bags made from

thin transparent foil with a refractive index matching that of water (Katco Ltd.,

United Kingdom). Each of the bags contained approximately 50 µL of the solu-

tion. The samples were air–saturated by bubbling atmosphere air in for approx-

imately 30 min using a common aquarium pump. When necessary, complete

deaeration of the sample was performed by addition of Na2 SO3 at final concen-

tration of 2 M [222].

Property Value
Chemical formula C20 H6 Br4 Na2 O5
Molar mass 647.89052
Absorption maximum (λa ) 524 nm
Fluorescence maximum (λ f ) 543.5 nm
Quantum yield 67%
Molar extinction coefficient at 524 nm 112,000 cm−1 /M
Molar extinction coefficient at 488 nm 30,000 cm−1 /M
CAS Number 17372–87–1

Table 5.1: Characteristic properties of Eosin Y [57, 223].

Partial oxygen removal, which is necessary for the TRAST contrast vs oxy-

gen concentration calibration curve, was achieved by titrating L–Ascorbic Acid,

78
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

catalysed by Ascorbate Oxidase [211, 224], according to:

L-Ascorbic Acid + O2 −→ 2 H2 O + 2 Dehydroascorbate (5.13)

L–Ascorbic Acid was purchased from Sigma–Aldrich and dissolved 65 µM. Ascor-

bate Oxidase from Cucurbita sp. (lyophilised powder purchased from Sigma–

Aldrich) was reconstituted in 20 mM sodium phosphate (NaH2 PO4 ) buffer. The

enzyme has an activity of 1,000 to 3,000 units per mg protein. One unit will ox-

idize 1.0 µM of L–Ascorbic Acid to dehydroascorbate per min at pH 5.6 at 25o C .

The samples were placed in larger bags that can hold approximately 800 µL. Ab-

solute values of oxygen concentration were measured using an pre–calibrated

optical fibre oxygen sensor (FireSting2, Pyro–Science) while at the same time,

fluorescence intensity were measured using different pulse widths using the

SPIM setup. The bags were left open at the top to allow for probe insertion (Fig-

ure B.1, Appendix B).

Three units of Ascorbate Oxidase were added prior to placement of the sam-

ple in the microscope and aliquots of L–Ascorbic Acid were injected and gently

mixed inside the bag by pipetting, before the onset of acquisition. Measure-

ments were taken during the course of atmospheric oxygen uptake, while the

absolute values were monitored by the probe. Measurements were repeated by

replenishing the solution.

Hydrogel beads

Calcium alginate hydrogels were prepared using the methods described in [174].

Alginate acid (Acros Organics, U.K.) solutions (3% wt/vol) in water were pre-

pared and autoclaved. Filter–sterilised CaCl2 (100 mM) solution was poured in

a 6 well plate (Nunclon Flat, Thermoscientific) in columns approximately 2 mm

tall. Droplets of 10 – 20 µL of alginate acid were dropped into the well using

a pipette. The cross linking was almost instantaneous. The drops were left to
79
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

gelate in the CaCl2 for approximately 20 min before they were washed twice

with sterile water. The beads were then kept in sterile water, and imaging ex-

periments were carried out the same day. For imaging, the beads were trans-

ferred to heat–sealed bags made from thin transparent foil with a refractive in-

dex matching that of water (Katco Ltd., United Kingdom) that included the bulk

dye solution.

Biofilms

A flow cell setup system similar to the one described at ref. [177] was used.

A square Fluorinated Ethylene Propylene (FEP) tube (inner diameter: 2.1 mm,

outer diameter: 2.3 mm) was used as the flow cell (Adtech Polymer Engineer-

ing Ltd, UK). P. aeruginosa (PAO1) were grown overnight in LB medium (Difco)

40 g/L in a shaking incubator at 37o C, 200 rpm. The culture was diluted in min-

imal medium ([15 mM (NH4 )2 SO4 , 34 mM Na2 HPO4 , 22 mM KH2 PO4 , 58 mM

NaCl, 1 mM MgCl2 , 0.1 mM CaCl2 and 0.001 mM FeCl3 ], supplemented with 2

g/L of Glucose plus 2 g/L of Casamino acids) to an OD600 ≈ 2.0 and 300 µL were

injected into the tubing, near the entry port of the flow cell. The inoculum re-

mained in the flow cell to settle for 1 h without flow, at room temperature. The

flow of the minimal medium was then resumed at a rate of ≈ 6 ml/h. A 6–Roller

peristaltic pump (Langer Instruments, U.S.A.) was used to supply the flow.

The biofilms were allowed to grow on the surfaces of the flow cell. Prior to

imaging, flow was stopped and one of the four faces of the flow cell was carefully

cut out using a surgical scalpel and placed in a thin foil, heat–sealing bag which

contained approximately 50 µL of the dye (Eosin Y, 2 µM). The substratum ma-

terial (FEP tube) was chosen due to its optical and mechanical properties; that

is, its refractive index is close to that of water, and it is a rigid material to sus-

tain flow during the growing phase, yet light enough to be placed in the FEP bag

80
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

(a) Detection
Illumination objective z
(b) y
objective
x

Figure 5.6: (a) Illustration of mounting concept for bacterial biofilms in SPIM. One of
the surfaces of the square tube which has attached biofilm on, is cut and
removed using a scalpel and put in a bag made of thin FEP matching the
refractive index of water and filled with dye solution. Drawing is not to scale.
(b) Schematic showing the optical sectioning of a colony.

during imaging.

All connecting FEP tubings (30 mm long, 2.1 mm ID and 2.3 mm OD) were

autoclaved and flushed with 10% (v/v) bleach, then thoroughly rinsed in sterile

milli–Q and then primed with the minimal medium before the start of every

experiment.

5.2.7 Acquisition and analysis

In a pulse train, the duty cycle (α) is given by α = w/T , where w is the pulse

duration and T the period (Figure 5.1a). The total illumination time is given by

Ti l l = N ·w, where N is the number of pulses within the train. In the TRAST mon-

itoring experiments of this study, the illumination time was kept constant (Ti l l =

100 ms). This was achieved by changing both the duty cycle and the number of

pulses accordingly, while the period of the pulse train was kept unchanged. The

pulse width duration, w, varied from 1 to 50 µs, the number of pulses, N , from

200,000 to 2,000 while the period, T , was kept constant at 50 µs. The total acqui-

sition time at each point of the curve can therefore be calculated by T ac q = N ·T .

The laser power before objective was 27.3 mW.

81
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

7000

Fluorescence intensity (AU)


6800

6600

6400

6200

6000

5800

5600

Baseline <1 minute 2 minutes 3 minutes 4 minutes


Interval between images

Figure 5.7: Eosin Y (2 µM) in double–distilled water (ddH2 O) samples repeatedly im-
aged after allowing different time intervals between images. Values shown
are the mean and standard deviations of the recorded intensity across 50 ×
50 pixel images. Three independent measurements were taken for each con-
dition. The baseline measurement is shown as a reference of the expected
values in absence of bleaching.

Up to fifteen images were taken per experiment at increasing pulse widths

to compile each TRAST curve (see Table A.1, Appendix A). The exposure time

of the camera was set equal to T ac q for each acquisition, so that intensity was

integrated during the full pulse train. A pause of approximately 2 min was in-

troduced between image acquisitions to ensure that dye in the focal volume has

been replenished, thus avoiding artefacts from bleaching. This was determined

by a calibration experiment where we repeatedly imaged samples at different

time intervals and we found that the intensity is fully recovered when at least

two minutes are allowed between images (Figure 5.7). The acquired images were

loaded and analysed by a custom–made Matlab script [225] that can perform a

pixel–wise fitting on the solution of the three–state model under the condition

of isodose illumination.

To calculate TRAST contrast, images using only the lowest duty cycle (α =

1%) and continuous illumination (α = 100%) were collected (see Appendix B.1).

Ti l l was kept constant, as above. The relative intensity decrease was then calcu-

lated in a pixel–wise manner, according to equation (5.9).

82
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

5.3 Results

5.3.1 Measurements in bulk solutions

To validate the setup, we performed TRAST measurements in homogeneous

bulk solutions of Eosin Y (2 µM) in double–distilled water (ddH2 O), taken in both

air–saturated and deaerated conditions. Complete deaeration was performed

by addition of Na2 SO3 . Figure 5.8a shows typical TRAST curves for measure-

ments in the aforementioned conditions. The error bars on the plot correspond

to the standard deviation of the intensity of each image, 20 × 20 pixels in size.

We found a significant increase in triplet relaxation time from 2.16 ± 0.02 µs to

16.52 ± 0.06 µs when oxygen was removed (Figure 5.8b,c). Our results are com-

pared to previously published results in Table 5.2. For the fitting, the initial value

of τ01 was calculated from (5.8), and the initial values for τi sc , τ10 , and τT were

given based on literature review.

Sample Method [Ref.] τT (µs) τi sc (ns)


Eosin Y, Present study 2.16 ± 0.02 1.86 ± 0.02
air–saturated solution SPIM–TRAST [213] 2.08 ± 0.04 1.34 ± 0.06
Confocal–TRAST [213] 2.38 ± 0.11 1.32 ± 0.14
Picosecond double– – 2.38 ± 0.28
pulse excitation [218]
Flash photolysis [216] – 1.19 ± 0.11
TRAST [215] 1.89 ± 0.04 1.18 ± 0.03
Eosin Y, Present study 16.52 ± 0.06 3.03 ± 0.01
deaerated solution TRAST [215] 29.41 ± 3.40 2.96 ± 0.25

Table 5.2: Triplet relaxation and intersystem crossing times for Eosin Y in air–
saturated and deaerated solutions, together with values from litera-
ture review.

Next, we measured the TRAST contrast in solutions of varying oxygen con-

tent, prepared by titration of Ascorbic Acid, in order to acquire a calibration

83
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

1 20
Aerated (b)
0.9 Aerated
Deaerated 15
0.8

0.7 10
Intensity (AU)

0.6 5
0.5
0
0.4 2.05 2.15 2.25
0.3 25
0.2 Deaerated (c)
20
(a)
0.1
5 10 15 20 25 30 35 40 45 50 15
0.02 10
Residual

0.01
0 5
-0.01
0
-0.02
5 10 15 20 25 30 35 40 45 50 16.3 16.5 16.7
Pulse width (μs) Triplet time (μs)

Figure 5.8: (a) TRAST curves from measurements in homogeneous, bulk solutions of
Eosin Y (2 µM) in double–distilled water, in air–saturated and de–aerated
conditions. (b, c) Extracted triplet times from pixel–wise fitting for the air–
saturated and deaerated solution, respectively. After [196]. © IOP Publish-
ing. Reproduced with permission. All rights reserved.

90

80
Relative decrease in intensity (%)

70

60

50

40

30

Oxygen concentration [μM]

Figure 5.9: TRAST contrast in samples of varying oxygen content. Two independent
measurements of Eosin Y (2 µM) in double–distilled water (open and solid
diamonds) and linear fits. After [196]. © IOP Publishing. Reproduced with
permission. All rights reserved.

84
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

curve between absolute values of oxygen concentration and TRAST contrast

readings, as shown in Figure 5.9. The error bars on the x–axis were derived from

the fluctuation of the oxygen content as measured by the probe over the time

of the measurement (including the pause between acquiring the low duty cycle

image and the continuous illumination image). The error bars on the y–axis de-

note the standard deviation between pixels (128 × 128 pixel images). Two inde-

pendent measurements on freshly prepared samples are shown (open and solid

diamonds) together with their linear fits (dashed lines). The slope of the linear

fit was (−0.96 ± 0.07) 10−3 . We investigated the effect of viscosity by measuring

TRAST contrast in solutions of increasing sucrose concentrations [226]. No sig-

nificant change was found for samples of viscosities of 1.30 and 1.80 cP, where

the slope was (−1.06±0.15) 10−3 and (−0.77±0.45) 10−3 , respectively, calculated

from two independent measurements in fresh samples for each case.

5.3.2 Measurements in heterogeneous samples

Since differences in the material properties of the sample other than its oxy-

gen content might affect the fluorescence intensity, next, we performed control

measurements in non–homogenous samples that, however, do not produce or

consume oxygen locally. Thus, calcium alginate hydrogel beads, which are com-

monly used as model synthetic matrices in biofilm studies as their composition

resembles that of biofilms [174, 180, 181], were prepared and studied.

We selected a ROI that includes the interface of the bead with the aqueous

solution (Figure 5.10a). A full TRAST imaging experiment showed that the triplet

relaxation lifetimes inside the hydrogel bead were not different from those in the

surrounding solution (Figure 5.10b), leading to a single–population distribution

(Figure 5.10c). The extracted values of triplet relaxation time and intersystem

crossing (τT = 2.09 ± 0.08 µs and τi sc = 1.92 ± 0.11 ns, respectively) resemble

those found in fully aerated bulk solutions. This is an expected result, since there
85
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

2.4 100
(a) (b) (c) Triplet time (μs)
2.3
80
2.2
60
2.1

2.0 40

1.9 20

1.8
0
μs 1.8 2.0 2.2 2.4

65 2.4 1
(d) (e) (f) 0.99
2.3
60
2.2 0.98

2.1 0.97
55

2 0.96

50 0.95
1.9

1.8 0.94
45
% ns 0.93

Figure 5.10: (a) Transmission image of a calcium alginate (3% w/w) hydrogel bead in
Eosin Y (2 µM) solution. (b) Triplet relaxation time map and (c) frequency
histogram. (d) TRAST contrast image. (e) Frequecy histograms of the ex-
tracted intersystem crossing times. (f) R2 of the fit. The pixel size on panels
(b-f) is 1.6 × 1.6 µm (after 4 × 4 pixel binning). White pixels represent areas
where the fitting has failed. Scale bar: 8 µm. After [196]. © IOP Publishing.
Reproduced with permission. All rights reserved.

was no production or consumption of oxygen in the sample, and oxygen diffuses

freely within the hydrogel. A TRAST contrast image is shown in Figure 5.10d. The

average contrast inside and outside the bead is 57±2% and 60±2%, respectively.

A map of the intersystem crossing times is shown in Figure 5.10e and lastly, R 2 ,

as a measure for goodness of fit, is shown in Figure 5.10f.

5.3.3 Measurements in bacterial biofilms

We used TRAST imaging to extract the transient kinetics, and TRAST contrast

to quantify the oxygen distribution inside a P. aeruginosa biofilm that had been

growing under continuous flow of minimal medium for ≈ 80 h after cell inocu-

lation. Prior to imaging, the flow was stopped and the samples were transferred

to the FEP bag for imaging. Images were taken ≈ 3 h after flow was stopped.

The height of the microcolonies was in the range of 30 – 60 µm and all images

86
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

were taken at mid–point depth for each microcolony (geometry shown in Figure

5.6b).

Figure 5.11(I) a single biofilm colony imaged in an air–saturated solution.

A wide range of triplet relaxation times were found, evident from the pixel–to–

pixel variations, as extracted from local fitting with equation (5.3). However, the

average triplet relaxation time inside the colony was significantly longer com-

pared to that outside (5.53 ± 1.17 µs and 3.41 ± 0.23 µs, respectively).

After complete oxygen removal by the addition of Na2 SO3 , there was no sta-

tistical difference in the triplet relaxation times inside and outside the colony

(9.09 ± 0.27 µs and 9.15 ± 0.17 µs, respectively), as shown in Figure 5.11(II). In-

stead, a narrow distribution of triplet times was observed (Figure 5.11c) whose

values were significantly longer than those measured prior to deaeration. Sim-

ilarly, the values of intersystem crossing times where limited to 4.75 – 5.68 ns

after deaeration, versus 2.10 – 7.45 ns before.

Using equation (5.9), TRAST contrast was calculated for this colony, and is

shown in figure 5.12a. It was found that inside the biofilm the average value was

significantly larger compared to outside (68 ± 7% and 56 ± 3%, respectively). On

the other hand, TRAST contrast was almost uniform (77±1% and 79±1%, inside

and outside of a colony, respectively) when measured after complete oxygen re-

moval (results not shown).

Lastly, using the calibration curve derived earlier (Figure 5.9), and after set-

ting the value of τT 0 as extracted from the deaerated biofilm measurement, a

map of absolute oxygen concentrations was constructed and shown in Figure

5.12b. It can be seen that oxygen gradients have been established, and oxygen

depletion extended outside the area where bacteria lie (dashed lines). The nec-

essary assumptions for converting TRAST contrast to absolute values of oxygen

87
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

I
11

5
y
3
x (a) (b) (c) μs
70 1
(d) 7
60 Triplet time (μs) 0.99
6.5
6 0.98
50
5.5 0.97
40 5
0.96
4.5
30 4 0.95

20 3.5 0.94
3
0.93
10 2.5
(e) 2 (f) 0.92
0 ns
2 4 6 8 10

II
10. 5

10

9.5

8.5
(a) (b) (c) μs
100 5.7 0.997
(d) Triplet time (μs) 5.6
0.996
80 5.5
5.4 0.995
60 5.3
0.994
5.2
5.1 0.993
40
5
0.992
4.9
20 0.991
4.8
(e) (f)
0 ns
8 8.5 9 9.5 10 10.5

Figure 5.11: A P. aeruginosa biofilm colony imaged in solution of Eosin Y (2 µM) in mini-
mal medium. Panel I: Air–saturated conditions. Panel II: Deaerated condi-
tions. (a) SPIM images. (b) Fluorescence images taken with 100 ms contin-
uous illumination. (c) Triplet relaxation time maps and (d) frequency his-
tograms, showing the pixel–to–pixel variation, extracted from full TRAST
experiments. (e) Intersystem crossing time maps. (f) R2 of fit. Images taken
≈ 80 h post–inoculation. The pixel size was 0.8×0.8 µm (after 2×2 pixel bin-
ning) and white pixels represent areas where the fitting has failed. Dashed
lines mark the boundaries of the colonies to aid visualisation. Scale bar:
4 µm. After [196]. © IOP Publishing. Reproduced with permission. All
rights reserved.

88
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

90 300

2 00
50

15
80 250

0
10
10
25 200

0
70 0

20 150
60 50 0

15
10 100

0
50 5 0 10

100
10

10
0 50
50
40
(a) (b)
10

50
10
0 0

10

0
% μM

0
Figure 5.12: (a) TRAST contrast image and (b) contours of oxygen concentration in-
side the colony of Figure 5.11(I). Dashed lines mark the boundaries of the
colony to aid visualisation. Scale bar: 4 µm. After [196]. © IOP Publishing.
Reproduced with permission. All rights reserved.

concentrations in living systems are discussed in the next section.

5.4 Discussions

Reducing the irradiance requirements The first published protocols for TRAST

imaging required long, up to 3 min, acquisition times and high laser excitation

powers, up to 500 kW/cm2 [49], which can lead to photodamage of the sam-

ple. To date, a number of steps have been taken to decrease these requirements

so that TRAST imaging can find more applications in biomedical studies. In a

study published in 2010, the exposure time was reduced to 1.2 ms per image,

however the irradiance remained rather high, at 60 kW/cm2 [211]. More recent

studies [213, 215] utilised Eosin Y that has a higher triplet yield, and managed

to decrease the laser irradiance requirement to 0.44 kW/cm2 and ≈ 10 kW/cm2 ,

however, their protocols still required over 20 and up to 57 images per TRAST

curve, respectively.

In this thesis, to further reduce illumination time, we have proposed and val-

idated a different acquisition protocol and explored its applicability for quan-

tifying oxygen concentrations. Specifically, we calculate and use TRAST con-

89
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

trast, that is the relative decrease in fluorescence intensity between two images:

one taken with a low duty cycle (α = 1%) and one with continuous illumination

(α = 100%), as per equation (5.9). This allows the total illumination to be low-

ered by a factor of up to 25 in comparison to previous studies. Measurements

presented in our study were performed at an irradiance of ≈ 7 kW/cm2 and the

total illumination time for each image was 200 ms.

It can be argued that the robustness of the fit could improved by increasing

the number of acquired images. However, acquiring additional images leads

to larger light doses and longer acquisition times, which can be deleterious to

samples. This trade–off has indeed been discussed in a previous study, where

authors note that they had to compromise on image quality by reducing laser

power in order be able to acquire additional images [211]. It should be also

noted that by using TRAST contrast as a measure, one does not have to use equa-

tion (5.3) for fitting; instead, the simpler equation (5.9) can be used, leading to

reduction of post–processing computational time as well.

Quantifying local oxygen concentration with TRAST contrast Our simula-

tions indicate that, when transition rates characteristic of Eosin Y are consid-

ered, a linear relationship holds between TRAST contrast and absolute oxygen

concentration (Figure 5.5).

We first tested and validated this relationship in calibration experiments with

bulk dyes (Figure 5.8). Subsequently, we examined its applicability in heteroge-

neous samples by imaging alginate hydrogel beads whose composition is simi-

lar to that of biofilms [174, 180, 181]. We found that local differences in the ma-

terial properties of the sample did not affect the measurements (Figure 5.10), an

expected result since TRAST contrast is a concentration–independent method

[49].

90
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

For the correct application of the calibration curve on complex samples,

both its slope and intercept must be known. To estimate oxygen concentrations

in biofilms we have here used the slope extracted from bulk dye measurements

and the intercept from deaerated biofilm measurements.

Oxygen concentration distribution inside biofilm colonies Single colonies

of P. aeruginosa biofilms were studied. First, full TRAST experiments showed

that the triplet relaxation times inside the biofilm are significantly longer than

those measured in the surrounding medium. Second, we used TRAST contrast

to quantify the oxygen concentration. It should be noted here, that in complex

living systems, a number of environmental factors and dynamic processes can

potentially affect the triplet–singlet formations, therefore, one needs to be care-

ful with the interpretation of results. In order to assess any possible cross–talk

artefacts from other processes, we deliberately depleted oxygen in the sample

and measured the triplet relaxation times. We no longer found any gradients

or local differences (Figure 5.11(II)c–d), and the biofilm became indiscernible

in the triplet relaxation time map, implying that, in our case, oxygen concentra-

tion is the major factor affecting triplet relaxation time and TRAST contrast. This

is a similar assumption to those made in previous published studies [215, 211].

The absolute values of τT 0 in the deaerated biofilms were different than those

found in deaerated bulk dye solutions. This discrepancy could be attributed to

by–products or other mechanisms associated with biofilm processes.

Figure 5.12 shows that in the biofilm imaged here, anoxic areas existed within

the colony and that oxygen depletion extended beyond the area of bacterial

cells. This finding is in agreement with a previous study that measured oxygen

concentration at the base of biofilms. The authors found that P. putida colonies

have anoxic centers and a “relatively homogeneous oxygen distribution of 5%

to 7% air saturation” when they measured below colonies, under low flow to

91
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

stagnant conditions [128]. An older study using microprobes has also reported

gradients of oxygen depletion towards the deeper parts of biofilms [117]. These

anoxic regions could be attributed to the respiration of oxygen–consuming bac-

teria that results in rapid reduction in the oxygen concentration from the surface

to the interior of the micro colonies. Further, it has previously been shown that

genes associated with anaerobic respiration were induced in the interior of mi-

cro colonies [82].

Interestingly, we also found areas inside the biofilm with higher oxygen con-

centrations compared to those close to the boundary. This could be attributed

to transport channels that are known to exist in colonies [44, 228], as well as

to the porosity of colonies [229]. By convection through these channels, in-

ner parts of biofilms can receive nutrients and metabolites. Furthermore, it is

known that some colonies have hollow parts in their interior, due to dispersal

of the bacteria [85]. In this way, the bacteria on the outer surface of the micro-

colony would consume and reduce the local oxygen content, while oxygen could

leak into the interior of the colonies through the channels described above.

Overall, the cross sections presented here show a highly heterogeneous oxy-

gen distribution with steep gradients both inside and outside the visible bound-

aries of the colony (Figure 5.12). Oxygen concentration reaches air–saturation

levels at approximately 15 – 20 µm away from the visible colony edge. It should

be noted that unlike any of the existing studies, our measurements were done at

approximately half the depth of a 60 µm thick mushroom–like colony.

Avoiding photo–bleaching artefacts TRAST imaging is based on relative dif-

ferences in intensities. For the correct interpretation of TRAST measurements,

it is crucial that there are no photobleaching artefacts that further decrease flu-

orescence. If no precausions or corrections against photobleaching are taken,

92
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

the perceived fluorescence decrease would be overestimated, leading to an un-

derestimation of oxygen concentration. In previous studies which investigated

phenomena happening in fast (tens of seconds) timescales, bleach correction

was retrospectively performed. Such an approach can decrease the total exper-

imental time to a few seconds, however, it introduces the need of “extensive im-

age processing” [211]. In our measurements we allow for dye replenishment in

the region of interest instead (Figure 5.7). In all results presented here, a pause

of 2 min was introduced between images, which is the limiting factor for our to-

tal experimental time. Despite being slower than protocols that include correc-

tions for image bleaching, ours is suitable to study phenomena that happen on

longer time scales, such as the formation of anaerobic areas in growing biofilms.

A post–acquisition bleach correction could readily be employed, if necessary.

Therefore, the theoretical temporal resolution of the TRAST contrast method is

limited only by the acquisition time (T ac q ) needed for two images: one with a

pulse train of low duty cycle (in our experiments 200,000 pulses of 1% duty cy-

cle with 50 µs period, i.e. Ti l l = 100 ms and T ac q = 10 s) and one with constant

illumination with the equivalent dose (Ti l l = T ac q = 100 ms).

Limitations The main limitation in the presented methodology is that the mea-

surements on biofilms were done in a closed system, that is the thin foil sealed

bag. In such arrangements it is expected that nutrients will be depleted over

time, influencing the activity and oxygen consumption rate. Additionally, since

there is no flow of media during imaging, the system is dominated by diffusion.

This could contribute to our observation of gradients extending into the bulk

liquid. It is possible that the oxygen profile of biofilms imaged under flow would

be quite different, and this is something worth investigating in future studies.

Lastly, in order to convert the TRAST contrast to absolute values of oxygen

concentration, the slope of the calibration curve needs to be known in a pixel–

93
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

wise manner, while in our analysis we have used a constant value that was calcu-

lated from calibration measurements in bulk aqueous solutions. Nonetheless, it

has been reported that the local viscosity in L. lactis biofilms is “not modified in

comparison with that of water” [227]. Another study has measured the viscosity

of the EPS matrix of S. maltophilia to be 1.50 cP, using FCS [230]. We found no

significant change in the slope of the linear fit when we measured bulk solutions

up to 1.80 cP, prepared by addition of sucrose. No data about local dynamic vis-

cosities in P. aeruginosa biofilms exist to our knowledge, however, based on the

above, it is reasonable to assume a viscosity of our biofilm samples close to that

of water.

5.5 Conclusion

In this chapter we have implemented TRAST imaging on a SPIM system, driven

by its capability to provide optical slicing for thick samples with minimal pho-

totoxicity. For the first time, we have shown high–resolution maps of triplet re-

laxation times inside bacterial biofilms. Furthermore, we have defined TRAST

contrast and presented a new acquisition protocol that allows quantification

of molecular oxygen concentrations based on a calibration curve that was vali-

dated for different micro–environmental conditions. In contrast to most other

existing techniques for measuring oxygen, the technique presented here is not

limited to measuring close to the surface or the substrate of biofilms. Using

TRAST contrast, we were able to image the oxygen concentration profile inside

a P. aeruginosa biofilm, in a non–destructive way. Steep concentration gradients

and zones of complete deaeration inside the colony were found, that would im-

possible to identify at such high resolution with other methods.

We believe that the approach taken here can be used to further our under-

standing of the role of oxygen during biofilm development and life cycle, and

94
CHAPTER 5. SPIM – TRAST FOR O2 MEASUREMENTS

that this method has potential to be used as a diagnostic tool in the future, such

as for example to identify aerobic or anaerobic zones in multispecies biofilms

[231] or to study oxygen consumption rates of genetically modified strains of

bacteria.

In future studies, TRAST imaging could work in tandem with other tech-

niques that can provide information about local diffusion and dynamics – such

as for example FCS or anisotropy measurements that can also be established on

camera systems [201, 232]. This could allow accounting for possible artefacts

due to local heterogeneities within the imaged region. Similarly, TRAST imaging

could be combined with sensors that can monitor other factors that could affect

the results, such as for example pH, or other quenchers like thiol derivatives.

95
Chapter 6

Afterword

6.1 Overall conclusions

This thesis aimed to introduce novel spectroscopy techniques in the study of

bacterial biofilms. It is now known that active biofilms are dynamic systems

whose development is characterised and driven by physical and chemical gra-

dients. Hence, methods that can provide spatially and temporally resolved in-

formation are of great value, even more so when they can be applied in vivo.

Two questions in particular were raised and addressed: “How is the stiff-

ness of colonies organised internally?”, and, “How is oxygen distributed inside

biofilms?”. Focus was put on studying these two parameters as they are both

associated with biofilm dispersal which is of high interest in strategising meth-

ods for inhibiting and removing biofilms. To answer these questions, two novel

spectroscopic techniques, namely Brillouin Imaging and Transient State (TRAST)

imaging were employed and further developed.

Regarding stiffness, no correlation with colony size was found when com-

paring the averages of Brillouin shifts of two–dimensional cross–sections of 24

randomly selected colonies growing under continuous flow of minimal media.

Given the high degree of complexity of the samples and the large number of vari-

96
CHAPTER 6. AFTERWORD

ables, a multivariate analysis could be more helpful to reveal quantitative data.

However, the sample size and the spatial–averaging of the Brillouin shift pro-

hibit such an approach. To this end, we focused our study on single colonies,

where we observed two distinct spatial patterns: in smaller colonies, stiffness

increased towards their interior, indicating a more compact structure of the cen-

tre of the colony, whereas, larger (over 45 µm) colonies were found to have less

stiff interiors. We have also identified differences in the stiffness profile of a sin-

gle colony when imaged at different time points. Regarding oxygen distribu-

tion, we have found that oxygen in the deeper parts of the colony is depleted,

which was also known from older studies using microsensors. However, only

with the combination of TRAST with SPIM and after a calibration procedure, as

was done here, was it possible to quantify the steep oxygen gradients that exist

within the colony and extend outside its boundaries. The above highlights that

spectroscopy combined with optical microscopy can provide valuable informa-

tion to microbiologists that would not be available otherwise.

We believe that our results can be beneficial to the study of the complex dy-

namics that drive biofilms and their dispersal. For example, it has been found

that c–di–GMP, a dynamic intracellular signaling molecule, is associated with

the onset of dispersal. More specifically the concentration of c–di–GMP is de-

creased when P. aeruginosa biofilms are exposed to nitric oxide (NO)–releasing

compounds that cause dispersal [233]. Further, a very recent study, using a ra-

tiometric, image–based quantification method, showed that c–di–GMP distri-

bution within colonies is heterogeneous. In particular, high amounts of c–di–

GMP were found to localise towards the outer boundaries of mature colonies,

whereas in smaller colonies a more uniform distribution was found to exist [234].

The localisation of the decrease of c–di–GMP – a cue for dispersal onset – in the

centres of mature colonies, coincides with both our findings from Brillouin mi-

97
CHAPTER 6. AFTERWORD

croscopy that indicate softer cores in larger colonies, as well as with our TRAST

imaging results that found anoxic zones within colonies.

6.2 Limitations and outlook

For the techniques presented here to find wider applications in microbiology,

progress has to be made in two domains: First, a rigorous theoretical under-

standing of the phenomena underlying the measured quantities has to be de-

veloped. As discussed, in the case of Brillouin microscopy the exact connection

between the measured Brillouin shift and the mechanical modulli in heteroge-

neous samples is, to date, not known. Likewise, in TRAST imaging, the contri-

bution of other factors in the relative changes of triplet relaxation times needs to

be quantified and accounted for. Second, further engineering refinement of the

techniques is needed to allow for more robust and repeatable measurements to

be made. In Brillouin imaging, the current bottleneck is temporal resolution due

to the weak Brillouin signal. This can be overcome in the future by employment

of additional background suppression techniques, such as molecular absorp-

tion cells [157, 158] or notch filters [161]. On the other hand, in SPIM – TRAST

imaging the current bottleneck was the development of a flow cell, suitable for

SPIM, that allows for measurements under constant flow.

Despite these limitations, we envision an exciting future for these techniques

in microbiology, as they can readily be applied to any biofilm, either grown in

nature or developed in the laboratory. A possible combination of both presented

methods under the same platform could give answers relating to the degree of

spatiotemporal correlation between the onset of dispersal and the depletion of

oxygen or other metabolites, especially if markers that can identify transitions

between different stages of the biofilm lifecycle are used [82]. Further, Brillouin

microscopy could used to evaluate different levels of rigidity deriving from spe-

98
CHAPTER 6. AFTERWORD

cific genes and their mutants [235], or to assess how stiffness is affected at the

microscopical level by different production rates of extracellular polysaccha-

rides, such as Psl or alginate [236]. Other studies could investigate the differ-

ences in mechanical properties induced by biofilms growing on different sub-

strates [237], or, to study the internal changes in biofilm stiffness resulting from

exposure to different dispersal agents, which, to date, have only been done in

bulk [101]. Similarly, TRAST imaging can be used to study oxygen distribution

and consumption inside multispecies biofilms [231] or to monitor changes in

the physiology of bacteria occurring as a response to oxygen depletion [44].

Therefore, the methods that this thesis has dealt with bring unique value

with their high spatial resolution and their non–destructive nature, and, are

compatible with existing microbiological protocols. Hence, their adaptation in

defined future studies can help to shed some light on emerging questions re-

garding the biofilm mode of life.

99
Bibliography

[1] R. Hooke “Micrographia: or Some Physiological Descriptions of Minute


Bodies Made by Magnifying Glasses. With Observations and Inquiries
Thereupon,” London: J. Martyn and J. Allestry, First Edition (1665).

[2] C. Karlsson – Rosenthal “The Beginning” Nature Cell Biol. 11(S6) (2009).

[3] N. Lane “The unseen world: reflections on Leeuwenhoek (1677) Concern-


ing little animals” Phil. Trans. R. Soc. B 370, 20140344 (2015).

[4] A. Leewenhoeck “Observation, communicated to the publisher by Mr.


Antony van Leewenhoeck, in a Dutch letter of the 9 Octob. 1676 here En-
glish’d: concerning little animals by him observed in rain-well-sea and
snow water; as also in water wherein pepper had lain infused” Phil. Trans
(1677).

[5] G. B. Airy “On the diffraction of an object–glass with circular aperture,”


Trans. Cambridge Phil. Soc. 5, 283–91 (1835).

[6] E. Abbe “Beiträge zur Theorie des Mikroskops und der mikroskopischen
Wahrnehmung,” Archiv für Mikroskopische Anatomie 370, 413–8 (1873).

[7] J. F. W. Herschel “On a case of superficial colour presented by a homoge-


neous liquid internally colourless,” Philos Trans R Soc London 135:143–5
(1854).

[8] G. G. Stokes “On the change of refractibility of light,” Philos Trans R Soc
London 142, 463–562 (1852).

[9] A. de Meijere “Adolf von Baeyer: Winner of the Nobel Prize for Chemistry
1905” Angewandte Chemie International Edition Volume 44(48), 7836–40
(2005).

[10] M. Renz “Fluorescence microscopy – A historical and technical perspec-


tive,” Cytometry A 83(9), 767–79 (2013).

[11] F. Zernike “Phase contrast, a new method for the microscopic observation
of transparent objects,” Physica 9, 974–86 (1942).

[12] W. J. Schmidt, “Doppelbrechung der Kernspindel und Zugfasertheorie der


Chromosomenbewegung,” Chromosoma 1, 253–64 (1939).

101
BIBLIOGRAPHY

[13] S. Inoué “Polarization optical studies of the mitotic spindle. I. The demon-
stration of spindle fibers in living cells. Chromosoma 5, 487–500 (1953).

[14] F. H. Smith “Microscopic interferometry,” Research (London) 8, 385–95


(1955).

[15] G. Nomarski “Microinterférométre différentiel à ondes polarisées,” J. Phys.


Radium 16, 9S–11S (1955).

[16] M. Minsky “Microscopy Apparatus” U.S. Patent #3013467 (1957).

[17] M. Minsky “Memoir on inventing the confocal scanning microscope,”


Scanning 10, 128–38 (1988).

[18] C. J. R. Sheppard, and A. Choudhury “Image Formation in the Scanning Mi-


croscope,” Optica Acta: International Journal of Optics 24, 1051–73 (1977).

[19] J. G. White, W. B. Amos, and M. Fordham, M “An evaluation of confocal


versus conventional imaging of biological structures by fluorescence light
microscopy,” J Cell Biol 105, 41–8 (1987).

[20] D. Axelrod “Cell-substrate contacts illuminated by total internal reflection


fluorescence,” J. Cell Biol. 89, 141–5 (1981).

[21] H. Siedentopf and R. Zsigmondy “Uber Sichtbarmachung und Größenbes-


timmung ultramikoskopischer Teilchen, mit besonderer Anwendung auf
Goldrubingläser,” Annalen der Physik 315(1), 1–39 (1902).

[22] A. H. Voie, D. H. Burns, and F. A. Spelman “Orthogonal-plane fluores- cence


optical sectioning: Three–dimensional imaging of macroscopic bio–logical
specimens,” J. Microsc. 170, 229–36 (1993).

[23] J. Huisken, J. Swoger, F. Del Bene, J. Wittbrodt, E. H. K. Stelzer, and F. D.


Bene “Optical sectioning deep inside live embryos by selective plane illu-
mination microscopy,” Science 305, 1007–9 (2004).

[24] M. Jemielita, M. J. Taormina, A. Delaurier, C. B. Kimmel, and R.


Parthasarathy, “Comparing phototoxicity during the development of a ze-
brafish craniofacial bone using confocal and light sheet fluorescence mi-
croscopy techniques,” J. Biophotonics 9, 1–9 (2012).

[25] P. Liu, S. Ahmed, and T. Wohland “The F–techniques: advances in receptor


protein studies,” Trends Endocrinol Metab. 19(5), 181–90 (2008).

[26] D. Magde, E. Elson, and W. W. Webb “Thermodynamic fluctuations in


a reacting system âĂŤ measurement by fluorescence correlation spec-
troscopy,” Phys. Rev. Lett. 29, 705–8 (1972).

[27] M. Ehrenberg, R. Rigler “Rotational brownian motion and fluorescence in-


tensify fluctuations,” Chemical Physics, 4(3), 390–401 (1974).

102
BIBLIOGRAPHY

[28] R. Rigler, J. Widengren “ Ultrasensitive detection of single molecules by flu-


orescence correlation spectroscopy,” Bioscience 3, 180–3 (1990).

[29] J. Ricka, T. Binkert “Direct measurement of a distinct correlation function


using fluorescence cross correlation,” Phys. Rev. A. 39, 2646–52 (1989).

[30] P. Schwille, F. J. Myer–Almes, and R. Rigler “Dual–color fluorescence cross-


correlation spectroscopy for multicomponent diffusional analysis in solu-
tion,” Biophys. J. 72, 1878–86 (1997).

[31] R. Peters, J. Peters, K. H. Tews, and W. Bähr “A microfluorimetric study of


translational diffusion in erythrocyte membranes,” Biochim. Biophys. Acta
367, 282–94 (1974).

[32] L. Stryer, and R. P. Haugland “Energy transfer: a spectroscopic ruler,” Proc.


Natl Acad. Sci. USA 58, 719–26 (1967).

[33] J. R. Lakowicz, H. Szmacinski, K. Nowaczyk, W. J. Lederer, M. S. Kirby, M.


L. Johnson “Fluorescence lifetime imaging of intracellular calcium in Cos
cells using Quin–2,” Cell Calcium 15, 7–27 (1994).

[34] W. Schrof, J. Klingler, S. Rozouvan, and D. Horn, “Raman correlation spec-


troscopy: A method for studying chemical composition and dynamics of
disperse systems,” Phys. Rev. E, 3:R2523–R2526 (1998).

[35] R. Begley, A. B. Harvey, and R. Byer. “Coherent anti–Stokes Raman spec-


troscopy,” Applied physics letters, 25(7), 387–90 (1974).

[36] R. Pecora, “Doppler Shifts in Light Scattering from Pure Liquids and Poly-
mer Solutions,” J Chem Phys, 40, 1604 (1964).

[37] H. C. Flemming and J. Wingender “The biofilm matrix,” Nature Rev. Micro-
biol. 8, 623–33 (2010).

[38] D. G. Davies, M. R. Parsek, J. P. Pearson, B. H. Iglewski, J. W. Costerton, E. P.


Greenberg “The Involvement of Cell–to–Cell Signals in the Development of
a Bacterial Biofilm,” Science 280(5361), 295–8 (1998).

[39] H. C. Flemming, J. Wingender, U. Szewzyk, P. Steinberg, S.A. Rice, and S.


Kjelleberg, “Biofilms: an emergent form of bacterial life” Nat. Rev. Micro.
14(9), 563–75 (2016).

[40] U. Römling and C. Balsalobre “Biofilm infections, their resilience to therapy


and innovative treatment strategies,” Intern Med. 272(6) (2012).

[41] X. Shi and X. Zhu “Biofilm formation and food safety in food industries,”.
Trends Food Sci. Technol. 20, 407–13 (2009).

103
BIBLIOGRAPHY

[42] R. A. N. Chmielewski and J. F. Frank “Biofilm Formation and Control in


Food Processing Facilities,” Compr. Rev. Food Sci. Food Saf. 2(1), 22–3
(2003).

[43] C. Schaudinn, A. Gorur, D. Keller, P. P. Sedghizadeh, J. W. Costerton “Peri-


odontitis: an archetypical biofilm disease,” J Am Dent Assoc. 140(8), 978–
86 (2009).

[44] R. M. Donlan “Biofilms: Microbial Life on Surfaces,” Emerg. Infect. Dis. 8,


881–90 (2002).

[45] U. Szewzyk, R. Szewzyk, W. Manz, K. H. Schleifer “Microbiological safety of


drinking water,” Annu Rev Microbiol 54, 81–127 (2000).

[46] I. B. Beech, and J. Sunner “Biocorrosion: towards understanding inter-


actions between biofilms and metals,” Curr. Opin. Biotechnol. 15, 181–6
(2004).

[47] K. J. Koski and J. L. Yarger “Brillouin imaging, ” Appl. Phys. Lett. 87, 061903
(2005).

[48] T. Sandén, G. Persson, P. Thyberg, H. Blom and J. Widengren “Monitor-


ing kinetics of highly environment sensitive states of fluorescent molecules
by modulated excitation and time-averaged fluorescence intensity record-
ing,” Anal. Chem. 79, 3330–41 (2007).

[49] T. Sandén, G. Persson and J. Widengren “Transient state imaging for mi-
croenvironmental monitoring by laser scanning microscopy,” Anal. Chem.
80, 9589–96 (2008).

[50] J. P. Boon, and P. A. Fleury in Prigogine and Rice (ed.) “The Spectrum of
Light Scattered by Fluids,” Adv. Chem Phys. XXIV, Academic Press, New
York (1973).

[51] T. Still “High Frequency Acoustics in Colloid-Based Meso- and Nanostruc-


tures by Spontaneous Brillouin Light Scattering,” Springer–Verlag Berlin
Heidelberg ISBN: 978-3-642-13483-8 (2010).

[52] S. Reiß, G. Burau, O. Stachs, R. Gutho, and H. Stolz “Spatially resolved Bril-
louin spectroscopy to determine the rheological properties of the eye lens,”
Biomed. Opt. Express 2(8), 2144–59 (2011).

[53] L. Brillouin “Diffusion de la Lumiére et des Rayonnes X par un Corps


Transparent Homogéne; Influence del’Agitation Thermique”, Annales des
Physique 17, 88 (1922).

[54] V. I. Kovalev and R. G. Harrison, “Suppression of stimulated Brillouin scat-


tering in high–power single-frequency fiber amplifiers,” Opt. Lett. 31(2),
161 (2006).

104
BIBLIOGRAPHY

[55] J. M. Vaughan, and J. T. Randall “Brillouin scattering, density and elastic


properties of the lens and cornea of the eye,” Nature 284, 489–91 (1980).

[56] J. R. Lakowicz “Principles of Fluorescence Spectroscopy, ” 3rd Edition,


Springer ISBN: 978-0-387-31278-1 (2006).

[57] J. M. Dixon, M. Taniguchi and J. S. Lindsey, “PhotochemCAD 2. A Refined


Program with Accompanying Spectral Databases for Photochemical Calcu-
lations,” Photochem. Photobiol. 81, 212–3 (2005).

[58] B. Valeur and M. N. Berberan–Santos “Molecular Fluorescence: Principles


and Applications,” 2nd Edition, Wiley–VCH Verlag & Co, SBN: 978-3-527-
32837-6 (2012).

[59] J. W. Costerton, G. G. Geesey, and K. J. Cheng “How bacteria stick,” Sci. Am.
238, 86–95 (1978).

[60] H. Lappin–Scott, S. Burton and P. Stoodley “Revealing a world of biofilms,


– the pioneering research of Bill Costerton” Nature Reviews Microbiology
12, 781–7 (2014).

[61] N. Hoiby, O. Ciofu and T. Bjarnsholt “Pseudomonas aeruginosa biofilms in


cystic fibrosis,” Future Microbiology 5, 1663–74 (2010).

[62] H. V. Nielsen, A. L. Flores–Mireles, A. L. Kau, K. A. Kline, J. S. Pinkner “Pilin


and Sortase Residues Critical for Endocarditis– and Biofilm– Associated
Pilus Biogenesis in Enterococcus faecalis,” Journal of Bacteriology 19, 4484–
95 (2013).

[63] M. Rousseau, H. M. Sharon Goh, S. Holec, M. L. Albert, R. B. H. Williams, A.


Molly Ingersoll, and K. A. Kline “Bladder catheterization increases suscep-
tibility to infection that can be prevented by prophylactic antibiotic treat-
ment,” JCI Insight. 1(15), e88178 (2016).

[64] T. Bjarnsholt “The role of bacterial biofilms in chronic infections,” APMIS


Suppl. 136, 1–51 (2013).

[65] Y. Deng and W. Lv “Biofilms and Implantable Medical Devices – 1st Edi-
tion,” Elsevier, ISBN: 9780081003824 (2014).

[66] D. Pavithra, and M. Doble “Biofilm formation, bacterial adhesion and host
response on polymeric implants–issues and prevention,” Biomed Mater.
3(3), 034003 (2008).

[67] S. Srey, I. K. Jahid, and S.–D. Ha “Biofilm formation in food industries: a


food safety concern,” Food Control 31 572–85 (2013).

[68] M. S. Doggett. “Characterization of fungal biofilms within a municipal wa-


ter distribution system,” Appl. Environ. Microbiol. 66, 1249–51 (2000).

105
BIBLIOGRAPHY

[69] S. J. Edwards and B. V. Kjellerup “Applications of biofilms in biore-


mediation and biotransformation of persistent organic pollu-
tants,pharmaceuticals/personal care products, and heavy metals,” Appl.
Microbiol. Biotechnol. 97:9909–9921 (2013).

[70] S. Sehar and I. Naz “Role of the Biofilms in Wastewater Treatment, Micro-
bial Biofilms - Importance and Applications,”, Dr. D. Dhanasekaran (Ed.),
InTech, DOI: 10.5772/63499 (2016).

[71] M. Whiteley, M. G. Bangera, R. E. Bumgarner, M. R. Parsek, G. M. Teitzel,


S. Lory, and E. P. Greenberg “Gene expression in Pseudomonas aeruginosa
biofilms,” Nature 413(6858), 860–4 (2001).

[72] X. Hu and J. Ahn “Differential gene expression in planktonic and biofilm


cells of multiple antibiotic–resistant Salmonella Typhimurium and Staphy-
lococcus aureus,” FEMS Microbiol Lett 325, 180–8 (2011).

[73] M. E. Davey, and G. A. O’Toole “Microbial biofilms: From ecology to molec-


ular genetics,” Microbiol. Mol. Biol. Rev. 64, 847–67 (2000).

[74] G. A. O’Toole “A resistance switch,” Nature 416(18), 495–6 (2002).

[75] K. Sauer, A. K. Camper, G. D. Ehrlich, J. W. Costerton, D. G. Davies “Pseu-


domonas aeruginosa displays multiple phenotypes during development as
a biofilm, ” J. Bacteriol. 184, 1140–54 (2002).

[76] O. Orgad, Y. Oren, S. L. Walker, and M. Herzberg “The role of alginate in


Pseudomonas aeruginosa EPS adherence, viscoelastic properties and cell
attachment,” Biofouling, 27(7), 787–98 (2011).

[77] J. Liu, A. Prindle, J. Humphries, M. Gabalda–Sagarra et al. “Metabolic co–


dependence gives rise to collective oscillations within biofilms,” Nature
30(523) 550–4 (2015).

[78] C. P. Lim, P. N. Q. Mai, D. Roizman Sade, Y. C. Lam and Y. Cohen “Biofilm


development of an opportunistic model bacterium analysed at high spa-
tiotemporal resolution in the framework of a precise flow cell,âĂİ npj
Biofilms and Microbiomes 2 (2016).

[79] L. Ma, K. D. Jackson, R. M. Landry, M. R. Parsek, and D. J. Wozniak “Analy-


sis of Pseudomonas aeruginosa conditional psl variants reveals roles for the
psl polysaccharide in adhesion and maintaining biofilm structure postat-
tachment,” J. Bacteriol. 188, 23, 8213–21 (2006).

[80] S. C. Chew, B. Kundukad, T. Seviour, J. R. van der Maarel, L. Yang, S. A. Rice,


P. Doyle ,and S. Kjelleberg “Dynamic remodeling of microbial biofilms by
functionally distinct exopolysaccharides,” mBIO e01536–14, 1–11 (2014).

[81] K. N. Kragh, J. B. Hutchison, G. Melaugh et al. “Role of multicellular aggre-


gates in biofilm formation,” mBio 7, e00237–16 (2016).

106
BIBLIOGRAPHY

[82] N. Barraud, D. J. Hassett, S.–H. Hwang, S. A. Rice, S. Kjelleberg, J. S. Webb


“Involvement of nitric oxide in biofilm dispersal of Pseudomonas aerugi-
nosa,” J. Bacteriol. 188, 7344–53 (2006).

[83] S. An, J. E. Wu and L.–H. Zhang “Modulation of Pseudomonas aeruginosa


biofilm dispersal by a cyclic–Di–GMP phosphodiesterase with a putative
hypoxia–sensing domain,” Appl Environ Microbiol. 76, 8160–73, (2010).

[84] J.B. Kaplan “Biofilm dispersal: mechanisms, clinical implications, and po-
tential therapeutic uses, ” J. Dent. Res. 89(3), 205–218 (2010).

[85] L. Ma, M. Conover, H. Lu, M. R. Parsek, K. Bayles, D. J. Wozniak “Assembly


and Development of the Pseudomonas aeruginosa Biofilm Matrix, ” PLOS
Pathogens 5(3) (2009).

[86] J. S. Webb, L.S. Thompson, S. James, T. Charlton, T. Tolker–Nielsen, B.


Koch, M. Givskov, and S. Kjelleberg “Cell death in Pseudomonas aeruginosa
biofilm development, ” J. Bacteriol. 185(15), 4585–92 (2003).

[87] S. M. Hunt, E. M. Werner, B. Huang, M. A. Hamilton, and P. S. Stewart “Hy-


pothesis for the role of nutrient starvation in biofilm detachment, ” Appl.
Environ. Microbiol. 70, 7418–25 (2004).

[88] R. Oliveira, L. Melo, A. Oliveira, R. Salgueiro “Polysaccharide production


and biofilm formation by Pseudomonas fluorescens: effects of pH and sur-
face material,” Colloids and Surfaces B: Biointerfaces 2(1): 41–6 (1994).

[89] B. Prakash, B. M. Veeregowda, G. Krishnappa “Biofilms: a survival strategy


of bacteria. Current Science,” 85(9): 1299–307 (2003).

[90] C. H. Tan, K. W. K. Lee, M. Burmolle, S. Kjelleberg, and S. A. Rice “All to-


gether now: experimental multispecies biofilm model systems” Env. Mi-
crobiology 19(1), 42–53 (2017).

[91] J. W. Costerton, K.-J. Cheng, G. G. Geesey, T. I. Ladd, J. C. Nickel, M. Das-


gupta, and T. J. Marrie “Bacterial biofilms in nature and disease,” Annu.
Rev. Microbiol. 41, 435–64 (1987).

[92] H. Anwar, J. L. Strap, and J. W. Costerton “Establishment of aging biofilms:


possible mechanism of bacterial resistance to antimicrobial therapy,” An-
timicrob. Agents Chemother. 36, 1347–51 (1992).

[93] P. Watnick, and R. Kolter “Biofilm, city of microbes,” J. Bacteriol. 182, 2675–
9 (2000).

[94] M. G. Mazza “The physics of biofilms - an introduction,” J. Phys. D: Appl.


Phys. 49(20), 203001 (2016).

[95] H. C. Flemming, T. R. Neu, and D. J. Wozniak “The EPS Matrix: The “House
of Biofilm Cells” J. Bacteriol. 189(22), 7945–7 (2007).

107
BIBLIOGRAPHY

[96] H. C. Flemming, and J. Wingender “Extracellular polymeric substances:


structure, ecological functions, technical relevance,” in G. Bitton (ed.), En-
cyclopedia of environmental microbiology, vol. 3. Wiley, New York, NY, pp
1223–31 (2002).

[97] D. McDougald, S. A. Rice, N. Barraud, P. D. Steinberg, and S. Kjelleberg


“Should we stay or should we go: mechanisms and ecological conse-
quences for biofilm dispersal,” Nature reviews. Microbiology. 10, 39–50
(2012).

[98] P. Stoodley, Z. Lewandowski, J. D. Boyl, and H. M Lappin-Scott “Struc-


tural deformation of bacterial biofilms caused by short-term fluctuations
in fluid shear: An in situ investigation of biofilm rheology, ” Biotechnol.
Bioeng. 65, 83–92 (1999).

[99] L. Hall–Stoodley, P. Stoodley “Evolving concepts in biofilm infections,” Cell


Microbiol. 11(7), 1034–43 (2009).

[100] M. Tallawi, M. Opitz and O. Lieleg “Modulation of the mechanical prop-


erties of bacterial biofilms in response to environmental challenges,” Bio-
mater. Sci. 5, 887–900 (2017).

[101] O. Lieleg, M. Caldara, R. Baumgaartel, and K. Ribbeck, “Mechanical ro-


bustness of Pseudomonas aeruginosa biofilms,” Soft Matter, 7, 3307–14
(2011).

[102] P. S. Stewart and M.J. Franklin “Physiological heterogeneity in biofilms,”


Nature reviews. Microbiology 6, 199–210 (2008).

[103] S. Kalmbach, W. Manz, and U. Szewzyk “Isolation of new bacterial species


from drinking water biofilms and proof of their in situ dominance with
highly specific 16S rRNA probes,” Appl. Environ. Microbiol. 63, 4164–70
(1997).

[104] I. W. Sutherland “The biofilm matrix – an immobilized but dynamic mi-


crobial environment,” Trends in microbiology 9, 222–7 (2001).

[105] B. Kundukad, T. Seviour, Y. Liang, S. A. Rice, S. Kjelleberg, and P. S. Doyle


“Mechanical properties of the superficial biofilm layer determine the ar-
chitecture of biofilms, ” Soft Matter, 12, 5718 (2016).

[106] P. Stoodley, R. Cargo, C. J. Rupp, S. Wilson, and I. Klapper “Biofilm ma-


terial properties as related to shear–induced deformation and detachment
phenomena,” J. Ind. Microbiol. Biotechnol. 29, 361–7 (2003).

[107] S. Chatterjee, N. Biswas, A. Datta, R. Dey, and P. Maiti “Atomic force mi-
croscopy in biofilm study,” Microscopy (Oxf), 63(4), 269–78 (2014).

108
BIBLIOGRAPHY

[108] O. Galy, P. Latour–Lambert, K. Zrelli, J.-M. Ghigo, C. Beloin, and N. Henry


“Mapping of bacterial biofilm local mechanics by magnetic microparticle
actuation, ” Biophys. J. 103(6), 1400–8 (2012).

[109] S. Aggarwal and R. M. Hozalski “Effect of strain rate on the mechanical


properties of Staphylococcus epidermidis biofilms, ” Langmuir 28, 2812-6
(2012).

[110] A. Cense, E. Peeters, B. Gottenbos, F. Baaijens, A. Nuijs, and M. Van Don-


gen “Mechanical properties and failure of Streptococcus mutans biofilms ,
studied using a microindentation device,” J. Microbiol. Methods. 67, 463–
72 (2006).

[111] P. C. Y. Lau, J. R. Dutcher, T. J. Beveridge, and J. C. Lam “Absolute quanti-


tation of bacterial biofilm adhesion and viscoelasticity by microbead force
spectroscopy,” Biophys. J. 96, 2935–48 (2009).

[112] K. Forier, A. S. Messiaen, K. Raemdonck, H. Deschout, J. Rejman, F. de


Baets, H. Nelis, S.C. de Smedt, J. Demeester, T. Coenye, and K. Braeckmans
“Transport of nanoparticles in cystic fibrosis sputum and bacterial biofilms
by single–particle tracking microscopy,” Nanomedicine (Lond) 8(6), 935–
49 (2013).

[113] N. Billings, A. Birjiniuk, T. S. Samad, P. S. Doyle, and K. Ribbeck “Material


properties of biofilms – key methods for understanding permeability and
mechanics, ” Rep Prog Phys. 78(3) (2015).

[114] P. So “Microscopy: Brillouin bioimaging, ” Nat. Photonics, 2(1), 13–4


(2008).

[115] T. Fenchel and B. Finlay “Oxygen and the Spatial Structure of Microbial
Communities,” Biol. Rev. 83, 553–69 (2008).

[116] A. Moya, X. Guimera, F. J. del Campo et al. “Profiling of oxygen in biofilms


using individually addressable disk microelectrodes on a microfabricated
needle,” Microchimica Acta 182(5–6), 985–93 (2015).

[117] D. de Beer, P. Stoodley, F. Roe and Z. Lewandowski. “Effects of biofilm


structures on oxygen distribution and mass transport,” Biotechnol. Bioeng.
43, 1131–8 (1994).

[118] G. Borriello, L. Richards, G. D. Ehrlich and P. S. Stewart. “Arginine or nitrate


enhances antibiotic susceptibility of Pseudomonas aeruginosa in biofilms
Antimicrobial agents and chemotherapy,” 50, 382–4 (2006).

[119] R. N. Glud, C. M. Santegoeds, D. de Beer, O. Kohls and N. B. Ramsing “Oxy-


gen dynamics at the base of a biofilm studied with planar optodes,” Aquatic
Microbial Ecology 14, 223–33 (1998).

109
BIBLIOGRAPHY

[120] K. M. Thormann, R. M. Saville, S. Shukla and A. M. Spormann “Induction


of Rapid Detachment in Shewanella oneidensis MR-1 Biofilms,” J. Bacteriol
187, 1014–21 (2005).

[121] N. P. Revsbech, and B. B. Jørgensen “Microelectrodes: Their Use in Mi-


crobial Ecology” in K. C. Marshall (ed.), Advances in Microbial Ecology,
Springer US, pp 293–352 (1986).

[122] O. S. Wolfbeis “Luminescent sensing and imaging of oxygen: Fierce com-


petition to the Clark electrode,” Bioessays 37, 921–8 (2015).

[123] R. N. Glud, J. K. Gundersen, N. P. Revsbech and B. B. Jørgensen. “Effects on


the benthic diffusive boundary layer imposed by microelectrodes,” Limnol.
Oceanogr. 39, 462–7 (1994).

[124] X. D. Wang and O. S. Wolfbeis “Optical methods for sensing and imaging
oxygen: materials, spectroscopies and applications,” Chemical Society Re-
views 43, 3666–761 (2014).

[125] J. R. Lakowicz and G. Weber “Quenching of fluorescence by oxygen. Probe


for structural fluctuations in macromolecules,” Biochemistry 12(21), 4161–
70 (1973).

[126] M. Staal, E. I. Prest, J. S. Vrouwenvelder, L. F. Rickelt and M. Kühl “A sim-


ple optode based method for imaging O2 distribution and dynamics in tap
water biofilms,” Water Research 45, 5027–37 (2011).

[127] R. N. Glud, N. B. Ramsing, J. K. Gundersen and I. Klimant “Planar op-


trodes: a new tool for fine scale measurements of two-dimensional O2
distribution in benthic communities,” Marine Ecology Progress Series 140,
217–26 (1996).

[128] M. Kühl, L. F. Rickelt and R. Thar “Combined Imaging of Bacteria and


Oxygen in Biofilms,” Applied and Environmental Microbiology 73, 6289–
95 (2007).

[129] M. Larsen, S. M. Borisov, B. Grunwald, I. Klimant and R. N. Glud “A sim-


ple and inexpensive high resolution color ratiometric planar optode imag-
ing approach: application to oxygen and pH sensing,” Limnology and
Oceanography: Methods 9, 348–60 (2011).

[130] G. Holst, O. Kohls, I. Klimant, B. König, M. Kühl and T. Richter “A modular


luminescence lifetime imaging system for mapping oxygen distribution in
biological samples,” Sensors and Actuators B: Chemical 51, 163–70 (1998).

[131] J. Huisken and D. Y. R. Stainier “Selective plane illumination microscopy


techniques in developmental biology,” Development, 136, 1963–75 (2009).

110
BIBLIOGRAPHY

[132] A. Karampatzakis, C. Z. Song, L. P. Allsopp, A. Filloux, S. A. Rice, Y. Cohen,


T. Wohland and P. Török “Probing the internal micromechanical proper-
ties of Pseudomonas aeruginosa biofilms by Brillouin imaging” npj Biofilms
and Microbiomes, 3:20, 1–7, (2017).

[133] J. Jonkman, and C. M. Brown “Any Way You Slice It – A Comparison of Con-
focal Microscopy Techniques,” Journal of Biomolecular Techniques, 26(2),
54–65 (2015).

[134] C. M. St. Croix, S. H. Shand, S. C. Watkins “Confocal microscopy: compar-


isons, applications, and problems,” BioTechniques 39, S2–S5 (2005).

[135] G. H. Patterson, D. W. Piston “Photobleaching in Two–Photon Excitation


Microscopy,” Biophysical Journal 78(4), 2159–62 (2000).

[136] E. Gross “Change of wavelength of light due to elastic heat waves at scat-
tering of liquids,” Nature 126, 201–2 (1930).

[137] R. Y. Chiao and B. P. Stoicheff, “Brillouin scattering in liquids excited by


the He–Ne maser,” J. Opt. Soc. Am. 54, 1286–7 (1964).

[138] G. B. Benedek, J. B. Lastovska, K. Fritsch, and T. Greytak, “Brillouin scat-


tering in liquids and solids using low–power lasers,” J. Opt. Soc. Am. 54,
1284–5 (1964).

[139] S.M. Shapiro, R. W. Gammon, and H. Z. Cummins “Brillouin scattering


spectra of crystalline quartz, fused quartz and glass,” Appl. Phys. Lett. 9,
157–9 (1966).

[140] R. Vacher and L. Boyer, “Brillouin scattering: A tool for the measurement
of elastic and photoelastic constants,” Phys. Rev. B 6(2), 639–73 (1972).

[141] J. D. Bass “Elasticity of grossular and spessartite garnets by Brillouin spec-


troscopy,”. J Geophys. Res. B, Solid Earth and Planets 94, 7621–8 (1989).

[142] G. D. Patterson “Light scattering from bulk polymers,” Annu. Rev. Mater.
Sci. 13, 219–45 (1983).

[143] S. A. Lee et al., “Elastic and photoelastic anisotropy of solid hf at high pres-
sure,” Phys. Rev. B, 34(4), 2799–806 (1986).

[144] S. K. Satija and C. Wang, “Brillouin scattering study of elastic anomaly


in (KCN) x (KBr) 1–x system,” Solid State Communications, 28(8) 617–9
(1978).

[145] R. Harley, D. James, A. Miller, and J. W. White “Phonons and the elastic
moduli of collagen and muscle,” Nature 267(5608), 285–7 (1977).

[146] S. Cusack, and A. Miller “Determination of the elastic constants of colla-


gen by Brillouin light scattering,” J. Mol. Biol. 135, 39–51 (1979).

111
BIBLIOGRAPHY

[147] S. J. Randall, and J. M. Vaughan “Brillouin scattering in systems of biolog-


ical significance,” Phil. Trans. R. Soc. Lond. A 293, 1402, 341–348 (1979).

[148] S. Lees, N. J. Tao, and S. M. Lindsay “Studies of compact hard tissues and
collagen by means of Brillouin light scattering,” Connect. Tissue Res. 24,
187–205 (1990).

[149] J. R. Sandercock “Some recent developments in Brillouin scattering, ” RCA


Rev. 36, 89–107 (1975).

[150] J. R. Sandercock, “Brillouin scattering study of SbSi using a double–


passed, stabilised scanning interferometer,” Opt. Commun. 2, 73–6 (1970).

[151] N. Berovic, N. Thomas, R. A. Thornhill, and J. M. Vaughan, “Observation


of Brillouin scattering from single muscle fibres,” Eur. Biophys. J. 17, 69–74
(1989).

[152] A. V. Svanidze, V. P. Romanov, and S. G. Lushnikov “Anomalous Behavior


of Brillouin Light Scattering at Thermal Denaturation of Lysozyme,” JETP
Letters 93(7) (2011).

[153] K. J. Koski, P. Akhenblit, K. McKiernan, and J. L. Yarger “Non–invasive de-


termination of the complete elastic moduli of spider silks,” Nature Materi-
als 12, 262–7 (2013).

[154] G. Scarcelli, and S. H. Yun “Confocal Brillouin microscopy for three–


dimensional mechanical imaging, ” Nat. Photon. 2, 39–43 (2008).

[155] M. Shirasaki “Large angular dispersion by a virtually imaged phased array


and its application to a wavelength demultiplexer, ” Opt. Lett. 21(5), 366–8
(1996).

[156] G. Scarcelli, P. Kim, and S. H. Yun “In vivo measurement of age–related


stiffening in the crystalline lens by Brillouin optical microscopy,” Biophys.
J. 101(6), 1539–45 (2011).

[157] Z. Meng, A. J. Traverso, and V. V. Yakovlev “Background clean–up in Bril-


louin microspectroscopy of scattering medium, ” Opt. Express 22(5), 144–8
(2014).

[158] A. J. Traverso, J. V. Thompson, Z. A. Steelman, Z. Meng, M. O. Scully, and V.


V. Yakovlev “Dual Raman–Brillouin Microscope for Chemical and Mechan-
ical Characterization and Imaging, ” Anal. Chem. 87, 7519–23 (2015).

[159] A. Fiore, J. Zhang, P. Shao, S. H. Yun, and G. Scarcelli “High-extinction vir-


tually imaged phased array-based Brillouin spectroscopy of turbid biolog-
ical media, ” Appl. Phys. Lett. 108, 203701 (2016).

112
BIBLIOGRAPHY

[160] G. Antonacci, G. Lepert, C. Paterson, and P. Török “Elastic suppression


in Brillouin imaging by destructive interference,” Appl. Phys. Lett. 107,
061102 (2015).

[161] P. Shao, S. Besner, J. Zhang, G. Scarcelli, and S. H. Yun “Etalon filters for
Brillouin microscopy of highly scattering tissues, ” Opt. Express 24(19),
4436–9 (2016).

[162] G. Lepert, R. M. Gouveia, C. J. Connon, and C. Paterson “Assessing corneal


biomechanics with Brillouin spectro-microscopy,” Farad. Discuss. 187,
415–28 (2016).

[163] G. Antonacci, M. R. Foreman, C. Paterson, and P. Török “Spectral broad-


ening in Brillouin imaging,” Appl. Phys. Lett. 103, 221105 (2013).

[164] G. Scarcelli, R. Pineda, and S. H. Yun “Brillouin optical microscopy for


corneal biomechanics, ” Invest. Ophthalmol. Vis. Sci. 53, 185–90 (2012).

[165] Z. Steelman, Z. Meng, A. J. Traverso, and V. V. Yakovlev “Brillouin spec-


troscopy as a new method of screening for increased CSF total protein dur-
ing bacterial meningitis, ” J. Biophotonics 8(5), 408–14 (2015).

[166] S. Mattana, S. Caponi, F. Tamagnini, D. Fioretto et al. “Viscoelasticity


of amyloid plaques in transgenic mouse brain studied by Brillouin mi-
crospectroscopy and correlative Raman analysis,” Journal of Innovative
Optical Health Sciences 10(5), 1742001 (2017).

[167] G. Antonacci, R. M. Pedrigi, A. Kondiboyina, V. V. Mehta, R. de Silva, C. Pa-


terson, R. Krams, and P. Török “Quantification of plaque stiffness by Bril-
louin microscopy in experimental thin cap fibroatheroma,” J. R. Soc. Inter-
face 12, 20150843 (2015).

[168] G. Scarcelli, W. J. Polacheck, H. T. Nia, K. Patel, A. J. Grodzinsky, R. D. Kam,


and S. H. Yun “Noncontact three–dimensional mapping of intracellular hy-
dromechanical properties by Brillouin microscopy, ” Nat. Meth. 12, 1132–
34 (2015).

[169] K. Elsayad, S. Werner, M. Gallemí, J. Kong, E. R. S. Guajardo et al. “Map-


ping the subcellular mechanical properties of live cells in tissues with fluo-
rescence emission – Brillouin imaging, ” Sci. Signal. 9(435), 1–13 (2016).

[170] G. Antonacci and S. Braakman “Biomechanics of subcellular structures by


non–invasive Brillouin microscopy,” Scientific Reports 6, 37217 (2016).

[171] J. Zhang, X. A. Nou, H. Kim and G. Scarcelli “Brillouin flow cytometry for
label–free mechanical phenotyping of the nucleus,” Lab Chip 17, 663–70
(2017).

113
BIBLIOGRAPHY

[172] Z. Meng. S. C. Bustamante Lopez, K. E. Meissner, and V. V. Yakovlev “Sub-


cellular measurements of mechanical and chemical properties using dual
Raman–Brillouin microspectroscopy,” J. Biophotonics. 9(3), 201–7 (2016).

[173] P.–J. Wu, I. Kabakova, J. Ruberti, I. Dunlop, C. Paterson, P. Török and D.


R. Overby “Brillouin microscopy, what is it really measuring?,” Submitted,
under review (2017).

[174] Y. Zhang, C. K. Ng, Y. Cohen and B. Cao “Cell growth and protein ex-
pression of Shewanella oneidensis in biofilms and hydrogel–entrapped cul-
tures,” Molecular BioSystems 10, 1035–42 (2014).

[175] L. Lambertsen, C. Sternberg, and S. Molin “Mini–Tn7 transposons for site–


specific tagging of bacteria with fluorescent proteins,” Environ. Microbiol.
6(7), 726–32 (2004).

[176] D. H. Figurski and D. R. Helinski “Replication of an origin–containing


derivative of plasmid RK2 dependent on a plasmid function provided in
trans,” Proc. Natl. Acad. Sci. U S A 76(4), 1648–52 (1979).

[177] C. Sternberg, and T. Tolker–Nielsen. “Growing and Analyzing Biofilms in


Flow Cells.” In: R. Coico, T. Kowalik, J. Quarles, B. Stevenson, R. Taylor, ed-
itors. “Current Protocols in Microbiology”. New Jersey, JohnWiley. 1B. 2.1–
1B. 2.15. 23 (2005).

[178] S. J. Pamp, C. Sternberg, and T. Tolker–Nielsen “Insight into the microbial


multicellular lifestyle via flow–cell technology and confocal microscopy, ”
Cytometry Part A 75(A), 90–103 (2009).

[179] R. G. Lyons “Understanding digital signal processing,”3rd Ed. Prentice


Hall, ISBN 0-13-702741-9 (1997).

[180] B. Cao, L. Christophersen, M. Kolpen, P. Ø. Jensen, K. Sneppen, N. Høiby,


C. Moser, and T. Sams “Diffusion Retardation by Binding of Tobramycin in
an Alginate Biofilm Model,” PLOS ONE, DOI:10.1371 (2016).

[181] L. J. Christophersen, H. Trøstrup, D. S. Malling Damlund, T. Bjarnsholt, K.


Thomsen, P. Ø. Jensen, H. P. Hougen, N. Høiby and C. Moser “Bead–size di-
rected distribution of Pseudomonas aeruginosa results in distinct inflam-
matory response in a mouse model of chronic lung infection,”. Clin. Exp.
Immunol 170(2), 222–30 (2012).

[182] R. Bakke and P. Q. Olsson “Biofilm thickness measurements by light mi-


croscopy,” Journal of Microbiological Methods 5, 93–8 (1986).

[183] R. Bakke, R. Kommedal, and S. Kalvenes “Quantification of biofilm accu-


mulation by an optical approach,” Journal of Microbiological Methods 44,
13–26 (2001).

114
BIBLIOGRAPHY

[184] S. W. Hermanowicz and J. J. Ganczarczyk “Some Fluidization Character-


istics of Biological Beds,” Biotechnology and Bioengineering, 25, 1321–30
(1983).

[185] K. S. Ro, J. B. Neethling “Biofilm density for biological fluidized beds,” Re-
search Journal WPCF, 63(5), 815–8 (1991).

[186] M. Tsezos and A. Benedekt “A Method For The Calculation Of Biological


Film Volume In A Fluidized Bed Biological Reactor,” Water Research 14,
680–93 (1980).

[187] J. Schmitt and H.C. Flemming “Water binding in biofilms,” Water Science
and Technology 39, 77–82 (1999).

[188] T. C. Zhang and P. L. Bishop “Density, Porosity, And Pore Structure Of


Biofilms,” Wat. Res. 28(11), 2267–77 (1994).

[189] J. Yan, G. A. G. Sharo, H. A. Stone, N. S. Wingreen, and B. L. Bassler “Vibrio


cholerae biofilm growth program and architecture revealed by single-cell
live imaging, ” Proc. Natl. Acad. Sci. U. S. A. E5337–43 (2016).

[190] B. Purevdorj–Gage, W. J. Costerton, P. and Stoodley, “Phenotypic differ-


entiation and seeding dispersal in non–mucoid and mucoid Pseudomonas
aeruginosa biofilms,” Microbiology 151, 1569–7676 (2005).

[191] S. M. Kirov, J. S. Webb, C. Y. O’may, D. W. Reid, J. K. Woo, S. A. Rice, and S.


Kjelleberg “Biofilm differentiation and dispersal in mucoid Pseudomonas
aeruginosa isolates from patients with cystic fibrosis,” Microbiology 153,
3264–74 (2007).

[192] J. Liu, R. Martinez–Corral, A. Prindle et al. “Coupling between distant


biofilms and emergence of nutrient time-sharing,” Science 356(6338), 638–
42 (2017).

[193] A. Persat, C. D. Nadell, M. K. Kim et al. “The Mechanical World of Bacte-


ria,” Cell 161(5), 988–97 (2015).

[194] G. Bodelón, V. Montes–Garcia, V. López–Puente et al. “Detection and


imaging of quorum sensing in Pseudomonas aeruginosa biofilm commu-
nities by surface–enhanced resonance Raman scattering,” Nature Materi-
als 15, 1203–11 (2016).

[195] R. N. Masyuko, E. J. Lanni, C. M. Driscoll et al. “Spatial organization of


Pseudomonas aeruginosa biofilms probed by combined matrix-assisted
laser desorption ionization mass spectrometry and confocal Raman mi-
croscopy,” Analyst. 139(22), 5700–8 (2014).

115
BIBLIOGRAPHY

[196] A. Karampatzakis, J. Sankaran, K. Kandaswamy, S. A. Rice, Y. Cohen and


T. Wohland “Measurement of oxygen concentrations in bacterial biofilms
using transient state monitoring by single plane illumination microscopy”
Biomed. Phys. Eng. Express 3 035020 (2017).

[197] T. A. Planchon, L. Gao, D. E. Milkie, M. W. Davidson, J. A. Galbraith, C.


G. Galbraith, E. Betzig “Rapid three–dimensional isotropic imaging of liv-
ing cells using Bessel beam plane illumination,” Nat Methods, 8(5) 417–23
(2011).

[198] B–C. C. Chen, W. R. Legant, K. Wang et al. “Lattice light-sheet microscopy:


Imaging molecules to embryos at high spatiotemporal resolution,” Sci-
ence, 346 (2015).

[199] P. J. Keller, F. Pampaloni, and E. H. Stelzer, “Life sciences require the third
dimension,” Curr. Opin. Cell Biol. 18, 117–24 (2006).

[200] J. G. Ritter, R. Veith, J. P. Siebrasse and U. Kubitscheck “High-contrast


single-particle tracking by selective focal plane illumination microscopy,”
Opt Express 16(10), 7142–53 (2008).

[201] T. Wohland, X. Shi, J. Sankaran and E. H.K. Stelzer “Single Plane Illumi-
nation Fluorescence Correlation Spectroscopy (SPIM–FCS) probes inho-
mogeneous three-dimensional environments,” Opt. Exp. 18(10), 10627–41
(2010).

[202] A. P. Singh, T. Wohland, J. W. Krieger, J. Buchholz, E. Charbon and J. Lan-


gowski “The performance of 2D array detectors for light sheet based fluo-
rescence correlation spectroscopy,” Opt. Exp. 21(7), 8652–68 (2013).

[203] J. W. Krieger, A. P. Singh, N. Bag, C. S. Garbe, T. E. Saunders, J. Langowski


and T. Wohland “Imaging fluorescence (cross-) correlation spectroscopy in
live cells and organisms,” Nature Protocols 10, 1948–74 (2015).

[204] X. W. Ng, C. The, V. Korzh and T. Wohland “The Secreted Signaling Protein
Wnt3 Is Associated with Membrane Domains In Vivo : A SPIM-FCS Study,”
Biophysical Journal 111, 418–29 (2016).

[205] K. Greger, M. J. Neetz, E. G. Reynaud, and E. H. K. Stelzer “Three-


dimensional Fluorescence Lifetime Imaging with a Single Plane Illumina-
tion Microscope provides an improved Signal to Noise Ratio,” Optics Ex-
press. 19(21), 20743–50 (2011).

[206] A. Costa, A. Candeo, L. Fieramonti, G. Valentini and A. Bassi. “Calcium


Dynamics in Root Cells of Arabidopsis thaliana Visualized with Selective
Plane Illumination Microscopy,” PLoS One. 8(10), e75646 (2013).

[207] E. Stelzer “Light–sheet fluorescence microscopy for quantitative biology,”


Nature Methods 12, 23–6 (2015).

116
BIBLIOGRAPHY

[208] V. E. Korobov and A. K. Chibisov “Primary processes in the photochem-


istry of rhodamine dyes,” J of Photochemistry 9(4), 411–24 (1978).

[209] G. Marriott, R. M. Clegg, D. J. Arndt–Jovin, and T. M. Jovin “Time re-


solved imaging microscopy. Phosphorescence and delayed fluorescence
imaging,” Biophys J. 60(6), 1374–87 (1991).

[210] J. Widengren, U. Mets, R. Rigler. “Fluorescence correlation spectroscopy


of triplet states in solution: a theoretical and experimental study,” J. Phys.
Chem. 99, 13368–79 (1995).

[211] M. Geissbuehler, T. Spielmann, A. Formey, I. Marki, M. Leutenegger, B.


Hinz, K. Johnsson, D. van de Ville and T. Lasser “Triplet imaging of oxygen
consumption during the contraction of a single smooth muscle cell (A7r5),”
Biophysical Journal 98, 339–49 (2010).

[212] T. Spielmann, H. Blom, M. Geissbuehler, T. Lasser and J. Widengren


“Transient state monitoring by total internal reflection fluorescence mi-
croscopy,” J. Phys. Chem. B. 114, 4035–46 (2010).

[213] J. Mücksch, T. Spielmann, E. Sisamakis and J. Widengren “Transient state


imaging of live cells using single plane illumination and arbitrary duty cycle
excitation pulse trains,” J. Biophotonics 8(5), 392–400 (2014).

[214] M. Geissbuehler, Z. Kadlecova, H.-A. Klok, and T. Lasser “Assessment of


transferrin recycling by Triplet Lifetime Imaging in living cells,” Biomedical
optics express. 3, 2526–36 (2012).

[215] T. Spielmann, L. Xu, A. K. B. Gad, S. Johansson and J. Widengren “Tran-


sient state microscopy probes patterns of altered oxygen consumption in
cancer cells,” The FEBS Journal 281(5), 1317–32 (2014).

[216] A. Penzkofer, and A. Beidoun “Triplet–triplet absorption of Eosin Y in


methanol determined by nanosecond excimer laser excitation and pi-
cosecond light continuum probing,” Chem. Phys 177(1), 203–16 (1993).

[217] A. Deshpande and N. Iyer “Correlation of fluorescence and lasing proper-


ties in Eosin Y,” J. of Luminescence. 46, 339–344 (1990).

[218] S. Reindl and A. Penzkofer “Triplet quantum yield determination by pi-


cosecond laser double–pulse fluorescence excitation,” Chem. Phys. 213,
429–38 (1996).

[219] A. Penzkofer, A. Beidoun and M. Daiber “Intersystem-crossing and


excited-state absorption in Eosin Y solutions determined by picosecond
double pulse transient absorption measurements,” J. Lumin. 51, 297–314
(1992).

117
BIBLIOGRAPHY

[220] R. Nilsson, P. Merkel and D. Kearns “Kinetic properties of the triplet states
of methylene blue and other photosensitizing dyes, ” Photochemistry and
photobiology. 16(2), 109–16 (1972).

[221] J. M. Vanderkooi, G. Maniara, T. J. Green, D. F. Wilson “An optical method


for measurement of dioxygen concentration based upon quenching of
phosphorescence,” J. Biol Chem 262(12), 5476–82 (1987).

[222] A. H. Devol “Bacterial oxygen uptake kinetics as related to biological pro-


cesses in oxygen de cient zones of the oceans,” Deep Sea Res. 25(2), 137–46
(1978).

[223] P. G. Seybold, M. Gouterman, J. Callis “Calorimetric, photometric and


lifetime determinations of fluorescence yields of fluorescein dyes,” Pho-
tochem Photobiol. 9(3), 229–42 (1969).

[224] L. W. Lo, C. J. Koch and D. F. Wilson “Calibration of oxygen–dependent


quenching of the phosphorescence of Pd–meso–tetra (4–carboxyphenyl)
porphine: a phosphor with general application for measuring oxygen con-
centration in biological systems,” Anal. Biochem. 236, 153–60 (1996).

[225] A. Karampatzakis and T. Wohland “A Matlab script for process-


ing TRAST monitoring data” (2016) [Retrieved 5 October 2016]
www.dbs.nus.edu.sg/lab/BFL/trast.html

[226] J. F. Swindells, C. F. Snyder, R. C. Hardy, P. E. Golden “Viscosities of sucrose


solutions at various temperatures: tables of recalculated values,” United
States. National Bureau of Standards.

[227] E. Guiot, P. Georges, A. Brun, M. P. Fontaine–Aupart, M. N. Bellon–


Fontaine, R. Briandet “Heterogeneity of diffusion inside microbial biofilms
determined by fluorescence correlation spectroscopy under two-photon
excitation,” Photochemistry and photobiology 75(6), 570–8 (2002).

[228] J. N. Wilking, V. Zaburdaev, M. D. Volder, R. Losick, M. P. Brenner, and


D. A. Weitz “Liquid transport facilitated by channels in Bacillus subtilis
biofilms,” PNAS 110, 848–52 (2013).

[229] M. Wagner, D. Taherzadeh, C. Haisch, H. Horn “Investigation of the


Mesoscale Structure and Volumetric Features of Biofilms Using Optical Co-
herence,” Biotechnol Bioeng. 107(5), 844–53 (2010).

[230] R. P. Briandet, P. Lacroix–Gueu, M. Renault, S. Lecart, T. Meylheuc, E. Bid-


nenko et al. “Fluorescence Correlation Spectroscopy To Study Diffusion
and Reaction of Bacteriophages inside Biofilms,” Appl. Environ. Microbiol.
74(7), 2135–43 (2008).

[231] S. Mitri, J. B. Xavier, and K. R. Foster “Social evolution in multispecies


biofilms,” PNAS 108(2), 10839–46 (2011).

118
BIBLIOGRAPHY

[232] P. N. Hedde, S. Ranjit, and E. Gratton “3D fluorescence anisotropy imag-


ing using selective plane illumination microscopy,” Optics Express 23(17),
22308–17 (2015).

[233] N. Barraud, D. Schleheck, J. Klebensberger, J. S. Webb, D. J. Hassett, S.


A. Rice, and S. Kjelleberg “Nitric oxide signaling in Pseudomonas aerug-
inosa biofilms mediates phosphodiesterase activity, decreased cyclic di–
GMP levels, and enhanced dispersal,” J. Bacteriol. 191, 7333–42 (2009).

[234] H. A. S. Nair, S. Periasamy, L. Yang, S. Kjelleberg, and S. A. Rice “Real Time,


Spatial, and Temporal Mapping of the Distribution of c–di–GMP during
Biofilm Development,” Journal of Biological Chemistry 292, 477–87 (2017).

[235] J. Y. Daher, J. Koussa, S. Younes, R. A. Khalaf “The Candida albicans Dse1


Protein Is Essential and Plays a Role in Cell Wall Rigidity, Biofilm Forma-
tion, and Virulence,” Interdiscip Perspect Infect Dis. 504280 (2011).

[236] K. Kovach, M. Davis–Fields, Y. Irie et al. “Evolutionary adaptations of


biofilms infecting cystic fibrosis lungs promote mechanical toughness by
adjusting polysaccharide production,” npj Biofilms and Microbiomes 3
(2017).

[237] J. Li, K. Hirota, T. Goto, H. Yumoto, Y. Miyake, T. Ichikawa “Biofilm for-


mation of Candida albicans on implant overdenture materials and its re-
moval,” Journal of Dentistry 40, 686–92 (2012).

119
Appendix A

Typical TRAST illumination scheme

Here we include instructions and a lookup table that can be used to recreate the

illuminating conditions for a TRAST monitoring experiment.

Steps:

1. Decide the total illumination time (Ti l l ). For Eosin Y bulk solutions in the

range of 1 – 2 µM illuminated with a 488 nm laser at ≈ 5 mW before objective,

100 ms is a good starting point.

2. Decide the period of the pulse train (T). This should be longer than the time

scale of the kinetics you want to measure. For our experiments, 50 µs is a suit-

able value.

3. Create a lookup table, similar to the one shown in Table A.1. Depending on

the number of images you wish to acquire, create a distribution of pulse widths

(w) that is denser near the time scale of the kinetics you want to measure and

extends up to the period of the pulse train. The duty cycle (α) for each image

can be calculated as α = T /w and the number of pulses for maintaining con-

A-1
APPENDIX A. TYPICAL TRAST ILLUMINATION SCHEME

stant total illumination as N = Ti l l /w.

4. For each image to be acquired, the acquisition time of your camera has to

cover the integrate the whole train of pulses, therefore it needs to be set to T ac q =

N ·T.

Image no. Pulse width Duty cycle Number of pulses Pulse train length
w (µs) α (%) N Tac q (s)
1 0.5 1 200,000 10.000
2 1 2 100,000 5.000
3 2 3 66,667 3.333
4 2 4 50,000 2.500
5 3 6 33,333 1.667
6 4 8 25,000 1.250
7 5 10 20,000 1.000
8 8 16 12,500 0.625
9 10 20 10,000 0.500
10 13 25 8,000 0.400
11 16 32 6,250 0.313
12 20 40 5,000 0.250
13 30 60 3,333 0.167
14 35 70 2,857 0.143
15 40 80 2,500 0.125
16 50 100 2,000 0.100

Table A.1: Lookup table for TRAST image acquisition settings.

A-2
Appendix B

Protocol for TRAST Contrast vs

Oxygen calibration

In this appendix, instructions, considerations and photographs for recreating

the parallel measurements with a fibre sensor and SPIM are presented. Results

from these measurements are used to construct the calibration curve for quan-

tifying oxygen concentrations (Figure 5.9, page 84).

Steps:

1. Prepare suitable FEP bags. The bags need to be able to hold approximately

800 µL and should be left open at the top to allow insertion of the probe (Figure

B.1a).

2. Calibrate your optical fibre sensor according to manufacturer specifications.

In this thesis a FireSting2 (Pyro–Science) sensor was used and was calibrated by

the two–point method: first, the air–saturation point was set by measurement in

water after bubbling atmosphere air for approximately 30 min using a common

aquarium pump. The zero–point was set by measurement in a water sample

with Na2 SO3 at final concentration of 2 M.


B-1
APPENDIX B. PROTOCOL FOR TRAST CONTRAST VS OXYGEN CALIBRATION

Figure B.1: (a) FEP bags that can hold approximately 800 µL. The tape is added for
structural support. A standard–sized 2–µL Eppendorf tube is shown for
comparison. (b) A trial run on bench top using the probe sensor. (c-d) Par-
allel measurements with an optical sensor and SPIM.

3. Perform a trial run on the bench using only the fibre sensor to make sure that

the titration experiment works as expected and oxygen concentrations are cap-

tured. Hold the bag with a reverse tweezer and place the sensor tip in the middle

of the bag. Pipette Ascorbic Acid and mix gently while measuring (Figure B.1b).

B Make sure the tip of the probe does not come in contact with the walls

of the bag at any time.

4. Transfer the set–up to the SPIM. Place the bag in the centre of the chamber

and hold it steadily using a reverse tweezer. Ensure that the bag is clean and

non–wrinkled to avoid distorting the light sheet. Lower the sensor tip near the

middle of the bag (Figure B.1c–d).

B-2
APPENDIX B. PROTOCOL FOR TRAST CONTRAST VS OXYGEN CALIBRATION

B Make sure that the probe tip does not lie in the laser beam path and

does not come in contact with the bag walls. The bag should not touch

objective or the walls of the chamber.

5. Measure oxygen concentration continuously with the point sensor while per-

forming the titration. At each point, capture two images, one taken while ex-

citing the sample with a low duty cycle pulse train (eg: 200,000 pulses, pulse

width = 0.5 µs, period = 50 µs, illumination time = 100 ms) and one with con-

tinuous wave illumination while keeping the total illumination time equal as

before. Calculate the relative decrease in fluorescence intensity between these

two images (using eq. 5.9, page 74) and correlate it with the absolute value of

oxygen concentration that was measured by the probe sensor at that time point.

B-3

You might also like