investigation of membrane potentials in bacterial biofilms

202
Investigation of membrane potentials in bacterial biofilms' communication and stress response A thesis submitted to The University of Manchester for the degree of Doctor of Philosophy in the Faculty of Science and Engineering 2020 Johanna A. Blee Department of Physics and Astronomy

Upload: others

Post on 04-Nov-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Investigation of membrane potentials in bacterial biofilms' communication and

stress response

A thesis submitted to The University of Manchester for the degree of

Doctor of Philosophy in the Faculty of Science and Engineering

2020

Johanna A. Blee

Department of Physics and Astronomy

2

Blank page

2

Table of contents

Table of contents 2

List of figures and tables 5

Abstract 1

Declaration 2

Copyright and ownership of intellectual property rights 3

Acknowledgements 5

Publications 6

1 Introduction 7

1.1 Bacterial biofilms 7

1.1.1 Overview 7

1.1.2 The extracellular polymeric substance 9

1.1.3 Biofilm lifecycle 10

1.1.4 Biofilm coordination and regulation 12

1.1.5 Biofilm tolerance and response to stress 13

1.2 P. aeruginosa and B. subtilis 14

1.3 Bacterial ion channels 19

1.4 Membrane potentials 20

1.5 Membrane potentials in biofilms 23

1.6 Outline 24

2 Background and methodology of experimental techniques 26

2.1 Fluorescence microscopy 26

2.1.1 Theory 26

2.1.2 Experimental methodology 30

2.2 Microbiological techniques 33

2.2.1 Background 33

2.2.2 Experimental methodology 36

2.3 Mathematical modelling of excitable systems 40

2.3.1 Theoretical background 40

2.3.2 Modelling methodology 46

3 Spatial propagation of electrical signals in circular biofilms 47

3.1 Overview 47

3.2 Introduction 47

3.3 Materials and methods 52

3

3.3.1 Cell culture and growth 52

3.3.2 Biofilm growth 53

3.3.3 Microscopy 55

3.3.4 Dyes 55

3.3.5 Data analysis 56

3.3.6 Modelling 57

3.4 Results 63

3.4.1 Electrical signalling in circular B. subtilis biofilms (experimental results and

characterisation) 63

3.4.2 Modelling the propagation of electrical signals in circular biofilms 68

3.5 Discussion 79

3.6 Conclusions 82

4 Membrane potentials, oxidative stress and the dispersal response of bacterial

biofilms to 405 nm light treatment 83

4.1 Overview 83

4.2 Introduction 84

4.3 Materials and Methods 85

4.3.1 Cell culture and growth 85

4.3.2 Cell preparation for microscopy 86

4.3.3 ROS scavengers 89

4.3.4 Microscopy 90

4.3.5 Data analysis 92

4.3.6 Mathematical modelling 94

4.4 Results 95

4.4.1 Physical response of P. aeruginosa cells to 405 nm light treatment 95

4.4.2 Membrane potential changes for P. aeruginosa in response to 405 nm light

stress 101

4.4.3 The response of fixed cells 111

4.4.4 Probing the dynamics and timescales of the hyperpolarisation response 112

4.4.5 Response in the presence of scavengers 114

4.4.6 The response of B. subtilis to 405 nm light 116

4.4.7 Hodgkin-Huxley model for the stress response 119

4.5 Discussion 125

4.6 Conclusions 129

5 Measuring c-di-GMP levels and the membrane potential response of Pseudomonas

aeruginosa exposed to oxidative stress 131

5.1 Overview 131

4

5.2 Introduction 131

5.3 Materials and methods 134

5.3.1 Cell culture and growth 134

5.3.2 The c-di-GMP reporter PA01 pCdrA::gfpc and the GFP control strain PA01:gfp 136

5.3.3 Transformation of pCdrA::gfpc into P. aeruginosa PA01 by electroporation 138

5.3.4 Plate reader assay 139

5.3.5 Cell preparation for single cell fluorescence microscopy 140

5.3.6 Microscope set-up 141

5.3.7 Fluorescent dyes and GFPs 141

5.3.8 405 nm light treatment and photobleaching using 488 nm light 141

5.3.9 Irradiance/dose measurements 142

5.3.10 Data analysis 142

5.4 Results 143

5.4.1 Confirming the suitability of P. aeruginosa PAO1 pCdrA::gfpc as a reporter of c-

di-GMP levels 143

5.4.2 Changes in c-di-GMP levels in response to 405 nm light 144

5.4.3 Changes in c-di-GMP levels in response to H202 146

5.4.4 Membrane potential response to H202 149

5.5 Discussion 150

5.6 Conclusions 154

6 Conclusions 156

7 Bibliography 161

Word count: 42, 710

5

1

List of figures and tables

Figure 1.1. Two pictures showing problematic biofilm growth. (a) Biofilm growth in a silicone

catheter, removed from patient after blockage4. (b) Microbial-induced corrosion in a

pipeline5. ........................................................................................................................... 8

Figure 1.2. Schematic diagram showing the five main stages of biofilm growth. Cells are

shown in red and EPS in yellow. (Stage I) Initial cell attachment: planktonic cells reversibly

attach to the surface, often by their poles. (Stage II) Irreversible attachment: cells attach to

the surface and begin to grow and divide colonising the surface. This transition is associated

with a loss of motility and an increase in the production of EPS. (Stage III) Aggregation: cells

continue to grow and divide forming cell clusters and aggregates. (Stage IV) Biofilm

formation: the cell density increases and cells begin to attach to surface cells which are

encased in an EPS. (Stage V) Mature biofilm formation: the biofilm has a complex three-

dimensional structure, with cells embedded in a complex EPS. Following biofilm maturation,

the biofilm disperses and cells return to the planktonic state, facilitating colonisation of new

surfaces. ......................................................................................................................... 10

Figure 1.3. Scanning electron micrograph of a B. subtilis biofilm on a chickpea root38.

Scanning electron micrograph of a P. aeruginosa biofilm on glass wool39. ....................... 15

Figure 1.4. Flowcharts of two biofilm regulation feedback loops. (a) Wsp feedback loop

involved in regulation of P. aeruginosa biofilm growth. (b) Feedback loop showing the

regulation of B. subtilis biofilm growth via Spo0a. ........................................................... 17

Figure 1.5 Illustration showing the distribution of potassium, sodium and chlorine ions

across a typical phospholipid cell membrane in a eukaryotic cell. .................................... 21

6

Figure 2.1. Fluorescent properties of a typical fluorophore. (a) Jablonski diagram showing

the electronic states of a fluorophore and its transitions from one to another energy level.

The thicker lines represent electronic energy levels, while the thinner lines denote the

various vibrational energy states (rotational energy states are ignored). (b) Spectral profile

of a fluorophore showing the Stokes shift observed between the excitation to emission

profiles. ........................................................................................................................... 27

Figure 2.2. Schematic diagram of the custom-built Olympus IX-71 inverted fluorescence

microscope. The laser beams were guided into the microscope by a combination of regular

and dichroic mirrors. The lasers were selectively filtered by a cube that contained a Semrock

Brightline full-multiband laser filter set. Fluorescence was detected using an ORCA-Flash4.0

LT PLUS Digital CMOS camera. ........................................................................................ 32

Figure 2.3. Step-by-step schematic showing the basic process used to culture bacterial cells.

Cells were streaked on to an agar plate, which was then incubated overnight. A single colony

from the plate was then picked from the plate and used to inoculate the culture which, after

further incubation, was used for further cell culture or to grow a biofilm, depending on the

experiment. ..................................................................................................................... 36

Figure 2.4. Schematic of the two different experimental set-ups used to grow biofilms. (a)

CellASIC ONIX microfluidic experimental set-up. (b) Syringe pump flow cell experimental set-

up. .................................................................................................................................. 39

Figure 2.5 Simple model of a cell membrane with a capacitor (𝐶𝑚) in parallel with a resistor.

....................................................................................................................................... 41

Figure 3.1. Proposed mechanism of active propagation of potassium through B. subtilis

biofilms83. The initial trigger for potassium release via Yug0 channel is metabolic stress, due

glutamate limitation. External potassium depolarizes neighbouring cells, limiting glutamate

7

uptake and thus produces further metabolic stress. This cycle results in the active

propagation of potassium through the biofilm. ............................................................... 49

Table 3.I. Recipes and sources for the culture media used in this chapter. ........................ 52

Figure 3.2. Illustrative figure showing how biofilms were grown in the CellASIC ONIX Y04D

plate (not to scale). (a) Schematic of a whole CellASIC ONIX Y04D plate, showing the four

identical, separate chambers, each with 6 inlet wells, a waste outlet well and a cell inlet

well. (b) Cell culture chamber, with six media inlets, waste outlet, cell inlet and six cell traps.

(c) Representative image of a circular B. subtilis biofilm grown overnight in a microfluidic

chamber. Biofilm cells were stained with the membrane potential dye ThT. .................... 54

Figure 3.3. Electrical wavefront from a B. subtilis biofilm. ThT fluorescence observed at 4 µm

from the centre of the biofilm as a function time. Signals from all angles are shown in blue

and the average signal is shown in red. ........................................................................... 57

Figure 3.4. Normalised cell density as a function of radial distance from the biofilm centre

for our experimental centrifugal wavefront data (red), centripetal wavefront data (black)

and agent-based fire-diffuse-fire model (blue). The centripetal biofilm had a larger radius

(~150 m) than the centrifugal biofilm (~90 m). ............................................................ 62

Figure 3.5. Electrical signal propagation through a two-dimensional biofilm. Schematics

show the spread of (a) centrifugal (‘away from the centre’) and (b) centripetal (‘towards the

centre’) electrical wave fronts through a biofilm. (c) The electrical signal given by ThT

fluorescence as a function of time at five different biofilm radii (r = 2 µm, 10 µm, 15 µm, 100

µm and 150 µm) from fluorescence microscopy experiments. .......................................... 63

Figure 3.6. Propagation of centrifugal and centripetal electrical signals through B. subtilis

biofilms. (a) and (b) ThT fluorescence intensity as a function of time and radial distance for

8

a biofilm in which an electrical signal has originated from (a) the biofilm centre (centrifugal)

and (b) the biofilm edge (centripetal). (c) The signals’ fluorescence energy density as a

function of radial distance for the centrifugal wavefront (red) shown in (a) and for the

centripetal wavefront (black) shown in (b), fitted with sigmoids (Equation 3.5). (d) Radial

distance for the maximum intensity as a function of signal mean time for the centrifugal

wavefront (red) shown in (a) and the centripetal wavefront (black) shown in (b), fitted with

power laws (Equation 3.7). .............................................................................................. 66

Figure 3.7. Fire-diffuse-fire model of electrical signal propagation through a B. subtilis

biofilm (Equation 3.8). (a) A plot of 𝑔(𝜈) as a function of 𝜈 for a range of different potassium

decay rates (𝛾). 𝑔(𝜈) is a function which may be used to determine the model’s stability and

thus find constantly propagating solutions to the FDF model (Equation 3.10). (b) The

potassium signal produced by our FDF model of a biofilm (Equation 3.9). (c) The signal

amplitude of the potassium wave shown in (b). (d) The velocity profile (position of the signal

maximum as a function of time) of the signal shown in (b). ............................................. 70

Figure 3.8. Workflow showing the steps executed by our model per time step (Δt). Firstly,

CellSignal was used to update diffusing signalling molecules. Secondly, the cell states were

updated for each cell in the simulation. Finally, CellEngine was used to grow the whole

colony. ............................................................................................................................ 73

Figure 3.9. Snapshots from a Gro simulation of our agent-based fire-diffuse-fire model of a

two-dimensional circular B. subtilis biofilms shown in (a) three-dimensions and (b) two-

dimensions. Snapshots are shown for time since initial firing at the centre of the biofilm T=0,

17, 63 and 120 mins. (c) A magnified image of a potassium wave spreading out from the

centre of the biofilm simulated by our agent-based fire-diffuse-fire model where the

bacterial agents are clearly visible. .................................................................................. 74

9

Figure 3.10. Propagation of (a) centripetal and (b) centrifugal electrical waves produced by

our agent-based fire-diffuse-fire model. The potassium profiles were produced by our model

for a signal triggered at (a) the biofilm centre and (b) the biofilm edge. (c) Fluorescence

energy density, as a function of radial distance, of the centripetal signal (red) and of the

centrifugal signal (black) fitted with sigmoids (Equation 3.5). (d) Radial distance for the

maximum intensity as a function of the signal mean time for the centripetal signal (red)

shown in (a) and the centrifugal signal (black) shown in (b), fitted with power laws (Equation

3.7). For (c) and (d) data was averaged over three separate simulations.......................... 76

Figure 3.11. Kurtosis and skewness of the electrical signal as a function of radial distance.

(a) Kurtosis of our experimental centrifugal wavefront (red) and centripetal wavefront

(black). (b) Skewness of our experimental centrifugal wavefront (red) and centripetal

wavefront (black). (c) Kurtosis of our ABFDF model’s centrifugal wavefront (red) and

centripetal wavefront (black). (d) Skewness of our ABFDF model’s centrifugal wavefront

(red) and centripetal wavefront (black). .......................................................................... 77

Figure 4.1. Schematic showing the ibidi flow cells in which biofilms were grown. (a) Ibidi µ-

slide VI0.4 with six identical channels in which P. aeruginosa biofilms were grown. (b) Ibidi µ-

slide III perfusion flow cell slides with three identical channels in which B. subtilis biofilms

were grown. .................................................................................................................... 87

Figure 4.2. Schematic showing a top and side view of the agarose microscope slide set-up

for fluorescence microscopy. Bacteria were immobilised between the agarose medium and

the microscope coverslip. ................................................................................................ 89

Figure 4.3. Growth curves (OD600) for P. aeruginosa grown in TSB media with and without

10 µM ThT. ...................................................................................................................... 91

10

Figure 4.4. Representative images that depict P. aeruginosa cells stained with ThT at the

five stages of biofilm growth. (a) – (e) show representative cells at Stage I through to Stage

V. .................................................................................................................................... 94

Figure 4.5. Phase contrast images depict the transformational change seen in P. aeruginosa

cells at Stage III of biofilm growth before and after a dose of 1.8 J/ cm2 of 405 nm light. . 96

Figure 4.6. Dispersal response of Stage I P. aeruginosa cells to treatment by 0.1 J/cm2 of 405

nm and 488 nm light. (a) Representative brightfield images show the number of cells before

and after treatment. (b) Graph showing the number of cells before and after treatment. 97

Figure 4.7. (a) Biofilm residence probability as a function of time (or equivalently dose) of P.

aeruginosa biofilms, exposed to 120 ± 4 µW/cm2 405 nm light, for the five stages of biofilm

growth. Corresponding fits of the Kaplan-Meier estimator (𝑆(𝑡), equation (4.3)) shown as

pink dashed lines. (b) The hazard functions ℎ(𝑡) obtained from the Kaplan-Meier functions

(𝑆(𝑡)) shown in (a) at Stage I, II and IV of P. aeruginosa biofilm growth with corresponding

fits shown in red. (c) The cumulative hazard functions 𝐻(𝑡) obtained from the Kaplan-Meier

functions (𝑆(𝑡)) shown in (c) at Stage I, II and IV of P. aeruginosa biofilm growth with

corresponding fits shown in red. (d) The ratio between hazard constants 𝑎𝑡 and 𝑏𝑡 (equation

(4.9)) from hazard functions shown in (b) and (c) at Stages I and Stages II and IV of P.

aeruginosa biofilm growth i.e. growth phases with a significant dispersal of bacteria.

Averages were taken from at least 20 cells in the field of view. ........................................ 99

Figure 4.8. Average ThT fluorescence of Stage II P. aeruginosa cells irradiated with 120 ± 4

µW/cm2 405 nm light as a function of time (or equivalently dose).................................. 102

Figure 4.9. Average ThT intensity of Stage I P. aeruginosa cells as a function of time (or

equivalently dose) observed in response to 405 nm light and 488 nm light, at a constant

irradiance of 480 ± 6 µW/cm2. ....................................................................................... 103

11

Figure 4.10. Average DiSC3(5) intensity of Stage I (a) P. aeruginosa and (b) B. subtilis cells as

a function of time (or equivalently dose) observed in response to 200 ± 4 µW/cm2 405 nm

light (black) and in response to no treatment (red). ....................................................... 104

Figure 4.11. (a) Average cell ThT intensity as a function of time (or equivalently dose)

observed in response to 405 nm light at different stages of P. aeruginosa biofilm growth, in

the same media, at a constant irradiance of 120 ± 4 µW/cm2 with corresponding sigmoidal

fits to equation (4.11). (b) Average ThT fluorescence of mature P. aeruginosa biofilm cells as

a function of time (or equivalently dose) in response to 405 nm light. Data was collected for

a much longer time than that shown in (a) i.e. 900 mins compared to 1500 seconds. .... 108

Figure 4.12. Boltzmann sigmoidal fit parameters (half-maximal dose (D0) and slope constant

𝑥 as given by equation (4.11)) which define the average hyperpolarisation of P. aeruginosa

cells at the five stages of biofilm growth in response to 405 nm light at 120 ± 4 µW/cm2 of

405 nm light. ................................................................................................................. 109

Figure 4.13. (a) Individual cell ThT intensity as a function of time (or equivalently dose),

observed at Stage I of P. aeruginosa biofilm growth, in response to 120 ± 4 µW/cm2 405 nm

light. (b) Leaving time of individual cells from (a) as a function of half-maximal time, with

corresponding linear fit shown in red. (c) The average difference in the leaving time of cells

as a function of cell separation. (d) The average difference in the half-maximal time of cells

as a function of cell separation. ..................................................................................... 110

Figure 4.14. Average ThT fluorescence of trapped P. aeruginosa cells as a function of time

(or equivalently dose) in response to 405 nm light at a constant irradiance of 120 ± 4

µW/cm2. ....................................................................................................................... 112

Figure 4.15. Average ThT fluorescence of trapped P. aeruginosa cells as a function of dose

in response to 405 nm light at an irradiance of 120 ± 4 µW/cm2. The three different curves

12

represent the laser on constantly (black), the laser switched off for 1 min after 15 sec of

illumination (blue) followed by continuous irradiation, and the laser switched off for 1 min

after 1 min on illumination (red) followed by continuous irradiation. ............................. 113

Figure 4.16. Average ThT fluorescence of trapped P. aeruginosa cells as a function of (a)

time and (b) dose in response to 405 nm light at an irradiance of 120 ± 4 µW/cm2. The black

curves represent the laser being on constantly and the red curves represent the response

observed when the laser is turned on once every 0.1 min for 10 ms. .............................. 114

Figure 4.17. (a) Average cell ThT intensity as a function of time (or equivalently dose) of

Stage I P. aeruginosa cells with and without added scavengers (100 mM sodium pyruvate

and 200 U/ml catalase) exposed to 405 nm light at a constant irradiance of 120 ± 4 µW/cm2

with corresponding sigmoidal fits to equation (4.11) shown in blue. (b) Residence probability

(probability of surface cells remaining) of Stage I P. aeruginosa cells following 700 secs

(equivalent to 85 mJ/cm2) of 405 nm light treatment with and without added scavengers

(100 mM sodium pyruvate and 200 U/ml catalase). Errors bars show the standard error.

..................................................................................................................................... 115

Figure 4.18. Biofilm residence probability as a function of time (or equivalently dose) for a

B. subtilis biofilm exposed to 120 ± 4 µW/cm2 of 405 nm light, the Kaplan-Meier estimate

(equation (4.3)) is shown in red, with an inset of the corresponding cumulative hazard

function (equation (4.8)). Averages were taken from at least 20 cells in the field of view.

..................................................................................................................................... 117

Figure 4.19. Average ThT intensity of Stage II B. subtilis cells as a function of time (or

equivalently dose) in response to 120 ± 4 µW/cm2 of 405 nm light. ................................ 118

Figure 4.20. Average cell ThT intensity as a function of time (or equivalently dose) of Stage

I B. subtilis cells with and without added scavengers (100 mM sodium pyruvate and 200

13

U/ml catalase) exposed to 405 nm light at a constant irradiance of 120 ± 4 µW/cm2 with

corresponding sigmoidal fits to equation (4.11) shown in blue. ...................................... 119

Figure 4.21. ThT fluorescence as a function of (a) time and (b) dose in response to 405 nm

light at an irradiance of 120 ± 4 µW/cm2 produced by our Hodgkin-Huxley style model. The

black curves represent the laser being on constantly and the red curves represent the

response observed when the laser is turned on for 10 ms once every 0.1 min. ............... 122

Figure 4.22. (a) Three-dimensional hyperpolarisation curve shows the membrane potential

response to 405 nm light of biofilm cells predicted by our model for a range of input

irradiances (85 µW/cm2 - 575 µW/cm2) as a function of time. (b) ThT fluorescence as a

function of dose in response to 405 nm light at an irradiance of 85 µW/cm2 (black), 185

µW/cm2 (green), 285 µW/cm2 (red) and 385 µW/cm2 (blue) produced by our model. (c) ThT

fluorescence as a function of dose produced by our model in response to 405 nm light at an

irradiance of 120 µW/cm2 which is switched on and off. The black curve represents the dose

response simulated when the laser is on constantly, the blue curve represents the response

observed when the laser is switched on for 45 sec then off for 1 min then back on, and the

red curve represents the response observed when the laser is switch on for 1 min then off

for 1 min then back on. (d), (e) and (f) Hyperpolarisation curves show the membrane

potential with time (or equivalently dose) in response to 120 µW/cm2 of 405 nm light

predicted by our Hodgkin-Huxley style model with corresponding sigmoidal fits to equation

(4.11). (d) Shows simulations of the five stages of biofilm growth based on the assumption

that the rate of ROS production and decay were dependent on the metabolic state of cells

and the stage of biofilm growth. (e) and (f) show simulations of our original and adapted

model in the presence (black) and absence (red) of ROS scavengers, based on the

assumption that the addition of scavengers leads to an increase in the decay rate of ROS.

..................................................................................................................................... 123

14

Figure 5.1. Physiological functions of the intracellular secondary messenger c-di-GMP. C-di-

GMP is synthesised from 2 GTPs via diguanylate cyclases and is degraded into pGpG/GMP

via phosphodiesterases. Extracellular signals control the activity of these proteins and

therefore ultimately regulate the levels of intracellular c-di-GMP. Low levels of c-di-GMP are

associated with the promotion of planktonic behaviour (e.g. motility and acute virulence),

whereas high levels of c-di-GMP are associated with biofilm growth. ............................ 132

Table 5.I. Descriptions and details for the bacterial strains used in this chapter. ............ 135

Table 5.II. Recipes and sources for the culture media used in this chapter. ..................... 135

Figure 5.2. Sequence map of pCdrA::gfpc. (a) Addgene full sequence map for pCdrA::gfpc

created with SnapGene. Shown on the map are: unique 6+ cutters, primers, features and

translations. (b) Schematic showing the horizontal cassette map for pCdrA::gfpc showing

the cdrA promoter fused with the artificial optimized ribosomal binding site (RBSII). The

transcriptional fusion is followed by two transcriptional terminators (T0 and T1). ......... 137

Figure 5.3. Treatment of P. aeruginosa PAO1 pCdrA::gfpc with SNP at concentration of 0

µM, 62.5 µM and 125 µM. (a) Growth measurements given by the OD450. (b) Fluorescence

GFP measurements. ...................................................................................................... 144

Figure 5.4. Treatment of P. aeruginosa PAO1 pCdrA::gfpc and P. aeruginosa PAO1::gfp with

405 nm light. (a) Normalised GFP fluorescence per cell before, following and an hour after

treatment of cells with 3.6 mJ/cm2 of 405 nm light. (b) Photobleaching curves for P.

aeruginosa PAO1 pCdrA::gfpc exposed to 488 nm light at an irradiance of 120 ± 2 µW/cm2

(black) before and (red) after treatment with 3.6 mJ/cm2 of 405 nm light, with exponential

fits given by Equation 5.3. ............................................................................................. 146

15

Figure 5.5. Treatment of P. aeruginosa with 1 mM H202. (a) Ratios of the average GFP

fluorescence to OD600 as a function of time since inoculation for: P. aeruginosa PAO1

pCdrA::gfpc (▲), P. aeruginosa PAO1::gfp (■) and P. aeruginosa PAO1 (●), without H202

(black) and with 1 mM H202 (red). (b) Growth curves of: P. aeruginosa PAO1 pCdrA::gfpc

(▲), P. aeruginosa PAO1::gfp (■) and P. aeruginosa PAO1 (●), without H202 (black) and with

1 mM H202 (red). (c) Ratio of average GFP fluorescence to OD600 , at mid-exponential growth

phase (OD600 ≈ 0.5), without H202 (blue) and with 1 mM H202 (green) for: P. aeruginosa PAO1

pCdrA::gfpc (▲), P. aeruginosa PAO1::gfp (■) and P. aeruginosa PAO1 (●). The presented

errors are standard errors. ............................................................................................ 147

Figure 5.6. Treatment of P. aeruginosa PA01 pCdrA::gfpc and P. aeruginosa PA01::gfp with

H202. (a) Average cell GFP fluorescence of P. aeruginosa PA01 pCdrA::gfpc and P. aeruginosa

PA01::gfp, for a range of H202 concentrations (1 µM, 10 µM, 1 mM and 10 mM), normalised

to the average cell GFP fluorescence without H202. (b) The normalised average cell decay

time constant of photobleaching by 120 ± 2 µW/cm2 488 nm light (Equation 5.3) as a

function of H202 concentration. The inset shows the decrease in cell fluorescence due to

addition of 10 mM H202 in fluorescence microscopy images. The presented errors are

standard errors. ............................................................................................................ 149

Figure 5.7. Membrane potential response of P. aeruginosa treated with H202. (a) Change in

average ThT fluorescence per cell caused by addition of 10 mM H202. (b) Average normalised

ThT (membrane potential) dose response without H202 (black) and with 10 mM H202 (red) in

response to 120 µW/cm2 405 nm light. The presented errors are standard errors. ......... 150

1

Abstract

Bacterial biofilms pose a large threat to health. To understand this resilient and

coordinated form of bacterial growth in more detail the bacterial cells’ membrane

potentials were studied. In circular Bacillus subtilis biofilms, in addition to previously

described electrophysiological waves, which travelled from the centre of the biofilm out to

the edge (centrifugal), waves which travelled from the edge of the biofilms towards the

centre (centripetal) were also observed. New data analysis techniques and an agent-based

fire-diffuse-fire model were used to show that the spatial heterogeneity in bacterial cell

placements and curvature affected the propagation of wavefronts through the biofilm.

The membrane potentials and physical responses of Pseudomonas

aeruginosa and B. subtilis biofilms to 405 nm light were also investigated. It was found that

all cells exhibited membrane potential hyperpolarisations in response to 405 nm light. The

dynamics of these membrane potential changes depended on the stage of biofilm growth.

At the early stages of biofilm growth, cells also dispersed in response to 405 nm light. A

Hodgkin-Huxley style model was used to demonstrate that changes observed during biofilm

growth could explain the observed differences in membrane potential dynamics.

The secondary messenger cyclic di-guanosine monophosphate (c-di-GMP) is a

crucial regulator in biofilm growth in P. aeruginosa. Its role in regulating the oxidative stress

response of P. aeruginosa and the connection between c-di-GMP levels and membrane

potential were investigated using a fluorescence-based GFP reporter strain. Oxidative

stress induced changes in GFP and therefore the GFP-based reporter could not be reliably

used to measure the c-di-GMP levels at high levels of oxidative stress. At low levels of

oxidative stress, the reporter strain was used to show that oxidative stress induced an

increase in the levels of c-di-GMP. This indicates that P. aeruginosa does regulate oxidative

stress via this intracellular messenger and provides a mechanism that drives the dispersal

response of P. aeruginosa to 405 nm light.

Overall, it was shown that bacteria regulate their membrane potentials in response

to a range of different stresses. The data analysis and modelling techniques developed in

this thesis can be used to further study this emerging field of bacterial electrophysiology.

2

Declaration

No portion of the work referred to in the thesis has been submitted in support of

an application for another degree or qualification of this or any other university or other

institute of learning.

3

Copyright and ownership of intellectual property rights

i. The author of this thesis (including any appendices and/or schedules to this

thesis) owns certain copyright or related rights in it (the “Copyright”) and s/he

has given The University of Manchester certain rights to use such Copyright,

including for administrative purposes.

ii. Copies of this thesis, either in full or in extracts and whether in hard or

electronic copy, may be made only in accordance with the Copyright, Designs

and Patents Act 1988 (as amended) and regulations issued under it or, where

appropriate, in accordance with licensing agreements which the University has

from time to time. This page must form part of any such copies made.

iii. The ownership of certain Copyright, patents, designs, trademarks and other

intellectual property (the “Intellectual Property”) and any reproductions of

copyright works in the thesis, for example graphs and tables (“Reproductions”),

which may be described in this thesis, may not be owned by the author and

may be owned by third parties. Such Intellectual Property and Reproductions

cannot and must not be made available for use without the prior written

permission of the owner(s) of the relevant Intellectual Property and/or

Reproductions.

iv. Further information on the conditions under which disclosure, publication and

commercialisation of this thesis, the Copyright and any Intellectual Property

and/or Reproductions described in it may take place is available in the

University IP Policy (see

4

http://documents.manchester.ac.uk/DocuInfo.aspx?DocID=2442 0), in any

relevant Thesis restriction declarations deposited in the University Library, The

University Library’s regulations (see

http://www.library.manchester.ac.uk/about/regulations/) and in The

University’s policy on Presentation of Theses.

5

Acknowledgements

I would like to acknowledge the Engineering and Physical Sciences Research Council

for funding this PhD and making it possible.

I would like to extend my deepest thanks to both my Supervisors, Thomas Waigh

and Ian Roberts, for the time, guidance and wisdom they have provided me with during this

PhD.

I am extremely grateful to Marie Goldrick for sharing her seemingly infinite

microbiological knowledge. I would like to thank all my PhD friends who have given me the

knowledge, coffee, cake and banter to motivate me to the end of this PhD. I would

especially like to thank Hannah Perkins; you really were a large part of making this PhD

enjoyable.

I would finally like to thank my family. I would specifically like to thank my parents,

not just for their continual support and guidance, but for the curiosity and resilience they

instilled in me. Jamie I would like to thank you for, as always, keeping me grounded and

giving me perspective.

6

Publications

Blee, J. A., Roberts, I. S. & Waigh, T. A. Spatial propagation of electrical signals in circular

biofilms: A combined experimental and agent-based fire-diffuse-fire study. Phys. Rev. E 100,

52401 (2019).1

Blee, J. A., Roberts, I. S. & Waigh, T. A. Membrane potentials, oxidative stress and the

dispersal response of bacterial biofilms to 405 nm light. In press. Phys. Biol. 17, 036001

(2020).2

7

CHAPTER

ONE

1 Introduction

An overview will be given of the current understanding of bacterial biofilms with an

emphasis on the importance of regulation, communication and membrane potentials.

Having provided a contextual overview of the project the final section of this chapter will

be used to outline the thesis content.

1.1 Bacterial biofilms

1.1.1 Overview

Bacterial biofilms are communities of bacteria encased in a self-produced

extracellular polymeric substance (EPS). Bacterial biofilms are coordinated multicellular

communities3–5. This prevalent form of growth supports bacterial survival in a wide range

of settings, from the human body to oil pipelines, resulting in a whole host of problems

which are difficult to tackle (Figure 1.1).

8

Figure 1.1. Two pictures showing problematic biofilm growth. (a) Biofilm growth in a silicone

catheter, removed from patient after blockage6. (b) Microbial-induced corrosion in a

pipeline7.

The National Institute of Health estimates that approximately 65 % of all microbial

infections and 80 % of chronic infections are associated with biofilms8. Bacterial biofilms

require antibiotics and biocides at levels 500 to 5000 times higher than planktonic cells9.

This has led to a critical need for new techniques to tackle biofilm growth. Biofilms are also

costly to industry through processes such as biofouling10, adding an extra economic

motivation for combating this form of growth.

Not all biofilms are harmful and they play important, positive roles in the stability

of ecological systems. They also have beneficial applications11, for example, in the

production of industrial chemicals and in treating wastewater. In addition to current

applications, it is hoped they may have further uses, e.g. in the production of biopower12.

The devastating effects of biofilms, paired with the important applications, provide

strong motivation to study this mode of growth. This has been reflected in the large

increase in the study of biofilms in recent decades. It is hoped that a greater understanding

may enable the development of new strategies with which to combat destructive biofilms,

as well as allowing useful biofilms to be harnessed.

9

1.1.2 The extracellular polymeric substance

Central to biofilm growth is the biofilm’s self-produced EPS. Strains of bacteria that

are incapable of producing an EPS (and consequently a biofilm), form homogenous

colonies13. In comparison, bacterial biofilms usually form complex, mushroom-like

structures; with channels which distribute nutrients and remove waste14.

The EPS makes up 75-90 % of the biofilm: giving it structure, contributing to its

genetic regulation and controlling the flow of nutrients and toxins through it4,15. By

impairing the penetration of antibiotic agents and toxic substances the EPS acts as a

diffusion barrier protecting the bacteria it contains. Specific extracellular components, such

as enzymes, have also been shown to actively trap and damage penetrating antimicrobial

agents16,17. One method, employed by several species of bacteria to achieve a robust

biofilm structure, is to synthesise protein fibres that form a shell onto which the microbial

cells and other EPS compounds may attach18.

The EPS produced by the majority of bacteria consists primarily of polysaccharides

combined with eDNA (mitochondrial or nuclear DNA that has been released by the bacteria

into the environment), lipids and extracellular proteins. As well as common constituents,

several novel EPS components have emerged, such as the bacterial hydrophobin, BslA,

which forms a water-resistant ‘raincoat’ over the B. subtilis biofilm19. The exact composition

of the EPS depends on the microbial species and the environment e.g. presence of flow.

This in turn affects key properties of the biofilm, from its mechanical stability to its

hydrophobicity.

Many EPS components remain unquantified and the roles of those that have been

quantified often remain elusive. These components define many of the biofilm’s key

properties and behaviours. Therefore, expanding our understanding of the molecular

function of biofilm components is crucial in developing our knowledge of this mode of

growth.

10

1.1.3 Biofilm lifecycle

Figure 1.2. Schematic diagram showing the five main stages of biofilm growth. Cells are

shown in red and EPS in yellow. (Stage I) Initial cell attachment: planktonic cells reversibly

attach to the surface, often by their poles. (Stage II) Irreversible attachment: cells attach to

the surface and begin to grow and divide colonising the surface. This transition is associated

with a loss of motility and an increase in the production of EPS. (Stage III) Aggregation: cells

continue to grow and divide forming cell clusters and aggregates. (Stage IV) Biofilm

formation: the cell density increases and cells begin to attach to surface cells which are

encased in an EPS. (Stage V) Mature biofilm formation: the biofilm has a complex three-

dimensional structure, with cells embedded in a complex EPS. Following biofilm maturation,

the biofilm disperses and cells return to the planktonic state, facilitating colonisation of new

surfaces.

The details of biofilm growth vary depending on the bacteria, as well as the

environment, but the lifecycle can be broadly described using five stages of growth (Figure

11

1.2)11,20,21. The first step involves the reversible attachment of the planktonic and in some

cases motile cells to a surface. Secondly, these cells attach irreversibly to the surface. This

transition is associated with a loss of motility and an increase in the production of EPS. The

initial and long term attachment of a bacterium to a surface is dependent on the ratio of

attractive and repulsive forces22. These forces can vary significantly for different bacteria

and surfaces; e.g. surfaces with varying compositions and topographies. Attachment is

mediated by forces which are dependent on surrounding conditions, such as available

nutrients and the presence of flow. The interaction between a bacterium and a surface may

also be affected by intermediary factors, such as adhesins23. These not only ensure

attachment and adhesion in the presence of shear stress, but also allow surface specificity.

Some of the forces involved in bacterial and biofilm adhesion remain elusive.

Following attachment the bacterial cells grow and divide to form cell clusters and

aggregates. Additional cells attach to these surface cells. The biofilm continues to grow until

a mature biofilm, with a complex three-dimensional structure emerges. By this stage, the

cells in the biofilm are embedded in a complex EPS. Following biofilm maturation, the

biofilm disperses and cells return to the planktonic state, facilitating colonisation of new

surfaces. Dispersal, which plays a significant role in the spread of bacteria to new

environments, is the least well understood part of the biofilm lifecycle. Dispersion is

triggered and controlled by environmental cues and inter/intracellular signals. To date no

ubiquitous mechanisms of dispersal have been observed and so they remain difficult to

characterise24. In summary, the growth of a biofilm is controlled by a host of physical,

chemical and biological mechanisms, which ultimately determine the structure and

behaviour of the resulting biofilm.

12

1.1.4 Biofilm coordination and regulation

The entire process of biofilm growth, from initial attachment to dispersal, is tightly

regulated21. Biofilm growth relies on the coordination of behaviour between its constituent

bacteria; this is achieved via a complex network of signalling molecules and genetic cues25.

Environmental triggers and secreted quorum sensing molecules both play roles in

regulating the genetic transition of cells from the planktonic to the sessile state and back

again. This is then reinforced by positive feedback loops in which genetic changes cause

cells to produce further influencing factors, such as signalling molecules and enzymes.

One method by which bacteria are known to regulate their cooperative behaviour

is quorum sensing, where communication is achieved via signalling molecules. Quorum

sensing coordinates behaviour based on the local density of the bacterial population.

During this process signalling molecules bind to a receptor on the bacteria, triggering the

transcription of specific genes. A lot of research has focused around identifying different

signalling molecules and the pathways that lead from their production to the alteration of

gene expression. Quorum sensing was initially discovered as a method for regulating

bioluminescence, but has since been identified in association with a range of behaviours.

Some signalling molecules are specific to one microbial species, while others can regulate

communication between a diverse range of species. Several examples of eukaryotes

developing mechanisms to counteract quorum sensing have also been reported26. For

example, one study found that epithelial cells quench the activity of the P. aeruginosa

3OC12-homoserine lactone autoinducer27. It is hoped that methods employed by hosts to

interrupt microbial communication may be mimicked and adapted to target biofilms.

Variations between different biofilms do not only depend on the microbial species,

but also on the environmental growth conditions. Universal mechanisms, for example of

communication, offer particularly attractive targets from which to develop treatments, as

13

they may offer a more widespread solution, where only minor adaptations need to be made

depending on the type of biofilm.

1.1.5 Biofilm tolerance and response to stress

The prevalence and increased resistance of biofilm cells to environmental stress

stems from increased adaptation abilities, as well as due to increased protection offered by

the EPS and neighbouring cells. Mature biofilms are built from heterogeneous,

phenotypically distinct sub-populations, each of which fulfil distinct roles28. Individual cells

cooperate and compete within a complex framework, that is often likened to our own

multicultural cities. Specific phenotypes are associated with distinct locations and growth

stages, indicating spatio-temporal regulation28. Different phenotypes are expressed in

response to varying conditions (e.g. pH, O2 and nutrients) across the biofilm as well as

stochastic gene expression.

The heterogeneity of bacteria ensures survival in variable environments in a bet

hedging strategy. One example of specific cells which survive stressful conditions are

persister cells, which are found deep within biofilms29. They have a decreased metabolic

activity and so have a higher antibiotic tolerance as antibiotics predominantly target cell

growth. They also show higher tolerance to oxidative stress, for example, the viability of

stationary phase cultures of B. subtilis is not affected by treatment with 10 mM H2O2,

whereas the viability of exponential phase cells is reduced to approximately 0.01 %30.

Another protective genetic method used by bacteria is horizontal gene transfer. Transfer of

advantageous genes is key to the evolution of bacteria and their resistance to antibiotics.

The efficiency of this process is enhanced within biofilms31.

The resistance of bacteria to environmental stress is strongly dependant on the

mode of growth (planktonic vs biofilm). During planktonic growth individual cell

14

characteristics are crucial, whereas during biofilm growth, the influence of external

protective elements and surrounding cells is often more important. A combination of

different mechanisms are employed by bacteria and by biofilms to ensure survival and

adaptation to environmental stresses. This is demonstrated by the response of cells to

photooxidative stress. Individual characteristics, such as pigmentation, are key for

photoprotection at a range of wavelengths32,33, while external factors and surrounding cells

can influence both photoprotection and the magnitude of the evoked response to

photooxidative stress. Biofilm cells are more resistant to photoinactivation by light over a

range of wavelengths. It is expected that a range of factors contribute to this increased

resistance. Firstly, some biofilm cells enter a dormant metabolic state, which is known to

be associated with decreased ROS production. Higher levels of catalase, which significantly

protect against oxidative stress, have also been detected in biofilm cells34. Finally, biofilm

cells are encased in a complex EPS, which is known to play a major role in protection. Biofilm

matrix components, such as alginate, protect biofilms by shielding them from light35, while

other components, such as cellulose and alginate, also protect against reactive oxygen

species generated under stress36.

1.2 P. aeruginosa and B. subtilis

Two model bacterial species, P. aeruginosa and B. subtilis, were used in this project

(Figure 1.1). P. aeruginosa is a Gram-negative bacterium, which causes difficult to treat,

nosocomial infections, in particular, topical skin infections and chronic lung infections in

cystic fibrosis patients37,38. In addition, P. aeruginosa can cause biofouling on nano-filtration

devices involved in seawater desalination systems39.

15

B. subtilis is a Gram-positive, spore forming bacterium, that is ubiquitous in the

environment20. It has been studied in the laboratory for over a century, leading to the

domestication of commonly used strains, such as 168. These strains show a distinct

attenuation in their ability to form biofilms when compared to wild type strains, such as

NCIB361013. Domestication can introduce mutations which impair the bacteria’s ability to

swarm on surfaces and form robust structures13.

Figure 1.3. Scanning electron micrograph of a B. subtilis biofilm on a chickpea root40.

Scanning electron micrograph of a P. aeruginosa biofilm on glass wool41.

The regulation of biofilm formation is a complex process, which is still not fully

understood. It is affected by a large number of regulatory pathways and feedback loops. By

examining mutations in strains that are incapable of biofilm formation and by introducing

mutations into wild type strains, knowledge may be gained regarding the genetic regulation

of biofilm growth. One of the key bacterial secondary messengers in biofilm regulation is

c-di-GMP. Its role in regulating the transition from the planktonic to the sessile state and

back again (dispersal) has been extensively studied, especially in model organisms such as

P. aeruginosa42–44. C-di-GMP regulates a host of biofilm associated behaviour from flagella

rotation to exopolysaccharide production, surface adhesin expression and antimicrobial

resistance44–46. The versatility and adaptation capabilities of P. aeruginosa are linked with a

16

large array of complex regulatory networks, including a broad range of genes involved in c-

di-GMP production and degradation. Diguanylate cyclases (DCGs) and phosphodiesterases

(PDEs) are responsible for the biosynthesis and the degradation of c-di-GMP, respectively.

The catalytic domains of DGCs carry a GGDEF site and PDEs carry either an EAL or HD-GYP

domain. These domains are often seen in conjunction with a receiver or transmission

domain, indicating modulation of their activity by external/internal stimuli43. In some

proteins, both GGDEF and EAL domains are present, suggesting a dual function as a DGC

and a PDE. For example, in planktonic P. aeruginosa cells, MucR functions as a DGC,

whereas in biofilm cells, it acts as a PDE47. The P. aeruginosa genome encodes a large

number of DGCs and PDEs, which are modulated by a broad range of signals. In turn, these

proteins regulate a wide range of behaviours. An example of one of these proteins which

regulates and is regulated by biofilm growth is the DGC WspR (Figure 1.4(a)). When a

surface is sensed, the Wsp signal transduction complex phosphorylates WspR and triggers

c-di-GMP synthesis48. In turn, WspR phosphorylation triggers subcellular WspR

oligomerization and cluster formation, increasing the DGC activity49. C-di-GMP then binds

to the l-site inhibiting WspR activity50.

B. subtilis biofilm regulation is primarily dependant on the phosphorylation state of

Spo0a, which is controlled by multiple histidine kinases51. This multicomponent

phosphorelay is affected by a range of stimuli, such as osmotic pressure and potassium

leakage. Spo0a-P produces sinl, which controls the ratio of two transcriptional factors sinR

and slrR. SinR directly represses exopolysaccharide production and promotes flagellar

motility; while SlrR activates biofilm genes and represses motility. One example of an

operon involved in and complexly affected by biofilm formation in B. subtilis is mstX-Yug052.

Expression of mstX and the downstream potassium channel Yug0 is required for biofilm

development and overexpression of mstX may induce biofilm formation. Phosphorylation

of Spo0a is achieved through the histidine kinase KinC, which is activated by potassium

17

efflux through Yug0. SinR negatively regulates the mstX-Yug0 operon and so represses it in

the planktonic state (Figure 1.4(b)).

Figure 1.4. Flowcharts of two biofilm regulation feedback loops. (a) Wsp feedback loop

involved in regulation of P. aeruginosa biofilm growth. (b) Feedback loop showing the

regulation of B. subtilis biofilm growth via Spo0a.

18

Once biofilm formation has been initiated the biofilm’s structure and the

composition of its EPS depends on the genetic expression of cells within it, which in turn

depends on the strain and on the environmental conditions53. The EPS of B. subtilis biofilms

grown in sucrose-rich media, e.g. SYM, is distinctly different from the EPS when grown in

reduced media, e.g. MSgg54. The EPS of biofilms grown in sucrose-rich media is dominated

by the polysaccharide levan, while the EPS from biofilms grown in sucrose-poor media also

contains a significant amount of proteins, DNA and polysaccharides. There are also

differences observed between different experimental setups, for example, the presence of

flow can reduce the thickness of the observed biofilm55.

The main component of the B. subtilis EPS is usually an exopolysaccharide with a

large molecular weight, which is generally formed of the monosaccharides; glucose,

galactose or N-acetyl-galactosamine56. These are synthesised by proteins produced by the

15 gene epsA-O operon. There is limited compositional knowledge of this

exopolysaccharide due to large heterogeneity and due to challenges in polysaccharide

sequencing. The second largest constituent of the EPS is generally the primary protein

element TasA. TasA is an amyloid-like protein that forms fibres that bind cells together57.

Deletion of tasA does not affect surface adhered biofilm formation, implying TasA is not

required for submerged biofilms. The formation of these fibres requires a secondary

protein TapA, and in turn, the production of both these proteins requires the peptidase

SiPW58. This peptidase is multifunctional, with an additional role in the adherence of

submerged biofilms to the surface. The biofilm is assembled with assistance from a

hydrophobin protein BsIA which forms a protective ‘coat’ around the biofilm19.

The EPS of P. aeruginosa PAO1 biofilms contains three polysaccharides: alginate,

Psl and Pel polysaccharides. As for B. subtilis, the presence of different P. aeruginosa EPS

components, depends on the environmental conditions, such as the growth media36.

Alginate deletion mutants develop biofilms with a decreased number of viable cells55. It has

19

also been shown that exposure to oxidative stress induces the overproduction of alginate,

which protects the biofilms from oxidative radicals59. Biofilms of alginate or psl defective

mutants fail to form complex biofilm structures, suggesting these polysaccharides are

structurally important55. The Psl polysaccharide is involved in initial attachment and biofilm

formation60. Pel is essential for the formation of a pellicle at the air-liquid interface, as well

as the formation of wrinkled colonies61.

1.3 Bacterial ion channels

Structural studies of bacterial ion channels have formed the basis of our knowledge

of the general structure of ion channels62. This is because bacterial cells are uniquely suited

to genetic manipulation and have short replication times. They are also easily cultured in

the large quantities required for the production of ion channel proteins, used in structural

analysis, via X-ray crystallography or nuclear magnetic resonance63. Genome sequencing of

ion channels from different cell types (e.g. eukaryotic cells) has confirmed evolutionary links

with bacterial ion channels. Specific genetic sequences may be used to diagnose specific

channel types. For example, the K+ filter sequence-TXGY(F)GD, is used to identify potassium

channels64. Different bacterial ion classes have been explored via structural and

electrophysical methods. We shall firstly discuss ion specific channels, followed by a

discussion on mechanosensitive channels.

The three-dimensional structures of ion specific channels have been resolved to

atomic resolution (potassium, sodium, chloride)65–67. This has allowed identification of

channel structures, such as receptors, these have provided an insight into molecular

mechanisms. Computer models have been used to further understand mechanisms

involved in gating and physiological behaviour68.

20

Structural analysis has shown that most cation specific ion channels have a similar

basic structure. They are composed of four sub units which converge to form the gate which

faces into the cytoplasm69. Various gating principles determine the state of the channel: pH,

ligand binding and membrane potential. Activation of a channel depends primarily on

opening of the gating region, but also on the conductivity of the filter. The ion filter is

located near the outer surface of the cell’s membrane and controls the channel’s specificity.

Despite the depth of structural knowledge on bacterial ion channels, little is known

regarding their role. There are exceptions to this: calcium channels have been shown to be

involved in the extreme acid resistance response and several channels have been shown to

be involved in motility and biofilm formation70. However, the primary role of most channels

remains elusive.

In contrast, the role of bacterial mechanosensitive ion channels (the other main

class of ion channel) is well established71. They act as ‘emergency valves’, releasing solutes

in osmotically challenging conditions, as well as acting as sensors of the cell’s turgor

pressure. A range of different mechanosensitive channels have been characterised via

several different techniques72. For example, electrophysical permeation studies, using large

cations and electron parametric resonance spectroscopy combined with cysteine scanning

mutagenesis and site binding labelling revealed several different channels with varying

conductances73.

1.4 Membrane potentials

Cells maintain a potential across their cell membranes when they are in the resting

state. This membrane potential is established via the asymmetric distribution of ions across

the membrane and is controlled by ion channels. Selective ion channels allow specific ions

to travel down the diffusion gradient, resulting in charge separation. This produces an

21

electrical gradient which increases until it matches the chemical gradient and there is

electrochemical equilibrium74. The equilibrium potential for a single ion is the potential at

which that ion would have no net movement across a membrane if it was the only ionic

species present. This is commonly defined as the Nernst potential75 and is given by,

𝑉𝑋 =𝑅𝑇

𝑧𝐹𝑙𝑛[𝑋]0[𝑋]𝑖

, (1.1)

where 𝑉𝑋 is the Nernst potential for given ion 𝑋, 𝑅 is the universal gas constant, 𝑇 is the

temperature in Kelvin, 𝑧 is the valence of the ionic species and 𝐹 is the Faraday constant.

Equation 1.1 is only valid for one ionic species. However, it can generally be

assumed that the main contributions to a eukaryotic cell’s membrane potential at rest

come from potassium, sodium and calcium. The distributions of these ions across a typical

cell membrane in resting state are shown in Figure 1.5.

Figure 1.5 Illustration showing the distribution of potassium, sodium and chlorine ions

across a typical phospholipid cell membrane in a eukaryotic cell.

The Goldman-Hodgkin-Katz equation takes account of the contributions from all

three of these ions and so, in general, can be used to find a good approximation of a cell’s

equilibrium potential,

𝑉𝑒𝑞 =

𝑅𝑇

𝐹ln𝑃𝑁𝑎[𝑁𝑎

+]𝑜𝑢𝑡 +𝑃𝐾[𝐾+]𝑜𝑢𝑡 +𝑃𝐶𝑙[𝐶𝑙

−]𝑜𝑢𝑡𝑃𝑁𝑎[𝑁𝑎

+]𝑖𝑛 +𝑃𝐾[𝐾+]𝑖𝑛 +𝑃𝐶𝑙[𝐶𝑙

−]𝑖𝑛,

(1.2)

22

where 𝑉𝑒𝑞 is the membrane potential, 𝑃𝑖𝑜𝑛 is the permeability for that ion, [𝑖𝑜𝑛]𝑜𝑢𝑡 is the

extracellular concentration of that ion and [𝑖𝑜𝑛]𝑖𝑛 is the intracellular concentration of that

ion.

The difference between a cell’s membrane potential and the resting potential of an

ionic species leads to efflux/influx of ions under resting conditions, this is counteracted by

actively pumping ions down their electrochemical gradients. Therefore, in most cells, the

resting potential of a cell is established due to charge separation, but is maintained by

active transport of ions across the membrane.

A cell’s membrane potential and many of its crucial physiological processes are

fundamentally linked. The membrane potential depends on the distribution of ions across

the cell membrane and in turn the transport of ions across the membrane is dependent on

the membrane potential. Many key functions of a bacterial cell are dependent on its

membrane potential70,76–79:

1. Uptake of nutrients/ions/toxins.

2. Motility.

3. Cell division is dependent on the arrangement of cell division proteins, which is

determined by the membrane potential.

4. Metabolism.

5. Regulatory pathways and transcription factors.

6. Cell growth.

7. Adhesion.

8. Quality control during sporulation.

The proton motive force (PMF) is a form of metabolic energy which drives the

uptake of many compounds and can be applied to synthesise ATP via F0F1-ATPase. In

general, the proton motive force is generated by a negative membrane potential and an

23

alkaline pH gradient across the membrane. Many of a cell’s functions rely either directly or

indirectly on the existence of a proton motive force, without a negative potential driving

this force, the cell cannot survive. Membrane potential indicators are therefore often used

as a measure of cell viability.

A negative resting potential is not the only way cells use the membrane potential.

Environmental changes may directly (e.g. increases in external ion concentration) or

indirectly (e.g. opening of ion channel by stimulus) result in a change in membrane

potential. These variations may, in turn result in further responses/behavioural changes.

For excitable cells, a signal may be triggered to communicate changes to other cells, in a

process known as an action potential. These excitable cells exploit membrane potential

changes brought about via a change in an environmental variable of interest (stimuli) to

send signals.

1.5 Membrane potentials in biofilms

The cell’s membrane potential and many of its vital physiological processes are

inextricably linked. It is therefore unsurprising that cells in a biofilm respond to changes in

membrane potential and synonymously that the biofilm membrane potential depends on

the state of the biofilms’ cells.

The electrical activity of bacteria in a biofilm has been found to depend on its

growth stage as well its environment. One study observed electrical spiking in the

membrane potential of E. coli that was sensitive to physical and chemical fluctuations80.

Another study found that the cellular response to external electrical stimuli was influenced

by the cellular proliferative capacity81.

The behaviour, physiology and growth of biofilms are modified by electrical

potentials. Application of electric fields and currents can enhance the activity of

antimicrobial agents against biofilms, in a process known as the ‘bioelectric effect’9.

24

Evidence now also exists to support the ‘electricidal effect’82, a process by which electrical

currents affect biofilm viability in the absence of antimicrobial agents. These processes have

received a substantial amount of attention owing to their potential application in

electrochemically active materials for use in medical devices83.

Bacteria in a biofilm receive increased protection, resulting in a prevalent and

stable form of microbial life. Central to this mode of growth is the ability of bacteria to

coordinate behaviour and act as a multicellular organism. Despite this, a significant amount

of biofilm regulation remains poorly understood. The similarity between bacterial ion

channels (with unknown roles) and their excitable eukaryotic counterparts (Section 1.3) is

suggestive of an analogous role in electrical signalling. The strong connection between a

cell’s electrical activity and its state provides a mechanism by which electrical signalling may

influence the behaviour of the biofilm cells.

Even though ion specific channels in bacteria are highly amenable to structural

analysis, electrophysical measurements are technically difficult, especially within biofilms.

This has made studying electrical signalling in bacterial biofilms challenging. Using a

multielectrode array, electric spiking was found to correlate with biofilm formation, leading

to the suggestion of electrical signalling as a driver in biofilm sociobiology84. More recently

fluorescent probes were used by Prindle et al. (2015)85 to present the first direct evidence

of electrical signalling between bacterial cells in a biofilm. Further studies have built on this

original work to establish a new field of biofilm electrophysiology86–88.

1.6 Outline

The aim of the thesis was to investigate the role membrane potentials play in

regulating the stress response of bacteria in biofilms:

Chapter 2 summarises the background and methodology of the experimental techniques.

25

Chapter 3 is the first results chapter. Experimental results of electrical signalling in circular,

B. subtilis, biofilms are presented alongside a mathematical model, which is used to explore

these results. These results are then discussed, including a discussion of future work.

Chapter 4 is the second results chapter. The membrane potential changes and dispersal

events which occur in P. aeruginosa and B. subtilis biofilms exposed to 405 nm light are

presented. These results are then described in terms of a Hodgkin-Huxley style model. This

is followed by a discussion, which includes ideas for future work.

Chapter 5 is the third and final results chapter. Experiments measuring the c-di-GMP levels

and membrane potential of P. aeruginosa in response to oxidative stress are presented.

These results are then discussed along with ideas for future work.

Chapter 6 is the final conclusion. The results and conclusions from the three results

chapters are summarised. Possible future extensions to this project and the future direction

of the field are then discussed.

26

CHAPTER

TWO

2 Background and methodology of experimental techniques

This chapter will present the theoretical background and methodologies of the

experimental and mathematical techniques. This will be divided into three sections:

fluorescence microscopy, microbiological techniques and mathematical modelling.

2.1 Fluorescence microscopy

2.1.1 Theory

Fluorescence microscopy is a form of optical microscopy commonly used to

visualise and quantify fluorescent molecules in order to detect the distribution of proteins

or other molecules of interest89. Specificity and the non-invasive nature of fluorescence

microscopy makes it a powerful tool.

During fluorescence microscopy experiments the specimen is illuminated with

specific wavelengths of light which are close to the absorption peaks of the target

fluorophore. This light excites the fluorophore, moving it to an excited state, the

fluorophore then emits light as it relaxes back to the ground state (Figure 2.1(a)). The

wavelength of the emitted light is normally shifted to longer wavelengths than the

absorbed light according to Stokes law (Figure 2.1(b))90. This allows differentiation between

the emitted light and the illumination light.

27

Figure 2.1. Fluorescent properties of a typical fluorophore. (a) Jablonski diagram showing

the electronic states of a fluorophore and its transitions from one to another energy level.

The thicker lines represent electronic energy levels, while the thinner lines denote the

various vibrational energy states (rotational energy states are ignored). (b) Spectral profile

of a fluorophore showing the Stokes shift observed between the excitation to emission

profiles.

The electronic states of a fluorophore are usually represented by a Jabolinski

diagram, such as the one shown in Figure 2.1(a)89,91. The principle electronic states are the

singlet ground state (S0), the singlet excited states (SN, N=1,2,3…) and the excited triplet

states (TN, N=1,2,3…). Each of these principle states contain vibrational energy levels.

Fluorophores usually contain several aromatic groups, or other molecules with numerous

π bonds. These molecules cause additional degrees of freedom which increases the number

of vibrational and rotational states of a given state. While in a given state, the fluorophore

may occupy any of the associated vibrational states, depending on the atomic nuclei and

bonding orbitals. At the temperatures used in our experiments the rotational energy is

larger than the rotational energy spacing and so these states can be ignored90,92. However,

very few molecules have enough internal energy to exist in any state other than the lowest

vibrational level of the ground state, and thus, these cannot be ignored.

Most fluorophores can repeat the process of excitation and emission hundreds to

thousands of times before the molecule becomes irreversibly photobleached.

28

Photobleaching, is defined as the permanent loss of fluorescence due to photon-induced

damage93. The dynamics of photobleaching vary greatly between different fluorescent

proteins and are highly dependent on the environmental conditions. An important type of

photobleaching involves the interaction of the fluorophore with a combination of light and

O2, therefore the O2 availability is often one of the main environmental conditions which

affects photobleaching89,90,92. As well as having a direct effect on fluorescence through

photobleaching, light can also impact fluorescence by inducing other environmental

changes. For example, hydroxyl radicals generated by photolysis of H2O2 cause a decrease

in GFP fluorescence94.

A fluorophore’s properties, such as photoresistance, lifetime and size, may

significantly affect its suitability95. Three fundamental parameters, the extinction

coefficient (ε), the quantum yield (φ) and the fluorescence lifetime (τ), are usually used to

describe a fluorophore. The extinction coefficient is a direct measurement of the ability of

a fluorophore to absorb light. The quantum yield is the probability that an excited

fluorophore will produce an emitted photon. The quantum yield of a fluorophore depends

on environmental factors, such as the pH90. The fluorescence lifetime is a measure of the

average time that a fluorophore spends in the excited state and is defined as the time at

which the fluorescence intensity decays to 1/e of its initial intensity. In an ideal system

fluorescence decay is monoexponential, while in heterogeneous systems, such as cells,

decay is more complicated and often multiexponential. In addition, other processes besides

emission can cause relaxation from the excited to ground state. An example of such a

process is quenching, which, unlike photobleaching, is often reversible. Quenching can

occur by different mechanisms. Collisional quenching occurs when an excited fluorophore

is deactivated via contact with another molecule (the quencher). There are a wide variety

of molecules which act as collisional quenchers, including O2, halogens and amines91.

Collisional quenching generally affects the excited state lifetime, as well as the quantum

29

yield. The other main type of quenching is static quenching, which occurs in the ground

state, via the formation of non-excitable molecules between the quencher and fluorophore.

During static quenching the fluorescent emission is reduced, but the excited state lifetime

is unaffected.

Fluorophores can be broadly divided into four classes: organic dyes, biological

fluorophores, quantum dots and nanodiamonds. During this project an organic dye (ThT)

and a biological fluorophore (GFP) were used. A wide range of organic dyes with a broad

range of fluorescent properties have been developed. Some are used as dyes to stain

specific structures96, others are used as indicators (e.g. of ion concentrations)85,97, while

some are used to track reagents98. For the same purpose there may exist several possible

dyes, each demonstrating differing optimum conditions, or in many cases a cost related to

their performance.

Biological fluorophores are also used in a host of different applications. The most

famous biological fluorophore is the green fluorescent protein (GFP), which is widely used

across the life sciences as a reporter of gene expression98–100. As with other fluorophores,

GFPs undergoes photobleaching, following light exposure and this can complicate time-

lapse experiments. However, it has also been exploited in physical techniques, such as FRAP

(Fluorescence recovery after photobleaching)101 and FLIP (Fluorescence Loss in

Photobleaching)102, which use photobleaching to study the motion and/or diffusion of

cellular components. Challenges in photobleaching and photobleaching techniques have

led to extensive characterisation of GFPs.

Although fluorescence microscopy is a powerful technique, traditional approaches

cannot overcome the fundamental limit enforced by diffraction. This limits the resolution

according to the Abbe limit (𝐴𝐿),

𝐴𝐿 =𝜆

2𝑁𝐴, (2.1)

where 𝑁𝐴 is the numerical aperture and 𝜆 is the illumination wavelength.

30

𝑁𝐴 = 𝑛𝑠𝑖𝑛𝜃, (2.2)

where 𝑛 is the refractive index of the medium between the objective front lens and the

specimen and 𝜃 is the aperture angle.

Alternative techniques, such as Raman scattering, TEM and SEM, can be used to

achieve a better resolution103. This is useful for examining fine biofilm structures, but the

preparation required for such techniques often destroys the biofilms native structure and

is not compatible with dynamic studies. It is also possible to overcome the Abbe limit using

super-resolution microscopy techniques, such as Stochastic Optical Reconstruction

Microscopy (STORM)104. However, these techniques require reducing/oxidizing buffers

which make live cell experiments challenging and are incompatible with studies of reactive

oxygen species. Therefore, in vivo, dynamical studies of biofilm structure and behaviour

generally still use traditional fluorescence microscopy.

2.1.2 Experimental methodology

Fluorescence microscopy experiments were performed using two microscopes. A

Zeiss LSM 5 Pascal fluorescence microscope was used to study the propagation of electrical

signals in B. subtilis biofilms and an Olympus IX-71 inverted fluorescence microscope was

used for all other experiments. The Zeiss LSM 5 Pascal microscope was assembled with a

Zeiss Temperature module, Zeiss CO2 module and encased in an incubator. Excitation was

provided by: a 405 nm diode laser, a 458/477/488/514 nm argon ion laser, a 543 nm HeNe

laser and a 633 nm HeNe laser. All the moving components, such as emission filter wheels,

main and secondary dichroic beam splitters, the pinhole and the mechanical attenuators

for each laser line were computer controlled. Temperature control and data acquisition

were also computer controlled. The Zeiss AIM software was used to acquire time lapse

experiments. The autofocus capabilities of this system made it well suited for long-term

31

time lapses. The other main reason it was chosen was because it was equipped with the

CellASIC ONIX microfluidics system.

The Olympus IX-71 inverted fluorescence microscope (Figure 2.2) was custom-built

and primarily designed for stochastic optical reconstruction microscopy (STORM).

Excitation was provided by OBIS 405LX, OBIS 488LX or OBIS 647LX lasers. The laser lines

were directed into an optical fibre, which was physically oscillated at 10k Hz to reduce the

spatial coherence, which can otherwise cause inference patterns. The laser beams were

directed by a combination of regular and dichroic mirrors into the microscope. The beams

were selectively reflected by a cube that contained a Semrock Brightline full-multiband

laser filter set. The sample was mounted on top of an Olympus 100x TIRF lens, which was

itself mounted on a laser-controlled self-correcting MadCity piezoelectric stage (which

reduced drift in the z-direction). Olympus UPAON 100XOTIRFM (NA 1.49) immersion oil was

used between the sample and lens. Time lapse images were recorded using an ORCA-

Flash4.0 LT PLUS Digital CMOS camera (C11440-22CU) (82 % peak quantum efficiency, 6.5

μm x 6.5 μm pixel size, 2.5 ms exposure time). Image acquisition was performed using the

software HCimage Live or using Micro-Manager. The lasers were controlled using OBIS

LX/LS Scientific Remote in conjunction with either OBIS Connection software or Micro-

Manager. The microscope was enclosed by a Solent Solutions incubator and the entire set-

up was set on a Newport floating optical table.

32

Figure 2.2. Schematic diagram of the custom-built Olympus IX-71 inverted fluorescence

microscope. The laser beams were guided into the microscope by a combination of regular

and dichroic mirrors. The lasers were selectively filtered by a cube that contained a Semrock

Brightline full-multiband laser filter set. Fluorescence was detected using an ORCA-Flash4.0

LT PLUS Digital CMOS camera.

Thioflavin-T was used as an indicator of membrane potential. Thioflavin-T (ThT;

Sigma) is commonly used to stain amyloid fibres, however, its positive charge also allowed

it to be used as a Nernstian voltage indicator. Its suitability in the role as a membrane

potential indicator in bacterial cells was established by Prindle (2015)85, since then it has

become widely used as an indicator of membrane potential in bacteria81,86,87. It is cheaper

and for biofilm studies, often more sensitive, than other commonly used membrane

potential indicators (such as DiSC3(5)). Thioflavin-T is a benzothiazole salt, which is soluble

in water to 25 mM and in dimethyl sulfoxide (DMSO) to 100 mM. Fresh 2 mM stocks in

water were made up on the day of the experiments and added to the cells and media at a

final working concentration of 10 µM.

33

To monitor and compare c-di-GMP levels the GFP-based fluorescent reporter, P.

aeruginosa PA01 pCdrA::gfpc, was used. The GFP has been widely used across the life

sciences as a reporter of gene expression99,100,105.

2.2 Microbiological techniques

2.2.1 Background

In natural environments bacteria almost exclusively form multispecies biofilms. The

role one species can play on another can be significant106. For example, B. subtilis can

protect S. aureus from biocide action107,108. Despite this the majority of laboratory grown

biofilms contain a single species. The majority of studies also focus on ‘model species’109.

Focusing on highly amenable, model species has facilitated investigation of their

biofilms. Many basic mechanisms and structures observed in biofilms are ubiquitous, even

if their details vary depending on the exact system. For example, all bacteria in a biofilm are

contained in an extracellular polymeric substance (EPS), even if its exact composition varies.

Therefore, even though it may have limited the applicability of results, the characterisation

of model, mono-species biofilms has provided us with generic knowledge on the formation,

structure and growth of biofilms110. This is illustrated by the progression made studying the

two model species using in this project: P. aeruginosa and B. subtilis. P. aeruginosa biofilms

have been central to the study of c-di-GMP signalling 43 and B. subtilis biofilms have been

key in the examination of molecular mechanisms of biofilm formation and regulation20.

Biofilms are routinely grown in flow cells. Flow cells allow hydrodynamic conditions

to be employed in an easily controllable environment which can be changed rapidly and

chemostats improve the reproducibility of biofilm growth111.

34

Microfluidics Microfluidics involves the manipulation of fluids on the order of microlitres to

picolitres112. There are a host of unique properties which emerge at these scales.

Microfluidic approaches benefit from a larger flexibility in design compared to macroscale

equivalents113. This allows devices to be tailored to the requirements of the experiment and

produces unprecedented control over the flow conditions. At the microscale the flow

through a device changes from turbulent to laminar, increasing the predictability,

reproducibility and control. The system’s compact size means multiple channels may be

implemented on one device, allowing multiple experiments to be conducted

simultaneously. The small size and high throughput also save time and money (due to the

reduction in the quantity of reagents required). This makes microfluidics an attractive

alternative to traditional macroscale approaches.

However, the transition from macroscale experiments to microscale equivalents, is

challenging. Differences between the behaviour of fluids in microfluidic devices and the

properties of the materials used to make them demands consideration be given to the

protocols and approaches used. Most conventional macroscopic flow cell devices are made

either of glass or of polystyrene, which have both been widely accepted as biocompatible.

Microfluidic devices on the other hand are generally made of polydimethylsiloxane (PDMS).

PDMS has a higher gas permeability than traditional materials and is more permeable to

CO2 than O2 or N2114. Consideration must therefore be given to the gas levels within a device.

The effect of PDMS on cultured cells is still disputed. Some studies show that the

viability of cells grown in PDMS devices is compromised. For example, differences were

found between gene expression profiles of PC12 cells differentiated on PMMA versus

PMMA-PDMS surfaces115. It has been suggested that this may be due to contamination of

the media via leaching of artefacts from the PDMS. Uncross-linked oligomers may be

removed by further curing or additional preparation steps, and so the level of

35

contamination will depend on the PDMS used and the method of fabrication. In one study,

a continuous range of uncrosslinked PDMS was detected in deionized ultra-filtered water,

after incubation in a PDMS microfluidic device; despite efforts to extract uncrosslinked

PDMS with ethanol, in a Soxhlet extractor overnight116. PDMS is also known to absorb small

hydrophobic molecules and proteins116. This may affect the concentration of different

media and EPS components.

In microfluidic devices the PDMS surface area to media volume can be up to 30

mm2/μL, which is significantly higher than the standard macroscopic media culture to

surface area volume of 0.5 mm2/μL113. This can affect cell proliferation and the reagent

concentrations. For example, the proliferation of mouse mammary fibroblasts was

impaired at higher PDMS surface area to volume ratios and consumption of nutrients, such

as glucose, increased. This behaviour became more pronounced as the surface area to

volume ratio increased117.

In conclusion, studies show that the biocompatibility and behaviour of cell lines in

a PDMS device will depend on the devices’ size and preparation. Therefore, the viability,

behaviour and growth of each cell line in a specific microfluidic device must be individually

considered. In cases where results are found to be significantly affected, alterations, such

as surface coatings, may be made. In extreme cases, where the PDMS is found to

unavoidably interfere with the system, it is possible to fabricate microfluidic devices from

alternative materials. However, these options are limited and expensive.

Experimental challenges may arise due to the changes in the behaviour of the fluid

in microfluidic devices118. For example, in the presence of laminar flow, fluids mix only via

diffusion, which can make reactions harder to achieve119. There are also other basic

practical issues with using microfluidic devices. The small size makes devices more

susceptible to air bubbles and the tubing and chambers are more likely to become

36

clogged120. In many microfluidic devices a closed set up also prevents the biofilm being

accessed for molecular analysis.

In general microfluidics offers a novel way in which to grow bacteria under flow as

long as careful consideration is given to the adaptations that need to be made to

compensate for changes and challenges associated with handling fluids on this length scale.

2.2.2 Experimental methodology

Cell culture

Figure 2.3. Step-by-step schematic showing the basic process used to culture bacterial cells.

Cells were streaked on to an agar plate, which was then incubated overnight. A single colony

from the plate was then picked from the plate and used to inoculate the culture which, after

further incubation, was used for further cell culture or to grow a biofilm, depending on the

experiment.

37

The basic procedure for cell culture is shown in Figure 2.3. Bacterial stocks were

stored, 1:1(v/v) 50 % glycerol (Sigma; made with Milli-Q H20), in a freezer at -80°C. Cells

were streaked on medium agar plates. The agar plate mix was made by combining liquid

medium with 1.5 % (w/v) agar (Sigma) in a glass media bottle (Cole-Parmer) and autoclaving

it. The mixture was microwaved to dissolve it, before it was poured into plates. The streaked

plates were grown, overnight, in a static incubator at 37°C. A single colony was picked from

the plate and used to inoculate a glass universal containing liquid media, this was then

incubated and shaken at 37 °C. The length of time the sample was incubated for and the

media used depended on the experiment.

Biofilm growth Two different experimental set-ups were used to grow biofilms. Firstly, following

the work of Prindle (2015)85, thin, circular biofilms were grown in the CellASIC ONIX

microfluidics system (Figure 2.4(a)). The two-dimensional nature of these biofilms

simplified the data analysis and modelling. It also subjected the inner biofilm to nutrient

deprivation, which in turn caused electrical signalling.

The Y04D microfluidic plates consisted of four independent chambers, which were all

connected to six media wells, a cell inlet well and a waste outlet (Figure 3.2(a)). The flow

was controlled via a pneumatic system which used pressurized air to pump the media and

cell suspensions through the plate, resulting in high precision, even at low flow rates. The

pressure-driven system provided a stable, pulseless flow with fast response times. The gas

tube of the chamber manifold pumped in an atmospheric gas mixture. PDMS is more

permeable to CO2 than O2/N2, so it is possible that the gas levels in the chamber were not

exactly maintained114.

This experimental system was used to follow the work of Prindle (2015)85, but it was

unpredictable. The Y04D microfluidic plates were originally developed for yeast growth and

38

biofilm experiments were only possibly in these plates due to plate defects. These defects

included a rougher surface finish, which allowed the cells to adhere to the surface. These

defects have since been rectified and so these experiments are no longer possible. As

described in Section 3.3.2, the protocols for cell loading and washing had to be varied

depending on the specific plate. There were also issues with unpredictable flows, backflows

and cells clogging the plates. This experimental set-up was tested for P. aeruginosa, but the

highly motile nature of this bacteria caused even more problems with clogging. The opening

of the plate to change media or empty waste also sometimes introduced flows and bubbles.

Taken in conjunction, these issues meant the experimental system was not perfectly robust.

Therefore, the rest of the experiments were conducted in a more reproducible bespoke

flow system, based on a syringe-pump and Ibidi µ-slides (Figure 2.4(b)). The use of a syringe

pump avoided the issues of unpredictable flows associated with the pressure driven

system. The Ibidi µ-slides are microfluidic slides which have gas-permeable polymer

coverslips, this ensured that CO2/O2 exchange during cell culture was maintained. These

coverslips were treated with IbiTreat, which is Ibidi’s most common surface medication and

is used to promote the adhesion of cells. IbidiTreat is a chemically modified polymer surface

which is comparable to other standard tissue-treated surfaces121,122.

In addition to growing biofilms under flow, to monitor C-di-GMP levels, cells were

cultured in microtiter plates. Flow cell experiments were also supplemented with

experiments that immobilised cells on an agar microscope slide. Detailed protocols and

description of each of the experimental set-ups can be found in the relevant results

chapters.

39

Figure 2.4. Schematic of the two different experimental set-ups used to grow biofilms. (a)

CellASIC ONIX microfluidic experimental set-up. (b) Syringe pump flow cell experimental set-

up.

40

2.3 Mathematical modelling of excitable systems

2.3.1 Theoretical background

As described in Chapter 1 (Section 1.4), a variation in ion concentrations across a

cell membrane results in a potential difference. The regulation of this membrane potential,

by ionic channels, is one of the most important functions of a cell.

Different cells have different ion channels, leading to varying membrane potentials,

as well as responses. Many different types of excitable cells have been successfully

modelled following the work of Hodgkin and Huxley123–127.

Hodgkin-Huxley Model

The Hodgkin-Huxley model is a mathematical model that has been used and

adapted since it was originally described in 1952 to build a range of models of varying

complexity that define the electrical characteristics of excitable cells. The longevity of this

model is testament to its success in accurately simulating and parametrising a range of

systems, from the original squid axon to cardiac muscles126,128. In such a model each

element of a cell is defined in terms of its electrical properties, resulting in a set of coupled

differential equations, which are based on the current flowing through a cell’s membrane

and the flow of ions through the cell’s ion channels. A simple electrical circuit model of a

cell membrane (Figure 2.5) has the cell membrane as a capacitor in parallel with a resistor

(ionic current (𝐼𝑖𝑜𝑛)). The membrane capacitance (𝐶𝑚) is given by

𝐶𝑚 = 𝑄𝑉, (2.3)

where 𝑄 is the charge across the capacitor and V is the potential voltage required to hold

the charge.

41

Figure 2.5 Simple model of a cell membrane with a capacitor (𝐶𝑚) in parallel with a resistor.

If 𝐶𝑚 is constant, then the current across a capacitor (𝐼𝑐𝑎𝑝) equals

𝐼𝑐𝑎𝑝 =𝑑𝑄

𝑑𝑡, (2.4)

Therefore,

𝐼𝑐𝑎𝑝 = 𝐶𝑚𝑑𝑉

𝑑𝑡.

(2.5)

The sum of the capacitive current and the ionic current ((𝐼𝑖𝑜𝑛) must be zero if there

is no build-up of charge on either side of the membrane yielding the equation

𝐶𝑚𝑑𝑉

𝑑𝑡+ 𝐼𝑖𝑜𝑛 = 0

(2.6)

where 𝑉 = 𝑉𝑖 − 𝑉𝑒 is the membrane potential, 𝑉𝑖 is the intracellular potential and 𝑉𝑒 is the

extracellular potential. For neurons the two primary ionic currents are sodium and

potassium. All other ionic currents are small enough that they can be lumped together in a

generic ‘leakage current’. The sodium, potassium and leakage currents (𝐼𝑘 , 𝐼𝑁𝑎 , 𝐼𝑙 ) are all

considered linear and are given by,

𝐼𝑖𝑜𝑛 = 𝑔𝑖𝑜𝑛(𝑉 − 𝑉𝑖𝑜𝑛). (2.7)

42

Substituting Equation 2.7 into Equation 2.6 leads to the classic equation

𝐶𝑚𝑑𝑉

𝑑𝑡= −𝑔𝑘𝑛

4(𝑉 − 𝑉𝑘)𝑔𝑖𝑜𝑛 −𝑔𝑁𝑎𝑚3ℎ(𝑉 − 𝑉𝑁𝑎) − 𝑔𝑙(𝑉 − 𝑉𝑙),

(2.8)

where 𝑔𝑘 , 𝑔𝑁𝑎 and 𝑔𝑙 are the respective channel conductance constants and 𝑉𝑘, 𝑉𝑁𝑎 and

𝑉𝑙 are the specific ions reversal (Nernst) potentials.

Hodgkin and Huxley performed voltage clamp experiments to determine the

behaviour of the conductances. Secondary variables were then chosen to represent the

observed conductances. For potassium, the secondary variable 𝑛 was introduced, and the

fourth power was chosen as the smallest exponent which agreed reasonably with

experimental data. Sodium conductance was more complex than potassium conductance

in squid axons. Based on experimental data Hodgkin and Huxley suggested two secondary

variables, 𝑚 and ℎ, to represent sodium conductance. The secondary variables 𝑛,𝑚 and ℎ

were then assumed to obey the differential equation,

𝑑𝑖

𝑑𝑡= 𝛼𝑖(1− 𝑖) − 𝛽𝑖𝑖, (2.9)

where 𝛼𝑖 and 𝛽𝑖 are the transition rate constants. 𝛼𝑖 is the number of times per sec that a

gate in the shut state opens. 𝛽𝑖 is the number of times per sec that a gate in the open state

closes. Through fitting experimental data these were found to depend on the voltage in the

following manner

𝛼𝑛 =0.01(10− 𝑣)

exp (10− 𝑣10 )− 1

(2.10)

𝛽𝑛 = 0.125exp (−𝑣

80) (2.11)

𝛼𝑚 =

0.1(25− 𝑣)

exp(25 − 𝑣10

) − 1

(2.12)

43

𝛽𝑚 = 4exp (−𝑣

18) (2.13)

𝛼ℎ = 0.07exp (−𝑣

20) (2.13)

𝛽ℎ =

1

exp(30− 𝑣10 ) + 1

(2.14)

where 𝑣 is the potential deviation from rest (𝑉 − 𝑉𝑒𝑞) and the constants (e.g. 0.01, 10, 80)

were found by fitting with experimental data.

Following on from the work of Hodgkin and Huxley, FitzHugh (1960, 1961,

1969)125,129,130 suggested a simple qualitative description of the above equations and split

them into fast and slow variables. Separation in this way retained most of the model’s key

features while simplifying further mathematical analysis.

More detailed analyses have also been given by Rinzel (1978)123, Hassard (1978)131

and Sabah & Spangler (1970)132. These were all instrumental in forming a new field of

applied mathematics, ‘the study of excitable systems’, which is still widely studied today.

Subsequent models of increasing complexity which use multiple continuous state variables

have been developed to describe action potentials in a broad range of different cell

types128,133,134.

Reaction-Diffusion Equations

Reaction-diffusion equations can be used to describe the concentration of one or

more chemical species in space and time. They take the form of a semi-linear parabolic

partial differential equation,

𝜕𝒒

𝜕𝑡= 𝑫𝛁2𝑞 +𝑹(𝒒)

(2.15)

44

where 𝒒 = 𝑞(𝒙, 𝑡) is the concentration of chemical species, 𝑫 is a diagonal matrix of

diffusion coefficients and 𝑹(𝒒) takes account of local reactions.

The simplest form of the reaction-diffusion equation is for one species (𝑢) in two

dimensions (𝑥) and (𝑡) and is referred to as the Kolmogorov-Petrovsky-Piskunov equation,

𝜕𝑢

𝜕𝑡= 𝐷

𝜕2𝑢

𝑑𝑥2+𝑅(𝑢).

(2.16)

A wide range of different reaction-diffusion systems are found across a broad range

of different disciplines from physics to biology, engineering and medicine135,136. These

systems are described by a set of boundary conditions and one or more partial differential

equations, which often cannot be solved exactly. There are a range of different analytical137

and numerical138 methods which take advantage of variable transformations along with

stability analysis to provide a range of different solutions. Standard forms include travelling

waves, wavelike phenomena and self-organised patterns e.g. spirals and stripes124.

Different methods are often used to confirm a single solution. Less accurate

methods may still retain the solution to a sufficient approximation, while saving on

computational time and allowing further analytical analysis.

Agent-based modelling Excitable tissue is classically modelled using reaction diffusion equations as

described above. The diffusing species is given by a continuous variable represented using

partial differential equations (PDEs) and a system of nonlinear ordinary differential

equations describes all other state variables. The complex nature of these systems not only

hinders formal analysis, but also makes them computationally expensive, especially in

large-scale simulations.

Agent-based models (ABM) are a class of models used to simulate the interactions

of individual agents to obtain the global behaviour. A set of rules is created for the

interaction of each element and the emergent behaviour can be understood by updating

45

the agents’ behaviour. These models are well suited for modelling multicellular systems,

such as cell colonies, in which higher-level properties emerge from interactions between

constituent cells. Agent-based modelling has been employed to study a variety of biofilm

behaviours, from detachment139 to mutation rates140 and growth141.

Modelling excitable biofilms The recent discovery that electrical signalling plays a role in bacterial

communication has led to the development of several new models. The original work of

Prindle (2015) included a simple Hodgkin-Huxley style model which was used to show that

the behaviour could be explained via their proposed mechanism of potassium signalling.

This model was extended into space using a one-dimensional lattice. Subsequent studies

have built on this work and model. One study added metabolic components to the electrical

Prindle model to build a discretised one-dimensional reaction-diffusion model142. Another

study captured the oscillatory nature of biofilm expansion using a similar, one-dimensional,

minimal reaction-diffusion model that included bacterial growth, nutrient consumption and

electrical signalling143.

Other related models have also been used to describe associated behaviours, for

example an agent-based model was used to describe the attraction of motile bacteria

towards a biofilm87. In this model the electrophysiological model of Prindle (2015) was

combined with an adapted version of a bacterial mechanical agent-based model developed

in earlier work144. The dynamics of extracellular potassium were included in the model via

a reaction diffusion equation, in which it was assumed that the potassium was produced

and absorbed by the biofilm periodically. Additionally, the synchronisation of biofilm

growth between two distant biofilms was simulated by modelling the two biofilms as non-

linear phase oscillators86. In one study, a model based on percolation theory was developed

to describe how signals propagate through biofilms. This model predicted that propagation

46

is only possible when the community is organized near a critical phase transition between

a disconnected and a fully connected conduit of signalling cells88. A minimal delay-

differential equation (DDE) was also used to explain the experimentally observed

discontinuous emergence of biofilm oscillations at a critical size145.

2.3.2 Modelling methodology

During the early stages of this project the original model of Prindle (2015) was

extended to include nutrient components and to better fit experimental results. The results

produced by this model were consistent with those produced by the later model of

Martinez-Corral (2019)142, for this reason and because of the one-dimensionality of this

model, these results are not presented in this thesis. Instead this thesis focuses on two-

dimensional modelling of electrical signalling in biofilms which is more physically realistic.

In order to achieve two-dimensional modelling, the description of potassium firing by cells

was simplified and an agent-based model was employed. Firstly, electrical signalling in

circular biofilms was described using a one-dimensional fire-diffuse-fire model, such

models were originally developed to describe intracellular calcium propagation146. This

model was then extended into two-dimensions using agent-based modelling. This model

was simulated using the software package Gro147, which was developed to simulate

bacterial colonies.

In addition to agent-based models this project used a non-linear Hodgkin-Huxley

style model to understand how differences, such as the mode of growth, may affect the

membrane potential dynamics.

47

CHAPTER

THREE

3 Spatial propagation of electrical signals in circular biofilms

3.1 Overview

Biofilms are explored as a class of active excitable matter in which cell division is

the active process and the spiking of the individual bacterial cells is the excitable process.

It is demonstrated how new methods of signal analysis, combined with agent-based

modelling can be used to further understand this important new class of excitable matter.

Moment analysis is used to quantify the propagation of electrical wavefronts through

circular biofilms and agent-based models are used to simulate the propagation of

wavefronts and to compare different signals.

3.2 Introduction

The strong motivation for studying biofilm growth is detailed in Chapter 0.

Fundamental to this form of growth is the ability of bacteria to coordinate behaviour and

act as a multicellular organism. It is hoped that a better understanding of biofilm regulation

would allow the development of new techniques to tackle, as well as use, biofilms. Electrical

signalling has recently emerged as a regulator of biofilm growth. The initiation and

propagation of electrical signals in eukaryotic excitable tissues has been studied by

electrophysiologists, across a broad range of disciplines, for well over a century128,148,149. In

contrast, the study of bacterial electrical signalling is only in its infancy. This is due to

difficulties studying bacteria via traditional electrophysical methods, such as patch clamps,

48

owing to their smaller size. Prindle et al. (2015)85 overcame this issue by using fluorescence

microscopy to provide the first direct evidence of electrical signalling between bacterial

cells in a biofilm.

The Prindle (2015) study found that B. subtilis cells communicate nutrient stress via

electrical signalling85. The cells in the biofilm interior were starved of glutamate and this

triggered the opening of voltage gated Yug0 potassium channels. Outer biofilm cells

responded to this signal with periodic reductions in growth. This allowed sufficient nutrients

to reach the interior cells and increased the resilience of the entire community to chemical

attack.

Structural analysis of Yug0 revealed a TrkA gating region. Other K+ uptake systems

assembled with TrkA (TrKH/G) show regulation by binding of metabolic products, such as

NAD+/NADH or ATP, to TrkA150. Binding induces conformational changes in TrkA, which in

turn leads to changes in the activity of the ion channel151. Similarly, stress products

produced due to a lack of glutamate, such as excess NADH, may bind to the TrkA gating

region of Yug0, triggering it to open.

In response to a transient increase in external potassium, potassium was released

by wild type cells, but not by those of a Yug0 deletion strain. This supports the hypothesis

that a lack of glutamate triggered the opening of Yug0 in inner cells, initiating a potassium

wave and that this wave was then actively propagated by the further opening of Yug0

channels in other cells, due to the depolarisation caused by the potassium wave. The uptake

of glutamate into the cell by the GltP transporter is dependent on the proton motive force

and was therefore affected by the depolarisation caused by the potassium wave. It is

argued by Prindle (2015) that it was by this mechanism that potassium mediated electrical

signalling coordinated the metabolic stress (Figure 3.1)85.

49

Figure 3.1. Proposed mechanism of active propagation of potassium through B. subtilis

biofilms85. The initial trigger for potassium release via Yug0 channel is metabolic stress, due

glutamate limitation. External potassium depolarizes neighbouring cells, limiting glutamate

uptake and thus produces further metabolic stress. This cycle results in the active

propagation of potassium through the biofilm.

Subsequent studies have shown that the growth oscillations of two distant biofilms

were coupled through electrical signalling, resulting in a synchronization of growth

dynamics, which allowed the biofilms to resolve nutrient competition through time-

sharing86. This shows the range over which electrical signalling may coordinate behaviour

and demonstrates how biofilms use strategies comparable to those seen in engineered

systems.

The role of these electrical signals extends beyond this initial application. For

example, these electrical signals were found to alter the motility of interacting planktonic

B. subtilis and P. aeruginosa cells86. This implies that electrical signalling in biofilms may be

a generic form of bacterial communication. Potassium has been shown to regulate and

affect several key processes in B. subtilis cells. It is therefore expected that a range of other

roles for potassium signalling will emerge following investigation of the system developed

by Prindle (2015). More broadly there are also many other voltage-gated ion channels,

besides Yug0, that have been identified in the genome of a wide range of bacterial species62.

50

The primary role of many of these ion channels remains elusive. Communication is crucial,

not only in metabolic regulation and motility, but also in many other key processes in a

biofilm (Chapter 0). As membrane potentials have been shown to regulate a broad range

of bacterial and biofilm associated behaviour it is logical to infer that electrical

communication is likely to extend far beyond these initial studies. It is expected that signals

are initiated in response to a range of stimuli and that these signals elicit a broad range of

different responses. As most cells are responsive to changes in local ion concentrations and

potentials, it is also probable that electrical signalling could be more diverse than just inter-

kingdom signalling. For example, it is possible that electrical signalling could be involved in

host-bacteria interactions, as seen for quorum sensing26.

Contrary to the slow development of biofilm electrophysiology compared to

eukaryotic electrophysiology, the molecular study of bacterial ion channels has informed

the basis of our knowledge of the generic structure of many ion channels62. Genome

sequencing of ion channels from different cell types (e.g. eukaryotic cells) has confirmed

evolutionary links with bacterial ion channels. The signalling observed by Prindle (2015)

shares several similarities with electrical signalling in the nervous system, such as the use

of glutamate as a neurotransmitter152. Also, the role of potassium in the propagation of

metabolic stress, is analogous to the role it plays in driving the dilation of blood vessels in

mammalian brains in response to metabolic stress153. These similarities reinforce the

assumption that electrical signalling in bacteria is widespread and that bacteria can be used

to inform our current knowledge on electrical signalling in general.

If electrical signalling is found to be a universal mechanism of communication, it

would be a very attractive target from which to develop treatments, as they may offer

widespread solutions. In order to fully exploit electrical signalling in bacterial biofilms, it

may be possible to use tools developed by other more established fields of active excitable

matter.

51

Traditional biophysical methods (e.g. patch clamps, electrocardiograms,

electroencephalography etc.) have been combined with computer modelling to advance

the study of excitable tissues126,154,155. Mathematical models provide a powerful tool for

understanding complex biological systems. Models can be used to explain a system, to

study the interactions of different components and to make predictions on behaviours.

Since the pioneering work of Hodgkin-Huxley155, electrophysiologists have developed a host

of mathematical modelling techniques that have been used to probe, understand and

predict the behaviour of excitable systems (Section 2.3). For example, today’s models offer

mechanistic insights into a range of cardiac dynamics, across a range of species (including

humans) 128. Models are also used to predict the behaviour of altered states e.g. disease.

The study of active excitable matter, therefore, provides one of the best examples of how

experimental and mathematical methods can be combined to better understand complex

and intricate systems. Well-established techniques and tools developed for studying active

excitable tissues were used to inform our study of the relatively new class of active excitable

matter - bacterial biofilms. Moment analysis was used to quantify electrical signal

propagation in detail. Agent-based modelling was then used to investigate the role of

spatial effects on signal propagation. Agent-based models are uniquely placed to offer

insights into the behaviour of complex systems, such as biofilms, due to their ability to

integrate combinations of spatially heterogeneous processes156–159. These models are

commonly used to study how complex global behaviours emerge from the interaction of

‘simple’ behaviours of individual agents (cells) and are especially useful in the study of

multicellular systems. This has led to the development of several dedicated agent-based

model simulators for bacterial cell colonies. Agent-based modelling has been used to

understand a range of biofilm behaviours, such as detachment139, mutation rates140 and

growth141.

52

3.3 Materials and methods

3.3.1 Cell culture and growth

Experiments were conducted using B. subtilis NCIB 3610, a wild type strain that has

retained its ability to form biofilms. Biofilm growth was conducted following the protocols

of Prindle (2015)85. Cells were freshly streaked onto LB agar plates from glycerol stocks 1

day before the experiment and incubated at 37oC overnight. The next day, 3 ml of LB was

inoculated with a single colony. The inoculum was then incubated and shaken at 200 rpm,

at 37°C, for approximately three hrs or until the cells reached an OD600 of 0.7 - 1.2. The

optical density of cells was measured using a spectrophotometer. At this stage the cells

were centrifuged at 2,100 rcf for 1 minute and then resuspended in a minimal MSgg

medium to promote biofilm growth. The recipes for all media can be found in Table 3.I.

Table 3.I. Recipes and sources for the culture media used in this chapter.

Media recipe

LB

10 g/l NaCl, 5 g/l Yeast extract, 10 g/l Tryptone, supplemented with antibiotics as required.

LB agar 10 g/l NaCl, 5 g/l Yeast extract, 10 g/l Tryptone, 15 g/l agar, supplemented with antibiotics as required.

Msgg 5 mM potassium phosphate buffer (pH 7.0), 100mM MOPS buffer (pH 7.0 adjusted using NaOH), 2mM MgCl2, 700 µM CaCl2, 50 µM MnCl2, 100muM FeCl3, 1µM ZnCl2, 2µM thiamine HCl, 0.5% (v/v) glycerol and 0.5% (w/v) monosodium glutamate 160.

53

3.3.2 Biofilm growth

Biofilms were grown in the CellASIC ONIX microfluidics system, in Y04D microfluidic

plates, following the work of Prindle (2015)85. The Y04D microfluidic plates consisted of four

independent chambers, which were all connected to six media wells, a cell inlet well and a

waste outlet (Figure 3.2(a)). These chambers were originally developed for yeast cells and

were 3 x 3 mm, with a depth of 5 - 7 μm (Figure 3.2(b)). In this geometry biofilms grew in

two-dimensional circles, centred around the cell flow traps (Figure 3.2(c)). The shallowness

of the chambers ensured that cells remained within a single focal plane during perfusion

and prevented media reaching cells at the centre of the biofilm. The microfluidic chamber

plate was vacuum sealed to the microfluidic system, ensuring that each well was

independent and that the flow parameters were accurate.

The flow was controlled via a pneumatic system which used pressurized air to pump

the media and cell suspensions through the plate, resulting in high precision, pulseless flow

with fast response times even at low flow rates. The only relevant disadvantage of using a

pressure pump, as opposed to a syringe pump, was the possibility of backflows or unknown

flow rates161.

The chamber was placed on the Zeiss LSM 5 Pascal microscope, which was encased

in an incubator set to 30°C to promote biofilm growth. The gas tube of the chamber

manifold pumped in an atmospheric gas mixture. This filled specific air channels in the

microfluidic plate and entered the chamber through its permeable walls. PDMS is more

permeable to CO2 than O2/N2, so it is possible the gas levels in the chamber were not exactly

maintained114.

54

Figure 3.2. Illustrative figure showing how biofilms were grown in the CellASIC ONIX Y04D

plate (not to scale). (a) Schematic of a whole CellASIC ONIX Y04D plate, showing the four

identical, separate chambers, each with 6 inlet wells, a waste outlet well and a cell inlet

well. (b) Cell culture chamber, with six media inlets, waste outlet, cell inlet and six cell traps.

(c) Representative image of a circular B. subtilis biofilm grown overnight in a microfluidic

chamber. Biofilm cells were stained with the membrane potential dye ThT.

The Y04D plate chambers were primed with MSgg before loading, and columns 1

to 6 of the plate were preloaded with the required media or media/dye solutions (Figure

3.2). Immediately after resuspension in MSgg the cell suspension was pipetted into the cell

inlet well and the plate was sealed on to the microfluidic device. The bonding of the

55

microfluidic plates varied and so the protocol for cell loading had to be varied accordingly.

Cells were loaded in 5 s bursts at pressures of 4 - 5 Psi to establish an appropriate pressure

which could be used to trap cells under the cell traps. At the high pressures used during cell

loading the flexible top of the chamber was pushed upwards this allowed cells to flow into

the cell traps. When cell loading ended these cells were confined to the cell trap when the

top of the chamber relaxed, leading to their adherence in these regions, circular biofilms

grew out from these initial cells. After a pressure was chosen, cells were loaded in 10 s

bursts, with 10 s in between bursts. The number of loading bursts varied depending on the

plate and cells were loaded until a sufficient number were trapped under the cell trap

(usually 6 - 8 loading bursts). The cell chambers were then washed at 3 psi (channels 1 - 6)

for 1 min. Low flow rates (0.25 psi, channels 3 and 4) were used for the first 2 hrs to help

cells adapt after loading. This was followed by fast flow (2 psi, channels 2 and 5) for 6 - 10

hrs. An intermediate flow rate (1.5 psi, channel 3 or 4) was then used for the rest of the

experiment.

3.3.3 Microscopy

The Zeiss LSM 5 Pascal fluorescence microscope was used for phase contrast and

fluorescence imaging. In general, a 20x objective lens was used to conduct time lapse

studies and images were taken every 5 min.

3.3.4 Dyes

Thioflavin-T (ThT; Sigma) is commonly used to stain amyloid fibres96, however, its

positive charge also allows it to be used as a Nernstian voltage indicator. Its suitability in

the role as a membrane potential indicator in bacterial cells was established by Prindle

56

(201585). ThT was used at 10 µM. In this set-up, it has a three-fold higher sensitivity to

membrane potential changes than the commonly used membrane potential indicator

DiSC3(5)85.

3.3.5 Data analysis

Image analysis was conducted in Matlab (MathWorks) using custom made scripts.

As previously described (Section 3.3.2), the B. subtilis cells grew in thin circular biofilms

centred around the flow traps (Figure 3.2(c)) and this motivated the use of polar

coordinates. Figure 3.3 shows a representative ThT fluorescence profile as a function of

time at all angles in a single radius and the average signal obtained by averaging these

signals. The coefficient of variation of the average signal’s radial mean was always less than

0.07, justifying the assumption that the average radial signal was representative of the

whole radial signal. Matlab was used to subtract the background noise and smooth the data

using a moving average filter. The offset signal due to the cell trap was negligible. Graphs

and fits were produced in both Matlab and Origin. Matlab was used to produce contour

plots of the ThT fluorescence in space and time.

57

Figure 3.3. Electrical wavefront from a B. subtilis biofilm. ThT fluorescence observed at 4 µm

from the centre of the biofilm as a function time. Signals from all angles are shown in blue

and the average signal is shown in red.

The ThT signals were quantified using moments’ analysis, integration to obtain the

fluorescent energy density and Gaussian fitting. These were all conducted in Matlab using

custom made scripts. These parameters were then exported to Origin where they were

fitted and plotted.

3.3.6 Modelling

Matlab was used to create our simple FDF model. Custom made scripts were used to

check for stable solutions and to obtain final solutions. Matlab was also used to plot

solutions to the FDF model. Our agent-based fire-diffuse-fire model (ABFDF) model was

created using the extended version of Gro147. The data was then imported to Matlab, where

it was analysed using the same methods as the experimental data. Matlab and Origin were

then used to obtain fits and plots as previously described for the experimental data.

Biofilm growth was controlled by CellEngine (in the extended version of Gro).

CellEngine was developed to simulate large colonies and is optimised for rod shaped

58

bacterium, such as B. subtilis, making it well suited for our purposes. During each simulation

timestep, the cells grew, leading to overlaps. CellEngine resolved these overlaps using rigid-

body dynamics in two steps: collision detection and collision response162. The collision

detection stage identifies overlaps among bacteria and the collision response then

performs physical rearrangements163. Rigid-body dynamics were used for the computation

of linear and angular displacement of the bacteria. After initial overlap resolution there

were often other overlaps that then had to be resolved. Overlaps had to therefore be

resolved via an iterative process which displaces the overlaps outward of the colony. This

is a many-body problem, O(N2), where N is the number of bodies involved.

To compute a solution to this problem, in a way that is computational efficient, the

extended version of Gro implemented two assumptions. The first assumption was that the

position (location and orientation) of a cell depends mainly on the neighbouring cells. The

second assumption was that colonies grow out radially. These assumptions produced

approximate solutions based on local forces exerted by nearby bacteria and global forces

that act to push the bacteria outward from the colony. Nearby bacteria exert pressure

when growing and it was assumed that only the pressure from cells that lie within a distance

(k) from a specific cell needs to be considered (a local force approximation). The global force

was exerted radially from the centre of the colony and was proportional to the number of

cells located between the cell and the biofilm centre. The colony was split into radial rings

of one bacterium in width. These rings were grouped into sets of width (w) rings of a radius

k, which represent the length at which local forces are considered. The colony was

therefore transformed into circular subcolonies of width w. When the colony size increases,

so does the number of rings, but the width (w) of all the subcolonies doesn’t, so there was

a global solution of O(N).

The algorithm implemented in the extended version of Gro to overcome cell overlap

was executed at each timestep and involved two main stages: ring tagging and expansion.

59

The ring tagging stage starts from the colony edge and assigns each bacterium a ring. Ring

tagging is composed of two phases: edge detection and ring assignment. The expansion

phase was composed of two stages: relaxation and relocation. The approach implemented

by CellEngine to grow bacterial colonies works well for circular colonies, such as those

studied, but may not be suitable for modelling certain geometries (e.g. conjoined colonies

or a colony grown along a flat edge). The algorithms used for ring tagging and expansion

can be found in Gutiérrez et al. (2017)147.

Cell behaviour was controlled by a Probabilistic Timed Automata based library, CellPro,

which encapsulated and simulated gene expression. The release of potassium by a cell was

implemented by a set of rules, defined using CellPro, which simulated gene expression

using digital proteins with two possible states. When a protein was produced its value was

set to true and when it was absent/degraded it was set to false. Promoters, represented as

boolean functions (YES, NOT, AND, OR) regulated the protein-associated gene. These were,

in turn, regulated by transcription factors. At every time step during the simulation, every

bacterium had a state which was given by all its protein expression values. In our ABFDF

model, CellPro was used to trigger the release of potassium in response to a threshold

quantity of extracellular potassium.

The potassium propagation was controlled by CellSignal, using a finite element model.

At each time step of the simulation, diffusion and degradation were applied to update the

concentrations of the signals over a set of predefined grids. Algorithms 3.1 and 3.2 show

the algorithms and diffusion methods applied by CellSignal at each timestep. Further details

regarding the algorithms used in CellEngine, CellPro and CellSignal can be found in Gutiérrez

(2017)147.

60

Algorithm 3.1 Pseudocode of the main execution cycle of CellSignals147. This algorithm was

implemented at each simulation time step.

61

Algorithm 3.2 Pseudocode of the diffusion method implemented by the extended version of

Gro147. This subroutine applied diffusion to the signals in the grid (G), where r represents

each row and c each column. Diffusion was calculated according to a set of coefficients.

Custom values for M could be set in CellSignals. The subroutine updated the concentration

values of each cell after applying the finite element diffusion method.

The ABFDF simulations were performed using experimentally relevant parameters. The

two-dimensional nature of the simulations meant that there was no height in the grids used

to obtain the signal concentrations, to obtain real units a constant of proportionality is

therefore required. Table 3.II shows the parameters used in our ABFDF model. Figure 3.4

shows the cell density of a simulated biofilm, which was matched to the experimental cell

densities as closely as possible.

Table 3.II. Parameters used for agent-based model of electrical signalling in a biofilm in Gro.

Parameter Value

Simulation time conversion factor

3.5 min

Simulation time step (dt) 0.35 min

62

Pixel size 0.1 µm

Signal grid length 20 px

Signal grid cell size

Cell growth rate

Average cell division size

20 px2

0.034 fl/min

3.14 ± 0.071 fL

Potassium diffusion coefficient

0.4 (molecules·cellgrid)/dt

Potassium degradation coefficient 0.07 (molecules)/dt

Figure 3.4. Normalised cell density as a function of radial distance from the biofilm centre

for our experimental centrifugal wavefront data (red), centripetal wavefront data (black)

and agent-based fire-diffuse-fire model (blue). The centripetal biofilm had a larger radius

(~150 m) than the centrifugal biofilm (~90 m).

63

3.4 Results

3.4.1 Electrical signalling in circular B. subtilis biofilms (experimental

results and characterisation)

Figure 3.5. Electrical signal propagation through a two-dimensional biofilm. Schematics

show the spread of (a) centrifugal (‘away from the centre’) and (b) centripetal (‘towards the

centre’) electrical wave fronts through a biofilm. (c) The electrical signal given by ThT

64

fluorescence as a function of time at five different biofilm radii (r = 2 µm, 10 µm, 15 µm, 100

µm and 150 µm) from fluorescence microscopy experiments.

Following the work of Prindle (2015)85, B. subtilis circular biofilms were grown,

under flow, in the CellASIC ONIX microfluidic system and the membrane potential was

monitored using the membrane potential indicator Thioflavin-T (ThT) (see Section 3.3). In

addition to the previously described outward moving (centrifugal, Figure 3.6 (a)) electrical

waves, inward moving (centripetal, Figure 3.6(b)) electrical waves were observed. Both

signals were of comparable length, with similar profiles.

The circular nature of the biofilms motivated Prindle to use polar coordinates to

describe the electrical signals. Figure 3.3 is a representative ThT fluorescence profile as a

function of time and shows the signals at all angles in a single radius and the average signal

obtained by averaging these signals. The coefficient of variation of both signals’ radial mean

was always less than 0.07. It was therefore assumed that the average radial signal was

representative of the whole radial signal and so the ThT profiles were described in terms of

radius and time.

Figure 3.5 (c) shows ThT as a function of time observed at different radii for a typical

centrifugal electrical wave travelling outwards through a biofilm. To quantify such profiles

Prindle defined their half maximal position (in time) and their amplitude. I more accurately

quantified these profiles via moments’ analysis. Standard parameters were derived based

on the first 4 moments of the distributions i.e. the mean, the standard deviation, the

kurtosis and the skewness (Equations 3.1 – 3.4). The 𝑛th moment of a distribution 𝑓(𝑥) with

𝑁 points is given by

< 𝑥𝑛 > = ∑𝑥𝑖

𝑛

𝑁

𝑖=1

𝑓(𝑥).

(3.1)

65

The first quantity used to describe the radial signals was the mean (𝜇 ≡ 𝐸[𝑋]) , which is

given by the first moment. The standard deviation is given by the square root of the

variance, which is given by the second central moment

The third and fourth moments of a distribution can be used to calculate the skewness and

the kurtosis, which are measures of the distributions’ shape. The skewness is a measure of

the distribution’s symmetry and the kurtosis is an indicator of whether it is peaked and has

a heavy tail. The skewness is defined as

and the kurtosis is defined as

In addition to these four quantities, the fluorescent energy density, which is the

area under the ThT curves, was obtained. Figure 3.6 shows the membrane potential profile

as a function of time for a centrifugal wavefront (Figure 3.6(a)) and a centripetal wavefront

(Figure 3.6(b)). Figure 3.6 also shows the corresponding means (Figure 3.6(d)) and energy

densities (Figure 3.6(c)) as a function of radial distance. Biofilms grown in the microfluidics

system grew out from a cell trap, which led to the small gaps in the data seen in Figure 3.6.

The fluorescent energy density (𝐸𝑟(𝑟)) of both centrifugal and centripetal waves decreased

sigmoidally with radial distance from the biofilm centre,

where 𝑟0 is the half radial constant and 𝑥 is the slope constant, which describes the

steepness of the curve.

𝜎 = √𝐸[(𝑥 − 𝜇)2]. (3.2)

𝑆 = 𝐸[(𝑋 − 𝜇

𝜎)3]

(3.3)

𝐾 = 𝐸 [(

𝑋− 𝜇

𝜎)4

]. (3.4)

𝐸𝑟(𝑟) =1

1 + 𝑒(𝑟−𝑟0)/𝑥

(3.5)

66

Figure 3.6. Propagation of centrifugal and centripetal electrical signals through B. subtilis

biofilms. (a) and (b) ThT fluorescence intensity as a function of time and radial distance for

a biofilm in which an electrical signal has originated from (a) the biofilm centre (centrifugal)

and (b) the biofilm edge (centripetal). (c) The signals’ fluorescence energy density as a

function of radial distance for the centrifugal wavefront (red) shown in (a) and for the

centripetal wavefront (black) shown in (b), fitted with sigmoids (Equation 3.5). (d) Radial

distance for the maximum intensity as a function of signal mean time for the centrifugal

wavefront (red) shown in (a) and the centripetal wavefront (black) shown in (b), fitted with

power laws (Equation 3.7).

The average skewness of the centrifugal wavefront was -0.03 ± 0.18 and the

centripetal skewness was -0.32 ± 0.15. The skewness of the centrifugal wave decreased as

it travelled, whereas the skewness of the centripetal wave increased as it travelled (Figure

3.11(a)).

67

The centrifugal kurtosis was 3.78 ± 0.28 and the centripetal kurtosis was 2.29 ±

0.17. The centrifugal wave was leptokurtic (kurtosis > 3, more peaked than a Gaussian),

whereas the centripetal wave was platykurtic (kurtosis < 3, less peaked than a Gaussian).

Both signals were therefore not perfect Gaussians. However, after the data was smoothed

using moving averages, Kolmogorov–Smirnov tests were performed at the 5 % significance

level to confirm that the data was well approximated by Gaussians of the form

𝐹𝑟(𝑡) =

𝑎𝑟𝑒−(𝑡−𝑏𝑟)

2

𝑐𝑟

(3.6)

where 𝐹𝑟(𝑡) is the average ThT fluorescence as a function of time at the radius 𝑟, 𝑎𝑟 is the

average peak fluorescence of the signal at radius 𝑟, 𝑏𝑟 is the average time of the peak

amplitude of the signal at 𝑟 and 𝑐𝑟 is a constant related to the standard deviation giving the

width of the signal at 𝑟.

Gaussian fits were used to robustly obtain the signal amplitude as a function of

radial distance (𝑎𝑟). As expected, the signal amplitude followed the same profile as the

fluorescence energy density and decreased sigmoidally with radial distance from the

biofilm centre (Equation 3.5).

As previously mentioned, the signals’ mean was used to quantify when the signal

maximum reached a specific radial distance. Figure 3.6(d) shows the propagation of the

signal’s mean through the biofilm (velocity profiles of the mean). These velocity profiles

followed a power law dependence of distance on time,

where 𝐷 is the radial distance the signal has travelled, 𝜏 is the mean time (𝜇), which is the

time at which the signal maximum is reached at a specific 𝐷, and 𝐴0 and 𝛼 are both

constants.

𝐷 = 𝐴0𝜏𝛼, (3.7)

68

The centripetal wave had a larger exponent ( = 1.79 ± 0.03) than the centrifugal

wave ( = 1.42 ± 0.06). The centripetal wave, therefore, had a steeper velocity profile than

the centrifugal wave, which indicates that it travelled faster. The exponent of both the

centripetal and centrifugal wavefronts was significantly larger than 1 ( >1). This indicates

that these signals did not propagate at a constant velocity as previously described by Prindle

(2015). This apparent contradiction is caused by, the previously discussed, differences in

data analysis. Reanalysis of the Prindle (2015) data showed that their signals also did not

propagate at a constant velocity through the biofilm. It is widely recognised that curvature

can affect the propagation of a wavefront through an excitable medium124,164,165. For a

centripetal wavefront, different parts of the front will excite the same point concentrating

its activity, whereas, for a centrifugal wavefront, which expands as it travels, the energy to

excite its neighbours is spread out.

3.4.2 Modelling the propagation of electrical signals in circular biofilms

Fire-diffuse-fire model Calcium plays a crucial role in a broad range of cellular functions. For example,

calcium is central in muscle mechanics166 and cardiac electrophysiology167. Intracellular

calcium dynamics show highly complex spatiotemporal behaviour. A broad array of models

have been developed to explain this behaviour including fire-diffuse-fire (FDF) models146. I

used a similar model to describe potassium propagation in a biofilm. This model was built

on the assumption that once a threshold concentration of potassium (𝑘∗) was reached at a

single cell that the cell fired instantaneously releasing a fixed concentration of potassium

(𝜎). A potassium wave was propagated by the sequential firing of biofilm cells. In this FDF

model the potassium profile (𝐾 = 𝐾(𝑥, 𝑡), where 𝑥 is the position and 𝑡 is the time) obeys the

reaction-diffusion equation,

69

𝜕𝐾

𝜕𝑡= 𝐷𝑘

𝜕2𝐾

𝜕𝑥2− 𝛾𝐾 + 𝜎∑𝛿(𝑥 − 𝑖𝐿)𝛿(𝑡 − 𝑡𝑖

𝑖

)

(3.8)

where 𝐷𝑘 is the potassium biofilm diffusion coefficient, 𝑖𝐿 is the location of the ith cell, 𝑡𝑖 is

the time the threshold value of potassium (𝑘∗) is reached at the ith cell and 𝛾 is the

potassium decay rate.

Equation 3.8 has an exact solution in one dimension given by

𝐾(𝑥, 𝑡) =∑𝐾𝑖(𝑥, 𝑡)

𝑖

𝐾𝑖(𝑥, 𝑡) =𝜎𝐻(𝑡 − 𝑡𝑖)

√4𝜋𝐷(𝑡 − 𝑡𝑖)exp (−

(𝑥 − 𝑖𝐿)2

4𝐷(𝑡 − 𝑡𝑖) − 𝛾(𝑡 − 𝑡𝑖)))

(3.9)

where 𝐾𝑖(𝑥, 𝑡) is the potassium profile of a single cell and 𝐻 is the Heaviside function.

This simple FDF model was used to model the electrical waves described by Prindle

(2015) which propagates at a constant velocity through a biofilm85. To obtain solutions for

a wave which propagates steadily through the biofilm, the time between firing of

consecutive cells must be constant (𝑡𝑖 − 𝑡𝑖−1 = 𝜏) for all cells. If such a 𝜏 exists, then it is a

solution of the equation146

𝑘∗𝐿

𝜎=∑

1

√4𝜋𝑛𝜈

𝑛=1

exp (−𝑛

4𝜈− 𝛽2𝑛) ≡ 𝑔(𝜈)

(3.10)

where 𝜈 =𝐷𝜏

𝐿2 and 𝛽2 =

𝛾𝐿2

𝐷.

To model the results of Prindle (2015) their propagation timescale (𝜏) of 0.84 s was used

and it was assumed that the cell separation (𝐿) was 1 µm and that the potassium diffusion

coefficient in the biofilm was 70 % of its aqueous value (𝐷𝑘 is 1380 µm s-1)168. To determine

if constantly propagating solutions exist, 𝑔( 𝜈) (Equation 3.10) was plotted for a range of

different experimentally relevant parameters (Figure 3.7(a)). If 𝑘∗𝐿

𝜎> 𝑔( 𝜈)𝑚𝑎𝑥 propagation

fails and so only solutions 0 < 𝑔( 𝜈) < 𝑔( 𝜈)𝑚𝑎𝑥 were relevant. Solutions existed for all

the tested 𝛾 values (Figure 3.7(a)), proving that the FDF model may be used to obtain a

70

constantly propagating electrical signal in the biofilm within experimentally relevant

parameters. Figure 3.7(b) shows a representative electrical signal produced by the FDF

model using the parameters: 𝜏=0.84 s, 𝐿=1 µm, 𝐷𝑘=1380 µm/s, 𝛽=0.05. Figure 3.7(c) and

Figure 3.7(d) show the amplitude and propagation of this signal through the biofilm

respectively.

Figure 3.7. Fire-diffuse-fire model of electrical signal propagation through a B. subtilis

biofilm (Equation 3.8). (a) A plot of 𝑔(𝜈) as a function of 𝜈 for a range of different potassium

decay rates (𝛾). 𝑔(𝜈) is a function which may be used to determine the model’s stability and

thus find constantly propagating solutions to the FDF model (Equation 3.10). (b) The

potassium signal produced by our FDF model of a biofilm (Equation 3.9). (c) The signal

amplitude of the potassium wave shown in (b). (d) The velocity profile (position of the signal

maximum as a function of time) of the signal shown in (b).

This FDF model is a simple, one-dimensional model, which is limited by the

assumptions that the cells positions are modelled as delta functions and that the firing is

71

instantaneous. Its simplicity and exact solution make it useful for understanding simple

signals. However, the centripetal and centrifugal signals I observed experimentally could

not be explained by such a model.

Agent-based model Our experimental results (Section 3.4) show that electrical signals do not always

propagate constantly though the biofilm as originally proposed by Prindle (2015) (Figure

3.6). Agent-based modelling (ABM)156 was used to accurately describe more complex

behaviour. ABMs allow complex global behaviours to be understood which emerge as a

result of interactions between individuals in a multicellular community e.g. accurate

placement of bacteria in a two-dimensional biofilm with varying density.

A range of different software packages have been developed to simulate bacterial

colonies. The extended version of Gro147 was used because it is a simple and fast simulator,

which has been optimised for understanding the effects of colony spatial arrangement on

cell-cell communication. The extended version of Gro is based around five major

components:

1. A CellEngine is a physics engine which was developed to simulate large colonies and

is optimised for rod shaped bacterium, such as B. subtilis, making it well suited for

our purposes.

2. Prospect is an extension to CCL169, a guarded command based language developed

for modelling cooperative systems.

72

3. CellPro is a Probabilistic Timed Automata based library that simulates gene

expression dynamics using digital proteins. These proteins are then used to drive

cell behaviour.

4. The additional libraries CellNutrient and CellSignals can be used to control the

external environment. Signalling is implemented through a set of grids that store

the signal concentration at each grid location. At each time step of the simulation,

diffusion and degradation are applied to update the concentrations of the signals

over the whole grid using a finite element model and the cell states are then

updated.

5. Gro is the central subsystem of the simulator.

An agent-based model based on the FDF (ABFDF) model was built using the extended

version of Gro. Figure 3.8 shows the workflow executed for each timestep of our

simulations. In these simulations, potassium release was triggered either at the central cell

(centrifugal) or at an edge ring of cells (centripetal). The potassium wave was then actively

propagated through the biofilm, via the triggering of potassium release at other cells.

Following the FDF model, potassium release was triggered at a cell when a threshold

concentration of potassium was reached at that cell. To keep the model simple, it was

assumed that the potassium released by a single cell (𝑘𝑗(𝑡)) was described by a rectangular

function,

𝑘𝑗(𝑡) =

{

0 휀 0

𝑖𝑓 𝑡 < 𝑡𝑗

𝑖𝑓 𝑡𝑗 ≤ 𝑡 ≤ 𝑡𝑗 + 𝑡𝑟

𝑖𝑓 𝑡 > 𝑡𝑗 + 𝑡𝑟

(3.11)

73

where 휀 is the amplitude of the potassium concentration, 𝑡𝑗 is the time of release at cell j

and 𝑡𝑟 is the rise time, which is how long each cell fires potassium.

Figure 3.8. Workflow showing the steps executed by our model per time step (Δt). Firstly,

CellSignal was used to update diffusing signalling molecules. Secondly, the cell states were

updated for each cell in the simulation. Finally, CellEngine was used to grow the whole

colony.

74

Figure 3.9. Snapshots from a Gro simulation of our agent-based fire-diffuse-fire model of a

two-dimensional circular B. subtilis biofilms shown in (a) three-dimensions and (b) two-

dimensions. Snapshots are shown for time since initial firing at the centre of the biofilm T=0,

17, 63 and 120 mins. (c) A magnified image of a potassium wave spreading out from the

centre of the biofilm simulated by our agent-based fire-diffuse-fire model where the

bacterial agents are clearly visible.

75

Figure 3.9 shows a Gro simulation of our ABFDF model. The bacterial density profile

of the simulated biofilm (i.e. the areal concentration as a function of radial distance) was

matched with the experimental data (Figure 3.4). It was found that this was crucial in

ensuring that the propagation of electrical waves was accurately simulated.

Figure 3.10 shows the potassium profile produced by our Gro simulation for (a) a

centrifugal wavefront and (b) a centripetal wavefront. The fluorescent energy density and

amplitude of the centripetal and centrifugal model signals decreased sigmoidally with radial

distance from the biofilm centre (Figure 3.10(c)), in agreement with experimental data

(Figure 3.6(c)). However, the signals produced by our model did not decrease as rapidly

with radial distance as the experimental signals (Table 3.III).

The average kurtosis of the model centrifugal wavefront was 3.812 ± 0.004, which

matches our experimental results (3.78 ± 0.28). The average kurtosis of the model

centripetal signal was 3.21 ± 0.01, which, in agreement with our experimental results, was

lower than the average kurtosis of the centrifugal signal, but was still larger than our

experimental results (2.29 ± 0.17).

The centrifugal model wavefront had an average skewness of 0.548 ± 0.004 and the

model centripetal wavefront had an average skewness of 0.719 ± 0.001. Our model

wavefronts, therefore, have profiles with tails weighted towards the right, unlike our

experimental results, which had negligible skewness.

Closer examination of the changes in skewness and kurtosis observed with

propagation through the biofilm indicated that the differences between our model and

experimental results may be caused by differences in the size of biofilms and due to the

gaps in the experimental data caused by imaging problem at the cell trap. Figure 3.11 shows

the kurtosis and skewness of the experimental results ((a) and (b)) and model results ((c)

and (d)) as a function of radial distance from the biofilm centre. For both the model and

76

experimental wavefronts the skewness of the centrifugal wavefront was larger than the

centripetal wavefront, with a decrease in the skewness observed with the direction of the

wavefront. The kurtosis of the centrifugal wavefront was also larger than the centripetal

wavefront. The underlying trends across the biofilm were therefore comparable between

the model and experimental data.

Figure 3.10. Propagation of (a) centripetal and (b) centrifugal electrical waves produced by

our agent-based fire-diffuse-fire model. The potassium profiles were produced by our model

for a signal triggered at (a) the biofilm centre and (b) the biofilm edge. (c) Fluorescence

energy density, as a function of radial distance, of the centripetal signal (red) and of the

centrifugal signal (black) fitted with sigmoids (Equation 3.5). (d) Radial distance for the

maximum intensity as a function of the signal mean time for the centripetal signal (red)

77

shown in (a) and the centrifugal signal (black) shown in (b), fitted with power laws (Equation

3.7). For (c) and (d) data was averaged over three separate simulations.

Figure 3.11. Kurtosis and skewness of the electrical signal as a function of radial distance.

(a) Kurtosis of our experimental centrifugal wavefront (red) and centripetal wavefront

(black). (b) Skewness of our experimental centrifugal wavefront (red) and centripetal

wavefront (black). (c) Kurtosis of our ABFDF model’s centrifugal wavefront (red) and

centripetal wavefront (black). (d) Skewness of our ABFDF model’s centrifugal wavefront

(red) and centripetal wavefront (black).

78

Table 3.III. Fit constants of Equation 3.5 and 3.7 for our experimental and model

wavefronts.

Fit constant name

Fit constant symbol

Experimental centrifugal wavefront

Experimental centripetal wavefront

Model centrifugal wavefront

Model centripetal wavefront

Fluorescent energy

densities half radial

constant (a.u)

𝑟0 41.1 ± 0.5 44.7 ± 0.4 63.4 ± 0.1 68.4 ± 0.1

Fluorescent energy

densities slope

constant (a. u)

𝑥 7.5 ± 0.5 24.2 ± 0.4 13.5 ± 0.1 9.8 ± 0.1

Power law constant for

waves velocity profile

(µm/min)

𝐴0 0.081 ± 0.021

0.026 ± 0.003

2.441 ± 0.001

0.131 ± 0.006

Exponent constant for

waves velocity

profile (a. u)

𝛼 1.42 ± 0.06 1.79 ± 0.03 0.76 ± 0.01 1.38 ± 0.01

The velocity profiles (Figure 3.10(d)) were fitted with power laws (Equation 3.7). The

exponent () of the simulated centripetal wavefront was larger than the simulated

centrifugal wavefront, in agreement with our experimental results as well as the theory of

curvature effects on propagation165,170. The velocity of these waves was slow enough that

the colony can grow a significant amount during their propagation, it was therefore

important to use a model such as ours that can account for colony growth. However, our

model was not fully successful in describing the velocity of the signals. The exponents of

the power laws fitted to our model results were lower than those observed experimentally.

The simulated centripetal wave front had an exponent of α = 1.38 ± 0.01, whereas the

79

experimental exponent was α = 1.79 ± 0.03 (Table 3.III). The simulated centrifugal wave

front had an exponent of α = 0.76 ± 0.01, whereas the experimental wave front had an

exponent of α = 1.42 ± 0.06. It is likely that these differences were caused by the

complicated experimental set-up and by the model assumption that cells are randomly

arranged and do not grow in chains or other complex geometrical arrangements often

observed in biofilms. The cells in the outer biofilm were the least well represented by our

model as they were clustered and formed chains (Figure 3.2(c)).

In summary, our model demonstrates that the spatial arrangement of cells (the

curvature of the biofilm and variations in cell density) can explain most of the differences

in the propagation of the centrifugal and centripetal electrical wavefronts.

3.5 Discussion

Understanding the regulation and coordination of biofilm growth is crucial for the

development of treatments against this mode of growth. Universal mechanisms of

communication, such as electrical signalling, are especially attractive targets for treatment

development, as they can offer solutions which are widely applicable. Electrical signalling

could be used in applications which seek to harness biofilms, for example, in wastewater

treatments and fuel cells. In addition, it is hoped that the study of electrical signalling in

bacteria could inform our knowledge of electrical signalling in more complicated eukaryotic

systems e.g. brains, hearts and sensory organs.

Following the work of Prindle (2015), electrical signals in circular B. subtilis biofilms

were studied. In addition to the originally described centrifugal wavefronts, wavefronts

which originated at the biofilm edge i.e. centripetal wavefronts were observed. It remains

unclear what caused the centripetal waves that originated in the outer biofilm. However,

80

the similarities in the timescales of the centripetal and centrifugal signals, combined with

the effect of curvature, suggests that the signals were of a similar nature. Regardless of

whether these biofilms are caused by similar mechanisms, it is likely that electrical signalling

performs other roles in biofilm regulation. This is supported by the wide range of key

cellular processes which are influenced by membrane potential and also the broad range

of ion channels with unknown roles that are similar to eukaryotic cells with a range of gating

principles. I hypothesize that signals are triggered in response to a range of stimuli (e.g. a

variety of stress responses) and that these signals regulate a broad range of different

behaviours.

Moments analysis was used to fully quantify the electrical signals. This allowed us

to show that, contrary to previous beliefs, the electrical signals did not propagate constantly

through the biofilm. In contradiction to previous assertions, it was also found that the

fluorescence energy density and amplitude of wavefronts decreased with distance from the

biofilm centre. In addition, it was shown that, as previously described for other excitable

systems124,164,165, the centrifugal wave travelled slower than the centripetal wave,

demonstrating the effect of curvature on signal propagation.

Mathematical models can be used to further understand active excitable systems.

Agent-based modelling is uniquely placed to offer insights into the behaviour of complex

systems, such as biofilms, due to the ability to simulate global behaviour which emerges as

a result of interactions between multiple agents (in this case bacteria). An agent-based

model was built on a fire-diffuse-fire model and the assumption that individual cells

released potassium in response to a threshold concentration of extracellular potassium.

The potassium diffusion and degradation constants were homogenous across the biofilm

and the potassium release function of all cells in the biofilm was also constant (휀 in Equation

3.11). Our ABFDF model produced wavefronts which had characteristics matching our

81

experimental results. In support of our experimental findings, both the model centripetal

and centrifugal wavefronts were not propagated constantly through the biofilm. The

curvature affected the speed of the signal propagation, in line with our experimental

results, as well as theory165,170. The ABFDF model produced sigmoidal decreases in the

amplitude profiles, which again supported our experimental results. The changes in the

signal shape (quantified by the kurtosis and skewness) were also consistently produced by

the model. The ability of this model to successfully simulate these behaviours shows that

the spatial arrangement of the cells alone was enough to explain them. More subtle,

propagation characteristics (e.g. fractional power law velocity profiles) were not

quantitatively described by our model owing to the complex nature of biofilm growth.

Collectively these experimental and modelling results demonstrate how cell density and

curvature can influence the propagation of an electrical signal in a biofilm. More broadly

they show the power of detailed signal quantification and agent-based modelling in

interpreting electrical signals in biofilms.

Our ABFDF model can be adapted to describe different cell behaviours and

environmental conditions. It can, therefore, be used to model a wide variety of different

biofilms and signals. With the expected expansion of the field of bacterial

electrophysiology, such modelling techniques will become increasingly useful. An

improvement to our current model would involve adapting cell growth to better resemble

biofilm growth, e.g. the formation of chains of cells. A further addition to the current work

would involve studying electrical wave propagation in three-dimensional biofilms since

they are more commonly found in nature, although the interaction of electrical waves with

boundaries of the biofilm that have widely varying curvatures and experience varying time

delays is expected to markedly complicate matters.

82

3.6 Conclusions

Rigorous data analysis techniques and an ABFDF model were developed to describe

newly observed phenomena in the electrical signals of B. subtilis biofilms including the

effect of spatial heterogeneity in bacterial cell placements and curvature of propagating

wavefronts. Centripetal electrical wavefronts, which travelled inward from the edge of a

circular biofilm, were observed for the first time. These were quantified alongside

previously described centrifugal wavefronts using moments’ analysis. Agent-based

modelling was combined with a fire-diffuse-fire (FDF) model to realistically describe the

two-dimensional arrangements of bacteria in biofilms (beyond the results of simple analytic

one-dimensional models) and their emergent waves of electrical signalling activity. More

generally I have demonstrated the power of these methods in the emerging field of biofilm

active excitable matter.

83

CHAPTER

FOUR 4 Membrane potentials, oxidative stress and the dispersal

response of bacterial biofilms to 405 nm light treatment

4.1 Overview

Violet-blue 405 nm light has attracted increasing attention due to its intrinsic

antimicrobial effect, which bypasses the need for additional photosensitisers171,172. Despite

this, limited studies have been performed on the effect of 405 nm light on bacterial biofilms.

The response of P. aeruginosa and B. subtilis biofilms to photooxidative stress induced by

405 nm light was investigated. The dispersal and physical response of cells to 405 nm light

at different stages of biofilm growth were studied. Residence probabilities were used to

quantify the motile response of the bacteria to irradiation. The role of membrane potential

in the response of cells to 405 nm light was investigated using fluorescence microscopy and

a Hodgkin-Huxley style mathematical model.

84

4.2 Introduction

Biofilms are a prevalent and resilient form of bacterial growth. Biofilm growth

causes contamination of both biotic and abiotic surfaces. They are also associated with

difficult to treat chronic infections8. As described previously in the Introduction (Section

1.1.5) the protection afforded by biofilm growth is significant, leading to difficulties tackling

this form of growth.

Antimicrobial Photodynamic Therapy (PDT) has been extensively investigated as an

alternative technique for treating localized infections173 and as a decontamination method

for both industrial and clinical applications174. However, current PDT methods rely on the

use of chemicals and/or ultraviolet light. This causes issues with the sub-optimal uptake of

photosensitisers and a lack of selectivity for bacterial cells over host cells175.

Recent studies have shown that 405 nm light may provide a superior alternative to

current PDT methods, with a broad range of both, Gram-positive and Gram-negative

bacteria, being inactivated by 405 nm light176,177. The antimicrobial action of 405 nm light

does not require additional photosensitisers and studies have found that the dose levels

required to elicit a bactericidal response are not harmful to mammalian cells171. The

cytotoxic response of bacteria to 405 nm light involves the photoexcitation of intracellular

porphyrin molecules, which cause the generation of ROS (reactive oxygen species)171,178.

During photosensitization of the cell, the porphyrin molecules are converted to their triplet

state179, which then facilitates the production of ROS via either the Type I or the Type II

pathway. In the Type I pathway the porphyrin molecules react with the cellular

components, producing free radicals which then cause further reactions. In contrast, the

Type II mechanism involves the excited photosensitisers reacting directly with molecular O2

to produce singlet O2180. As a consequence, oxidative stress may lead to DNA or RNA

85

damage, lipid peroxidation, protein and nucleic acid oxidation, enzyme inhibition and the

activation of programmed cell death181–183.

Recent studies have shown that bacteria may regulate their membrane potentials

in response to stress85,184. This led us to investigate the role of membrane potential in the

oxidative response of bacterial biofilm cells to 405 nm light. To study contrasting responses,

a Gram-positive (P. aeruginosa) and Gram-negative (B. subtilis) were studied. As discussed

in the Introduction, both bacteria have been widely used to study biofilms. P. aeruginosa

was chosen for its clinical relevance and B. subtilis was chosen due to its reported resistance

to UV-C33, as well as its electrical activity.

4.3 Materials and Methods

4.3.1 Cell culture and growth

Experiments conducted on P. aeruginosa used the wild-type strain P. aeruginosa

PA01. The cells were freshly streaked onto TSA (Tryptic Soy Agar) plates and LB plates from

glycerol stocks two days before the experiment and incubated at 37oC overnight.

The next day, 10 ml of TSB (Tryptic Soy Broth) in a glass universal was inoculated

with a single colony. The inoculum was then incubated and shaken at 200 rpm at 37oC

overnight or until OD600 ≈ 2. The optical density of cells was measured using a

spectrophotometer. The cells were then diluted 1:2 with fresh TSB directly before injection

into the flow cell.

Experiments using B. subtilis were conducted on the strain NCIB 3610. The cells were

freshly streaked onto LB agar plates from glycerol stocks on the day before the experiment

and incubated at 37oC overnight. The next day, 3 ml of LB in a glass universal was inoculated

with a single colony. The inoculum was then incubated and shaken at 200 rpm at 37oC for

86

approximately 3 hrs or until the cells reached an OD600 ≈ 0.6. The cells were then centrifuged

at 2,100 rcf for 2 min and resuspended in a minimal MSgg medium to promote biofilm

growth. Media recipes can be found in Table 4.I.

Table 4.I. Recipes and sources for the culture media used in this chapter.

Media Recipe

LB 10 g/l NaCl,5g/l Yeast extract, 10 g/l Tryptone.

LB agar

10 g/l NaCl,5 g/l Yeast extract, 10 g/l Tryptone, 15 g/l agar.

TSB 30 g/l Tryptone soya broth.

TSA 30 g/l Tryptone soya broth, 15 g/l agar.

1 % TSB 10 g/l Tryptone soya broth.

PBS 8 g/l NaCl, 0.2 g/l KCl, 1.4 g/l Na2HPO4 and 0.24 g/l KH2PO4.

Msgg 5 mM potassium phosphate buffer (pH 7.0), 100mM MOPS buffer (pH 7.0 adjusted using NaOH), 2mM MgCl2, 700 µM CaCl2, 50 µM MnCl2, 100muM FeCl3, 1µM ZnCl2, 2µM thiamine HCl, 0.5% (v/v) glycerol and 0.5% (w/v) monosodium glutamate (Branda et al (2001)160).

4.3.2 Cell preparation for microscopy

Biofilm growth P. aeruginosa biofilms were grown in Ibidi µ-slide VI0.4 flow cells, with rectangular

size chambers of size 17 x 3.8 mm and height 0.4 mm (Figure 4.1(a)). This size allowed high

shear stresses to be exerted on cells, while still allowing plenty of space for biofilm growth.

87

The flow chamber was primed with media prior to the experiment, which was conducted

at room temperature. Cells were given an hour to adhere to the plates’ surface before the

flow of 1 % TSB was initiated at 30 µl/min.

B. subtilis biofilms were grown in Ibidi µ-slide III perfusion flow cell slides, which are

made of three channels, each containing two wells (Figure 4.1(b)). Experiments were

conducted in the first of these two wells. The wells were 5.5 mm in diameter and 1.2 mm

deep. The use of the wells reduced shear stress. The flow chamber was primed with media

prior to the experiment, which was conducted at 30oC. Cells were given an hour and a half

to adhere before the flow was initiated at 10 µl/min.

Figure 4.1. Schematic showing the ibidi flow cells in which biofilms were grown. (a) Ibidi µ-

slide VI0.4 with six identical channels in which P. aeruginosa biofilms were grown. (b) Ibidi µ-

slide III perfusion flow cell slides with three identical channels in which B. subtilis biofilms

were grown.

The media was supplied by the NE-1002X programmable microfluidics syringe

pump. Different culture protocols and flow cells were used for B. subtilis and P. aeruginosa

88

to optimise their growth. Different flow cells were used because B. subtilis requires lower

shear stress to form biofilms than P. aeruginosa. Data from different stages of biofilm

growth was drawn from a minimum of three independent experiments. Experiments to

compare different stages of biofilm growth were conducted in separate chambers to ensure

previous treatments did not affect the results.

Preparation of agarose microscope slides Following overnight culture, as previously described, fresh 1 % TSB medium (P.

aeruginosa) or Msgg (B. subtilis) was reinoculated 1:10 with the overnight culture. This

fresh culture was grown and shaken at 200 rpm, at 37°C, until the cells reached mid-

exponential phase (OD600 ≈ 0.5). This cell culture was then diluted to OD600 ≈ 0.05 using fresh

1 % TSB/ Msgg medium. Microscope slides were prepared an hour in advance of the cells

reaching mid-exponential phase, following an adapted version of the protocol of Jong et al.

(2011)105, which was developed for single cell imaging of B. subtilis. To make the microscope

slides, two glass microscope slides (Thermo Scientific; 76 x 26mm) were cleaned using 70

% ethanol and Milli-Q water, before a gene frame (ABgene; 17 x 28 mm) was attached to

the centre of one of the glass slides. To make the microscope slide agar, 1.5 % w/v high

resolution, low melting agarose (Sigma) was dissolved in 10 ml of 1 % TSB, before the

required antibiotics and/or H202 and/or ThT dye were added. While the agar medium was

still a liquid, 500 µl of it was pipetted into the centre of the gene frame, before the second

microscope slide was placed on top. The whole system was left to set at 4°C for 45 mins.

After it had set, the top microscope slide was slid off and two (or three) strips (≈5 mm wide)

were cut in the agar using a sterile razor blade. Two or three strips were cut to allow

experiments to be conducted in parallel, air gaps were left between the gene frame and

the strips to ensure cells had sufficient O2 (Figure 4.2). The cells were then loaded on to the

agar strips, 2.5 µl of cell culture was added to the top of the strip and dispersed over the

89

whole strip by gently turning the slide up and down. A microscope coverslip (Thermo

Scientific; 50 x 22 mm) was placed on top and adhered to the gene frame. The slides were

prewarmed for at least an hour prior to imaging to avoid issues arising from thermal drift.

Figure 4.2. Schematic showing a top and side view of the agarose microscope slide set-up

for fluorescence microscopy. Bacteria were immobilised between the agarose medium and

the microscope coverslip.

4.3.3 ROS scavengers

Sodium pyruvate and catalase stocks were prepared in advance. These stocks were

then added to the cell suspension and media half an hour before injection into the flow cell

at concentrations of 100 mM sodium pyruvate and 200 U/ml catalase. Experiments were

all conducted at the same time after initial inoculation to ensure biofilm growth did not

affect the results.

90

4.3.4 Microscopy

Microscope set-up Microscopy experiments were conducted using an Olympus IX-71 inverted

microscope with Olympus UPAON 100XOTIRFM (NA 1.49) immersion oil and a TIRF

objective lens.

Brightfield images were acquired using an LED (525 nm) which was focused onto

the sample using a condenser lens. For fluorescence imaging, the sample was illuminated

with an OBIS 405LX or OBIS 647LX laser. Time lapse images were recorded using an ORCA-

Flash4.0 LT PLUS Digital CMOS camera (C11440-22CU).

405 nm and 488 nm light treatment Cells were treated with OBIS 405LX and OBIS 488LX lasers. The laser power was

varied to provide the irradiance required at the sample. The power of the laser was set on

the PC using OBIS connection software and the laser power at the sample was measured

using a power meter. Brightfield microscopy images were obtained to track the number of

cells.

During experiments studying the effect of 488 nm light and during experiments

studying the effect of turning the 405 nm laser on and off, the lasers were controlled and

the time lapses collected using Micromanager.

Membrane potential indicators ThT is commonly used to stain amyloid fibres96, however, its positive charge allows

it to also be used as a Nernstian voltage indicator85–87. ThT was supplied by Sigma-Aldrich.

91

Fresh 2 mM stocks were made up on the day of the experiment and added to the cells and

media at a final working concentration of 10 µM. ThT can be excited at 405 nm, allowing it

to be used in conjunction with 405 nm light treatment. As shown in Figure 4.3 the addition

of 10 µM ThT did not impact the growth of P. aeruginosa PA01.

Figure 4.3. Growth curves (OD600) for P. aeruginosa grown in TSB media with and without

10 µM ThT.

DiSC3(5) was made up in 3 mM stocks in dimethyl sulfoxide prior to use and stored

at -20°C. DiSC3(5) was added to the cell suspension immediately before injection into the

flow cell at a final working concentration of 6 µM. P. aeruginosa cells were given 20 min, to

allow incubation and adherence, before time lapses were conducted. B. subtilis cells were

given an additional 20 min adherence period before media containing DiSC3(5) was added

at a final concentration of 9 µM. This was followed by a 20 min incubation period prior to

imaging. An OBIS 647LX laser at an irradiance of 1.8 mW/cm2 was used to excite DiSC3(5)

and time lapse images were recorded once every sec. In order to obtain a reasonable signal-

92

noise ratio the DiSC3(5) experiments had to be conducted in PBS. This allowed this dye to

be used to confirm hyperpolarisation results obtained using ThT, but prevented it being

used for long-term biofilm experiments. Cells were centrifuged for 10 min at 3,000 rcf and

then resuspended in PBS.

Irradiance/dose measurements The optical power was measured at the sample using a Thorlabs PM121D digital

power meter. The power across the area in which measurements were made was uniform,

so the irradiance was given by

𝐼 =𝑃

𝐴 ,

(4.1)

where 𝐼 is the irradiance, 𝑃 is the power at the sample and 𝐴 is the illumination area. The

dose (D) was given by

𝐷 =𝐼𝑡

1000

(4.2)

where is the D dose in J/cm2, I is the irradiance in mw/cm2 and t is the time of illumination

in s.

4.3.5 Data analysis

Image analysis was conducted using ImageJ (National Institutes of Health), Matlab

(MathWorks) and Origin (OriginLab Corporation). Background subtraction was done using

the ImageJ ‘rolling ball’ background plugin with a radius of 4 – 8 µm depending on the

experimental condition e.g. media. To obtain ThT curves the ‘plot Z-axis’ function on the

ImageJ image analysis toolbox was used. Signal processing was performed in Matlab,

including signal smoothing using a cubic weighted Savitzky-Golay filter, which removed

93

sinusoidal noise from the pump and instantaneous noise introduced when motile cells

passed behind the field of view. The errors presented are standard errors.

Prior to cell tracking, a bleach correction using a histogram matching method was

applied. This ensured that the increase in fluorescence with time did not affect the cell

tracking. Cells were tracked using the Fiji ImageJ plugin ‘TrackMate’ and tracks were then

exported to Matlab where further analysis was performed. This tracking was easier to

achieve than usual because it was only necessary to identify the number of cells per frame,

rather than needing to connect cells with their previous positions. Manual cell counting was

performed using the Fiji ImageJ plugin ‘Cell counter’, this method was used to confirm the

number of cells in the first and last frame of time lapses, it was also used to count the

number of cells before and after treatment with a single dose of light. Survival analysis (the

calculation of residence probabilities) was performed on the cell tracks using the Matlabs’

Statistics and Machine Learning Toolbox. Curve fitting and fit comparisons were performed

in Origin.

In order to compare different stages of biofilm growth five stages of biofilm growth

were defined based on previous literature (Figure 4.4)20,21,25,42,43:

Stage I - Initial attachment - Cells colonised the flow cell surface and reversibly

attached, often via their poles. This stage of biofilm growth was observed in the

first hour following inoculation of our system.

Stage II - Irreversible attachment - Following initial attachment, cells attached

irreversibly to the surface and began to form cell clusters. This occurred about 3 hrs

after inoculation of our system.

Stage III - Aggregation - Cells formed larger aggregates, this was associated with

cells becoming encased in an EPS. This was observed approximately 5 hrs after

initial inoculation of our experimental set-up.

94

Stage IV - Biofilm formation - The surface was densely populated by cells that were

part of large aggregates and biofilms, some cells had begun to attach to surface

cells to form a three-dimensional biofilm. This was seen approximately 10 hrs after

flow cell inoculation.

Stage V - Mature biofilm - Cells were part of a complex inhomogeneous biofilm,

containing channels and three-dimensional stacks. This stage of growth was

observed after at least 24 hrs of growth in our experimental set-up.

Figure 4.4. Representative images that depict P. aeruginosa cells stained with ThT at the

five stages of biofilm growth. (a) – (e) show representative cells at Stage I through to Stage

V.

4.3.6 Mathematical modelling

Our simple Hodgkin-Huxley155 model consisted of a set of coupled differential

equations (equations (4.14) to (4.21)). Following Prindle et al.85 and previous neuronal

models124,154,155, parameters were given representative values where possible, otherwise,

95

parameter scanning was implemented within an expected range (Table 4.II). Parameter

scanning, simulations, and plotting were all conducted using custom made Matlab scripts.

Table 4.II Fit parameters for our ROS membrane potential response model.

Fit parameter Original model Adapted model

𝑔𝑙𝑒𝑎𝑘 1 min-1 1 min-1

𝑔𝐼 50 min-1 50 min-1

𝑉𝑙𝑒𝑎𝑘 -156 mV -156 mV

𝑉𝐼 -300 mV -300 mV

𝑉0 -156 mV -156 mV

𝑆𝑡ℎ 2 2

𝛼0 2 min-1 0.5 min-1

𝛽0 0.009 min-1 -

𝜎𝑏

𝛽

200

- - 2

𝜎𝑠 2 2

𝐼𝑡ℎ

𝑎𝑡

𝛾𝑡

50 µW/cm2

0.03 µM/(min mV)

4 min-1

50 µW/cm2

0.03 µM/(min mV)

4 min-1

4.4 Results

4.4.1 Physical response of P. aeruginosa cells to 405 nm light treatment

To examine how localised 405 nm light affects P. aeruginosa cells, biofilms were

grown, under flow, in a microfluidic chamber and a finely focused laser was used to provoke

a response (see Section 4.3 for more details).

96

Figure 4.5. Phase contrast images depict the transformational change seen in P. aeruginosa

cells at Stage III of biofilm growth before and after a dose of 1.8 J/ cm2 of 405 nm light.

The physical response of cells to 405 nm light depended on the stage of biofilm

growth. Figure 4.7 shows the cell dispersal dose response to 120 ± 4 µW/cm2 of 405 nm

light, at the five stages of biofilm growth. Cell dispersal was specifically restricted to 405

nm light with no equivalent response observed to an equivalent dose of 488 nm wavelength

light (Figure 4.6). During Stage I and II of biofilm growth the cells left the surface in response

to 405 nm laser illumination at 120 ± 4 µW/cm2. Dispersed cells maintained a membrane

potential demonstrating their viability. At Stage III of biofilm growth, most cells did not

leave the surface in response to the laser. Initially, 10 % of the cells left the surface, but

these cells were then replaced by other cells. After approximately 250 min of treatment at

an irradiance of 120 ± 4 µW/cm2, which is equivalent to a dose of 1.8 J/ cm2, these cells

became physically altered by the laser treatment. Cells morphologically transformed from

their traditional rod shapes to coccoids (Figure 4.5). Physical shortening of cells was first

observed after 100 min of treatment at an irradiance of 120 ± 4 µW/cm2, which is equivalent

to a dose of 0.72 J/cm2. Cells fully transformed from traditional rod shapes to coccoids after

250 min of treatment at an irradiance of 120 ± 4 µW/cm2 which is equivalent to a dose of

1.8 J/ cm2. A similar morphological transformation has been observed in P. aeruginosa in

97

response to antibiotic stress185. This stress response has been linked to increased antibiotic

resistance, due to a change in the cells’ metabolic activity.

Figure 4.6. Dispersal response of Stage I P. aeruginosa cells to treatment by 0.1 J/cm2 of 405

nm and 488 nm light. (a) Representative brightfield images show the number of cells before

and after treatment. (b) Graph showing the number of cells before and after treatment.

98

At Stage IV, some cells dispersed, but most cells remained on the surface following

treatment and were not visibly altered by treatment. At Stage V of biofilm growth, almost

all cells remained on the surface following treatment. Surface cells on the bottom layer of

the biofilm stacks all remained after 6 hrs of illumination at 120 ± 4 µW/cm2, which is

equivalent to a dose of 2.6 J/cm2. These cells did not visibly change in response to the light.

The dispersal of bacteria at Stages I, II and IV of biofilm growth was quantified using

survival functions186. A survival function is defined as the probability that an event has not

occurred by a certain time. In order to avoid confusion with cell death, the survival

functions depicted in Figure 4.7(a) were defined as the biofilm residence probabilities,

which give the probability that the cells remain on the surface at a given time. This statistical

tool had not to my knowledge been used to describe bacteria before, but it provided a

robust description of their behaviour.

The Kaplan-Meier estimator is used to predict the survival function from

experimental lifetime data187. An important advantage of the Kaplan–Meier method is that

it can account for data censoring, specifically right censoring, which can occur if, for

example, a cell is not tracked correctly. The biofilm residence probability was well

estimated by the Kaplan-Meier function (Figure 4.7(a)). The biofilm residence probability as

a function of time (𝑆(𝑡)) given by the Kaplan-Meier estimator was defined as

𝑆(𝑡) = ∏(1 −

𝑑𝑖𝑛𝑖

𝑖:𝑡𝑖≤𝑡

)

(4.3)

where 𝑡𝑖 is the time until at least one cell leaves the surface, 𝑑𝑖 is the number of events at

time 𝑡𝑖 and 𝑛𝑖 is the number that were known to remain on the surface or that have not

been censored by time 𝑡𝑖.

Hazard functions are used to determine which periods in time have the highest and

lowest chance of an event happening186,188,189. The hazard function ℎ(𝑡) gives the

99

instantaneous failure rate of an individual (rate of leaving in our case) conditioned on the

fact that the individual survived until a given time:

ℎ(𝑡) = lim∆𝑡→0

𝑃(𝑡 ≤ 𝑇 < 𝑡 + ∆𝑡 ↓ 𝑇 ≥ 𝑡)

∆𝑡

(4.4)

where 𝑃(𝑡 ≤ 𝑇 < 𝑡 + ∆𝑡 ↓ 𝑇 ≥ 𝑡) is the probability of a single cell leaving the surface

between t and 𝑡 + ∆𝑡.

Figure 4.7. (a) Biofilm residence probability as a function of time (or equivalently dose) of P.

aeruginosa biofilms, exposed to 120 ± 4 µW/cm2 405 nm light, for the five stages of biofilm

growth. Corresponding fits of the Kaplan-Meier estimator (𝑆(𝑡), equation (4.3)) shown as

pink dashed lines. (b) The hazard functions ℎ(𝑡) obtained from the Kaplan-Meier functions

(𝑆(𝑡)) shown in (a) at Stage I, II and IV of P. aeruginosa biofilm growth with corresponding

fits shown in red. (c) The cumulative hazard functions 𝐻(𝑡) obtained from the Kaplan-Meier

100

functions (𝑆(𝑡)) shown in (c) at Stage I, II and IV of P. aeruginosa biofilm growth with

corresponding fits shown in red. (d) The ratio between hazard constants 𝑎𝑡 and 𝑏𝑡 (equation

(4.9)) from hazard functions shown in (b) and (c) at Stages I and Stages II and IV of P.

aeruginosa biofilm growth i.e. growth phases with a significant dispersal of bacteria.

Averages were taken from at least 20 cells in the field of view.

The hazard function was calculated from the Kaplan-Meier function using the

relation

ℎ(𝑡) = −𝑑

𝑑𝑡(ln (𝑆(𝑡)) .

(4.5)

Figure 4.7(b) shows the hazard rates for the biofilm residence probabilities described by

𝑆(𝑡). The cumulative hazard rate (𝐻(𝑡)) was calculated using

𝐻(𝑡) = − ln(𝑆(𝑡)) .

(4.6)

Figure 4.7(b) and Figure 4.7(c) show the hazard functions and cumulative hazard functions

derived from the Kaplan-Meier estimate with corresponding fits. The hazard functions

(ℎ(𝑡)) increased linearly with time at Stages I, II and IV of biofilm grow and were well

described by

ℎ(𝑡) = −𝑎𝑡 + 2𝑏𝑡𝑡

(4.7)

where 𝑎𝑡 is the intercept constant and 𝑏𝑡 is the slope constant.

The cumulative hazard functions were defined as

𝐻(𝑡) = −𝑎𝑡𝑡 + 𝑏𝑡𝑡2.

(4.8)

Mathematically this defines a highly non-Markovian statistical process. The increase in the

hazard function with time implies an increase in the chance of a cell leaving the surface with

time. Increases in survival hazard functions are commonly observed in radiation damage of

tissue in medical physics, although their origin is contentious190. The ratio of two different

biofilm residence hazard rates (ℎ(𝑡)1 and ℎ(𝑡)2) was dependent on time,

101

ℎ(𝑡)1ℎ(𝑡)2

=𝑎𝑡1 + 2𝑏𝑡1𝑡

𝑎𝑡2 + 2𝑏𝑡2𝑡

(4.9)

where 𝑎𝑡1, 𝑏𝑡1 and 𝑎𝑡2, 𝑏𝑡2 are the constants of functions 1 and 2 respectively (equation

(4.8)).

The non-proportionality of the hazard functions meant that different functions

could not be directly compared using the hazard ratio as a measure of survival. Instead 𝑎𝑡

ratios were used to compare initial leaving rates, and 𝑏𝑡 ratios were used to compare

dispersal rates (Figure 4.7(d)).

The 𝑎𝑡 ratio was 18.0 ± 0.4 between Stage I and Stage II of biofilm growth and 0.186

± 0.009 between Stage I and Stage IV of biofilm growth. The 𝑏𝑡 ratio was 41.075 ± 0.269

between Stage I and Stage II and 0.186 ± 0.002 between Stage I and Stage IV. The increase

in the hazard function constants 𝑎𝑡 and 𝑏𝑡 from Stage I to Stage II of biofilm growth implies

a larger initial and faster overall event rate. In contrast the decrease in the hazard function

constants (𝑎𝑡 and 𝑏𝑡) from Stage I to Stage IV of biofilm growth implies a lower initial and

overall slower dispersal rate. Overall, these results show that initially cells became more

physically responsive to 405 nm light with biofilm growth, but that as the biofilm matured

this responsiveness decreased until, by the stage of mature biofilm growth, no obvious

physical response was observed.

4.4.2 Membrane potential changes for P. aeruginosa in response to 405

nm light stress

To probe the role of membrane potential in the response of bacteria to 405 nm

light the membrane potential changes were monitored using the membrane potential

indicator dye Thioflavin-T (ThT). ThT was chosen instead of 3, 3’-dipropylthiadicarbocyanine

iodide (DiSC3(5)), a dye commonly used to measure membrane potentials in bacteria,

102

because it is better suited to measuring membrane potentials within biofilms and is three

times more sensitive 85,86,191. DiSC3(5) has also been shown to inhibit bacterial growth and

so is inappropriate for long term measurements, whereas at a working concentration of 10

µM, ThT does not inhibit P. aeruginosa growth (Figure 4.3).

Figure 4.8. Average ThT fluorescence of Stage II P. aeruginosa cells irradiated with 120 ± 4

µW/cm2 405 nm light as a function of time (or equivalently dose).

Across a range of different environmental conditions and physiological states, the

response of biofilm cells to 405 nm light was accompanied by an increase in ThT

fluorescence. More ThT is retained in a cell as it becomes more negatively charged and

therefore the observed increases in ThT implies an opposite change in the cell’s membrane

potential. This behaviour was specifically restricted to 405 nm light with no membrane

potential changes observed in response to an equivalent dose of 488 nm light (Figure 4.9).

103

An example of a typical membrane potential hyperpolarisation of an aggregate of P.

aeruginosa cells is shown in Figure 4.8.

Figure 4.9. Average ThT intensity of Stage I P. aeruginosa cells as a function of time (or

equivalently dose) observed in response to 405 nm light and 488 nm light, at a constant

irradiance of 480 ± 6 µW/cm2.

To confirm our results, the membrane potential dose response was measured using

DiSC3(5). This produced equivalent results to ThT (Figure 4.10), confirming the results and

suitability of ThT as a membrane potential indicator.

104

Figure 4.10. Average DiSC3(5) intensity of Stage I (a) P. aeruginosa and (b) B. subtilis cells as

a function of time (or equivalently dose) observed in response to 200 ± 4 µW/cm2 405 nm

light (black) and in response to no treatment (red). The errors presented are standard errors.

105

Hyperpolarisations were observed in response to all irradiances in our

experimental range (70 - 740 µW/cm2). These hyperpolarisations can be described by

Boltzmann sigmoidal curves of the form,

𝑇ℎ𝑇(𝑡) = 𝑇ℎ𝑇0+(𝑇ℎ𝑇𝑚𝑎𝑥 − 𝑇ℎ𝑇0)

1 + 𝑒(𝑡−𝑡0)/𝑦

(4.10)

where 𝑇ℎ𝑇(𝑡) is the fluorescence intensity as a function of time, 𝑇ℎ𝑇0 is the bottom plateau

constant, 𝑇ℎ𝑇𝑚𝑎𝑥 is the top plateau constant, 𝑡0 is the half time constant and 𝑦 is the slope

constant, which describes the steepness of the curve. When cells were constantly exposed

to 405 nm light the dose received by cells was directly proportional to the exposure time

and the irradiance (see Section 4.3.5). The ThT dose response can therefore be described

by an equivalent version of equation 4.10,

𝑇ℎ𝑇(𝐷) = 𝑇ℎ𝑇0+(𝑇ℎ𝑇𝑚𝑎𝑥 − 𝑇ℎ𝑇0)

1 + 𝑒(𝐷−𝐷0)/𝑥

(4.11)

where 𝑇ℎ𝑇(𝐷) is the fluorescence intensity as a function of dose 𝐷, 𝑇ℎ𝑇0 is the bottom

plateau constant, 𝑇ℎ𝑇𝑚𝑎𝑥 is the top plateau constant, 𝐷0 is the half dose constant and 𝑥 is

the slope constant, which describes the steepness of the curve.

To ascertain whether the membrane potential dose response of cells was

dependent on the irradiance of 405 nm light the irradiance was varied from 70 µW/cm2 -

750 µW/cm2 and corresponding hyperpolarisations were quantified using their sigmoidal

fit parameters. The sigmoidal fit parameters which describe the ThT fluorescence as a

function of time (𝑡0 and 𝑦) both decreased exponentially as a function of irradiance. The

decrease in both parameters with laser irradiance was well described by mono-

exponentials. The half time constant (𝑡0), decreased faster with increasing irradiance, than

the slope constant (𝑦). The exponential constant describing the dependency of 𝑡0 on laser

irradiance was 1.51 ± 0.09 times larger than the exponential constant of 𝑦. The sigmoidal

slope constant which describes the ThT fluorescence as a function of dose, 𝐷0, was not

106

dependent on the irradiancy. In contrast the half dose constant, 𝑦, increased linearly with

increasing laser irradiance.

To compare the irradiance dependency of the membrane potential responses at

different stages of biofilm growth and at different irradiances, two methods were used: one

involved comparing responses in different regions of the same flow cells and the other

involved shifting the power during treatment. Using the second method the linear regimes

of the responses were fitted with:

𝑇ℎ𝑇(𝐷) = 𝑚(𝐼)𝐷 + 𝑐

(4.12)

where 𝑇ℎ𝑇(𝐷) is the linear part of the ThT dose (𝐷) response, 𝑚(𝐼) is the irradiant

dependant slope of response and c is the intercept constant.

The slope constant m is dependent on the irradiance according to

𝑚(𝐼) = 𝐴𝑒𝐼𝑑𝑘

(4.13)

where 𝐴 is the exponential maximal constant and dk is the exponential slope constant.

The exponential maximal constant (𝐴) was dependent on the stage of biofilm

growth, but dk was constant. This meant that the ratio between the steepness of the

response at different stages of biofilm growth was not dependent on the irradiance, which

allowed direct comparison between different stages of biofilm growth at a set irradiance.

Figure 4.11 shows five hyperpolarisation curves observed in response to 405 nm

light, at a constant irradiance of 120 ± 4 µW/cm2, at the five previously defined stages of

biofilm growth (Section 4.3). The only stage of biofilm growth which was not well described

by a single sigmoidal function (equation 4.11) was Stage III. During this stage of growth,

sigmoidal behaviour was followed by an exponential phase. In order to compare this stage

of biofilm growth with other stages, the initial behaviour was fitted according to equation

(4.11) (shown in Figure 4.11).

107

The parameters governing the sigmoidal fits of the membrane potential dose

response varied with the stage of biofilm growth (Figure 4.12). The slope constant (𝑥) is

inversely proportional to the steepness of the response. From Stages I through to III, 𝑥

decreased as the cells responded faster to the laser light. There was a 67 ± 2 % decrease in

𝑥 from Stage I to Stage II and a 96 ± 9 % decrease in 𝑥 from Stage I to Stage III.

The cells responded slowest to the light during the later stages of biofilm growth

(Stages IV and V). This caused an increase in 𝑥, which was 34 ± 2 % higher in Stage IV than

in Stage I and 291 ± 7 % times higher in Stage V than Stage I. The 𝐷0 value defines the

position of the half-maximal membrane potential and followed a similar pattern, with the

only difference being that it was higher during Stage IV than Stage V. Cells which remained

on the surface depolarised back to their original values following hyperpolarisation at

longer time scales (>500 min, Figure 4.11(b)) demonstrating they were still viable.

Although the dose response of cells depended on the laser irradiance, the change

in the steepness of the dose response due to biofilm growth was not dependent on the

irradiance used to evoke the response (Section 4.3.5). This meant that the steepness ratios

of different biofilm growth dose responses obtained at 120 ± 4 µW/cm2 were

representative of the ratios observed across the entire range of irradiances tested. The

membrane potential dose response was 2.91 ± 0.02 times less steep in the mature biofilm

than in the initially adhered cells.

108

Figure 4.11. (a) Average cell ThT intensity as a function of time (or equivalently dose)

observed in response to 405 nm light at different stages of P. aeruginosa biofilm growth, in

the same media, at a constant irradiance of 120 ± 4 µW/cm2 with corresponding sigmoidal

fits to equation (4.11). (b) Average ThT fluorescence of mature P. aeruginosa biofilm cells as

a function of time (or equivalently dose) in response to 405 nm light. Data was collected for

a much longer time than that shown in (a) i.e. 900 mins compared to 1500 secs.

109

Figure 4.12. Boltzmann sigmoidal fit parameters (half-maximal dose (D0) and slope constant

𝑥 as given by equation (4.11)) which define the average hyperpolarisation of P. aeruginosa

cells at the five stages of biofilm growth in response to 405 nm light at 120 ± 4 µW/cm2 of

405 nm light.

Figure 4.11 was obtained by averaging the ThT intensity of cells in the region of

treatment and so is representative of the global behaviour. Figure 4.13(a) shows the spatial

heterogeneity in cell membrane potential response of Stage I biofilm growth cells in

response to 120 ± 4 µW/cm2 405 nm light. The half-maximal membrane potential

(𝐷0) varied between 82 s and 422 s (presented here in units of time rather than dose to

allow comparison with the leaving time). The leaving time (time at which a cell leaves the

surface) also varied (Figure 4.13(c)). Similar stochasticity is used as a strategy in gene

regulation 192.

There was a positive correlation, at the 0.05 significance level, between individual

cell half maximum membrane potential (𝐷0) and leaving time (Figure 4.13(b)). The

110

difference in 𝐷0 and leaving time of different cells had no statistically significant correlation

with cell separation. This confirmed that this was not an example of a coordinated stress

response across multiple cells (Figure 4.13 (c) and Figure 4.13(d)).

Figure 4.13. (a) Individual cell ThT intensity as a function of time (or equivalently dose),

observed at Stage I of P. aeruginosa biofilm growth, in response to 120 ± 4 µW/cm2 405 nm

light. (b) Leaving time of individual cells from (a) as a function of half-maximal time, with

corresponding linear fit shown in red. Each data point is one cell. (c) The average difference

in the leaving time of cells as a function of cell separation. (d) The average difference in the

half-maximal time of cells as a function of cell separation.

111

4.4.3 The response of fixed cells

To test what happened when cells were not free to leave the surface, an agarose

microscope slide set-up was used (Section 4.3.2). Cells were spotted onto a semi-solid

agarose medium. If left, unirradiated, these cells grew and divided to form microcolonies.

When treated with 405 nm light, the trapped bacterial cells became deformed and lysed.

These cells hyperpolarised, as observed previously in the flow cell set-up, before

depolarising as they were damaged/killed by the 405 nm light (Figure 4.14). Differences

between the flow cell set-up and this set-up, such as flow, could affect the levels of ROS.

However, the rate of hyperpolarisation of the initial cells in this set-up was comparable to

the initially adhered cells in the flow cell set-up, which suggests that experimental

differences did not have a significant effect on the membrane potential response.

The response of these cells was superficially similar in shape (an increase followed

by a decrease) to mature biofilm cells in the flow cell (which also did not disperse; Figure

4.11(b)). However, the two responses differed significantly. The ThT profile of initial cells

was much faster (Figure 4.14), with a peak at 1.33 ± 0.02 min, whereas for mature biofilm

cells (Figure 4.11(b)) the peak was at 102 ± 2 min. Even despite possible experimental

differences, this suggests that the differences in the membrane potential response of

mature biofilm cells was not caused solely by their inability to leave. The mature biofilm

cells also returned to membrane potential values comparable with their starting rest

potential values, indicating viability after treatment, whereas the initially adhered cells

became damaged by treatment when they were not free to leave and depolarised to values

lower than their original resting potentials.

112

Figure 4.14. Average ThT fluorescence of trapped P. aeruginosa cells as a function of time

(or equivalently dose) in response to 405 nm light at a constant irradiance of 120 ± 4

µW/cm2.

4.4.4 Probing the dynamics and timescales of the hyperpolarisation

response

To test how cells recovered from irradiation and to probe the mechanisms

underlying the hyperpolarisation response the laser was turned on and off. Cells were

treated with the laser; the laser was then switched back off and cells were given time to

recover before the laser was switched back on again. If the initial treatment length was

over a critical dose, then the cells continued to respond, even while the laser was switched

off. The increase in membrane potential was slower when the laser was switched off and

therefore the ThT response as a function of time was slower, leading to an increase in the

sigmoidal constants t0 and 𝑦. In contrast, the ThT response as a function of dose was altered

in the other direction, leading to increased sigmoidal constants D0 and 𝑥. This is because

113

the ThT fluorescence of cells increased, while the dose was constant (while the laser was

off). Once a critical threshold dose was crossed, the time/dose at which the treatment was

paused did not change the results. For example, Figure 4.15 shows the dose response

observed when the laser was switched off for 1 min after 15 sec and after 1 min, both

responses are comparable.

Figure 4.15. Average ThT fluorescence of trapped P. aeruginosa cells as a function of dose

in response to 405 nm light at an irradiance of 120 ± 4 µW/cm2. The three different curves

represent the laser on constantly (black), the laser switched off for 1 min after 15 sec of

illumination (blue) followed by continuous irradiation, and the laser switched off for 1 min

after 1 min on illumination (red) followed by continuous irradiation.

Cells were also treated with a laser which was pulsed on and off. Figure 4.16 shows

the response when the laser was switched on for 10 ms every 0.1 min. The ThT fluorescence

response as a function of time was slower when the laser is pulsed rather than being on all

the time (t0 and 𝑦 are larger). However, the dose received by these cells was so small, that

114

it was difficult to display both responses as a function of dose on the same graph (Figure

4.16(b)). These results show that pulsing the laser significantly increases the speed of the

dose responses (D0 and 𝑥 are much smaller), meaning that a relatively small total dose of

405 nm light can cause a significant response.

Figure 4.16. Average ThT fluorescence of trapped P. aeruginosa cells as a function of (a)

time and (b) dose in response to 405 nm light at an irradiance of 120 ± 4 µW/cm2. The black

curves represent the laser being on constantly and the red curves represent the response

observed when the laser is turned on once every 0.1 min for 10 ms.

4.4.5 Response in the presence of scavengers

Figure 4.17 shows the response of P. aeruginosa to 405 nm light in the presence

and absence of a scavenger mix consisting of 100 mM sodium pyruvate and 200 U/ml

catalase. The addition of ROS scavengers altered the membrane potential and dispersal

response of P. aeruginosa cells to 405 nm light, confirming that both responses were

associated with ROS. The hyperpolarisation dose response was slower and delayed in the

presence of ROS scavengers. There was a 74.1 ± 0.2 % increase in the half dose constant

(𝐷0) as well as a 73.2 ± 0.8 % decrease in the steepness (1

𝑥) in the presence of scavengers.

115

The dispersal response was also significantly reduced by the ROS scavengers with a 46 ± 11

% reduction in the number of cells that dispersed in response to 85 mJ/cm2 of 405 nm light.

Figure 4.17. (a) Average cell ThT intensity as a function of time (or equivalently dose) of

Stage I P. aeruginosa cells with and without added scavengers (100 mM sodium pyruvate

116

and 200 U/ml catalase) exposed to 405 nm light at a constant irradiance of 120 ± 4 µW/cm2

with corresponding sigmoidal fits to equation (4.11) shown in blue. (b) Residence probability

(probability of surface cells remaining) of Stage I P. aeruginosa cells following 700 secs

(equivalent to 85 mJ/cm2) of 405 nm light treatment with and without added scavengers

(100 mM sodium pyruvate and 200 U/ml catalase). Errors bars show the standard error.

4.4.6 The response of B. subtilis to 405 nm light

The Gram-positive bacterium B. subtilis was also studied. Early stage B. subtilis

biofilm cells were seen to disperse or become deformed in response to 405 nm light.

Dispersed cells maintained a membrane potential, signifying viability. In contrast to the

dispersal of P. aeruginosa cells, which was gradual, the clearance of B. subtilis cells was

sudden. This was reflected in the cumulative hazard function, which was much sharper for

B. subtilis (Figure 4.18) than for P. aeruginosa (Figure 4.7). The cumulative hazard function

of B. subtilis was initially flat, followed by a sharp increase, which represents a sudden

increase in the chance of cells leaving the surface. This is suggestive of a dispersal

mechanism which is activated above a threshold concentration of ROS with a corresponding

lag time e.g. it could be due to a coherence feed forward loop gene circuit193. In contrast,

the P. aeruginosa hazard function increased from the outset of treatment and was less

steep, this suggests that different mechanisms may govern the dispersal response of these

two bacterial strains. However, due to differences in the experimental protocols used for

the two species, we cannot rule out that differences were a consequence of different

experimental conditions.

117

Figure 4.18. Biofilm residence probability as a function of time (or equivalently dose) for a

B. subtilis biofilm exposed to 120 ± 4 µW/cm2 of 405 nm light, the Kaplan-Meier estimate

(equation (4.3)) is shown in red, with an inset of the corresponding cumulative hazard

function (equation (4.8)). Averages were taken from at least 20 cells in the field of view.

The response of B. subtilis to 405 nm light treatment was also accompanied by

membrane potential hyperpolarisations (Figure 4.19). These hyperpolarisations followed

the same profile as P. aeruginosa and were well described by sigmoidal fits as defined in

equation (4.11).

The addition of ROS scavengers (100 mM sodium pyruvate and 200 U/ml of

catalase) also altered the hyperpolarisation response of B. subtilis (Figure 4.20). The half

dose constant (𝐷0) increased by 65.7 ± 0.9 %, analogous to P. aeruginosa, in contrast

118

however the steepness of the response (1

𝑥) was not greatly affected by the addition of

scavengers in B. subtilis with only a 6.1 ± 0.1 % decrease observed.

Our results confirm that B. subtilis responds to 405 nm light, despite its established

resistance to UV light33.

Figure 4.19. Average ThT intensity of Stage II B. subtilis cells as a function of time (or

equivalently dose) in response to 120 ± 4 µW/cm2 of 405 nm light.

119

Figure 4.20. Average cell ThT intensity as a function of time (or equivalently dose) of Stage

I B. subtilis cells with and without added scavengers (100 mM sodium pyruvate and 200

U/ml catalase) exposed to 405 nm light at a constant irradiance of 120 ± 4 µW/cm2 with

corresponding sigmoidal fits to equation (4.11) shown in blue.

4.4.7 Hodgkin-Huxley model for the stress response

It is known that the oxidative stress response of cells may be affected by changes

in the responsiveness of cells and by differences caused by surrounding cells and EPS 34,36,194.

A Hodgkin-Huxley style model155 was used to test the hypothesis that changes in the

production and loss of ROS at different stages of biofilm growth caused the differences in

the observed membrane potential dose responses. Our model was similar in style to that

of the Prindle (2015)85 model, which was developed to describe the response of potassium

ion channels in B. subtilis. The bacteria were modelled as excitable cells with a membrane

120

potential (𝑉) given by

𝑑𝑉

𝑑𝑡= −𝑔𝐼𝑛

4(𝑉 − 𝑉𝐼)− 𝑔𝑙𝑒𝑎𝑘(𝑉 − 𝑉𝑙𝑒𝑎𝑘)

(4.14)

where it was assumed that the observed behaviour is described by one representative ion

channel (I) and a leakage current (leak). 𝑔𝐼 and 𝑔𝑙𝑒𝑎𝑘 are the respective channel

conductances. 𝑉𝐼 and 𝑉𝑙𝑒𝑎𝑘 are the ion and Nernst potentials respectively. 𝑛 is the ion

activation constant and is given by,

𝑑𝑛

𝑑𝑡= 𝛼(𝑆)(1− 𝑛) − 𝛽(𝑉)𝑛

(4.15)

where the opening rate 𝛼(𝑆) was taken to be dependent on the stress induced by ROS

(𝑆) following a Hill equation with the cooperativity parameter (n) set equal to 1,

𝛼(𝑆) =𝛼0𝑆

𝑛

𝑆𝑡ℎ + 𝑆𝑛

(4.16)

where 𝛼0 is the maximal opening rate and 𝑆𝑡ℎ is the threshold stress value.

Following Hodgkin-Huxley’s original work on squid axons the ion channel closing

rate 𝛽(𝑉) was assumed to be dependent on the voltage,

𝛽(𝑉) = 𝛽0𝑒−𝑉/𝜎𝑏

(4.17)

where 𝛽0 is the maximal closing rate and 𝜎𝑏 is the coefficient of the closing rates’ voltage

dependency. Both P. aeruginosa and B. subtilis contain voltage-gated ion channels62,85,195.

It was assumed that the stress induced by ROS was proportional to the difference

in the production and loss of the ROS in the cell. Following dynamical equations of neuronal

excitability130,154,155 and the subsequent work of Prindle (2015)85 the production of ROS was

represented by a threshold-linear function of laser irradiance (𝐼)

𝑑𝑆

𝑑𝑡=

𝛼𝑠(𝐼𝑡ℎ − 𝐼)

exp (𝐼𝑡ℎ − 𝐼𝜎𝑆

− 1)− 𝛾𝑠𝑆

(4.18)

where 𝛼𝑠 is the production constant and 𝛾𝑠 is the decay constant, both of which were

121

assumed to depend on the stage of biofilm growth. In our experiment, the irradiance was

constant so that

𝑑𝐼

𝑑𝑡= 0 .

(4.19)

Following Prindle (2015) a simple dependence between ThT fluorescence (𝐹) and level of

membrane hyperpolarisation (𝑉0 − 𝑉) was assumed,

𝑑𝐹

𝑑𝑡= 𝛼𝑡(𝑉0 −𝑉) − 𝛾𝑡𝐹

(4.20)

where 𝑉0 is the resting membrane potential, 𝛼𝑡 is the constant of proportionality between

ThT fluorescence and the level of membrane potential hyperpolarisation and 𝛾𝑡 is ThT

decay rate. The parameters which defined this model are shown in Table 4.II, where

possible these were matched with previously known results, the rest were chosen by

parameter fitting.

To test our model, both the input irradiance (𝐼) and the parameters 𝛼𝑠 and 𝛾𝑠 were

varied. For input irradiances in the range 85 - 575 µW/cm2, our model produced ThT

fluorescence profiles with sigmoidal fit parameters which had the same irradiance

dependencies as our experimental results; i.e. 𝑡0and 𝑦 decreased exponentially as a

function of 𝐼, 𝑥 was not dependent on 𝐼, and 𝐷0 increased linearly with 𝐼 (Figure 4.22(a)

and Figure 4.22(b)).

To further test this model and its assumptions, experiments in which the laser was

switched on and off were simulated. If the laser was turned off after an initial threshold,

then as seen experimentally, it made no difference when the laser was switched off. The

change in the simulated ThT dose response and its corresponding fit parameters were

comparable (Figure 4.22(c)). Pulsing the laser caused a much faster ThT dose response than

when the laser was on constantly (Figure 4.16). Simulations using a pulsed laser (10 ms

every 0.1 min to match experiments), also produced a ThT dose response which was much

faster than the response to constant irradiation (Figure 4.21(b)). This was a consequence of

122

the slowdown in the response observed in time (Figure 4.21(a)) not making up for how

small the total dose received was. However, the differences in the response of a pulsed and

unpulsed laser predicted by our model were not as drastic as the differences observed

experimentally. In summary, these results show that our model can be used to show that if

the laser is turned off after a threshold dose, the response continues. These results also

demonstrate that the model’s assumptions and equations, such as the representation of

ROS production as a threshold-linear function of laser irradiance, are good approximations,

but in their current state they do not fully encapsulate the behaviour.

Figure 4.21. ThT fluorescence as a function of (a) time and (b) dose in response to 405 nm

light at an irradiance of 120 ± 4 µW/cm2 produced by our Hodgkin-Huxley style model. The

black curves represent the laser being on constantly and the red curves represent the

response observed when the laser is turned on for 10 ms once every 0.1 min.

It was assumed that larger ROS production rates (𝛼𝑠) were observed when cells

were more metabolically active and that 𝛾𝑠 increased with biofilm growth, so that from

Stage I to Stage V: 𝛼𝑠 = 0.001, 0.05, 4, 0.0008, 0.0005 µM/(min mV) and 𝛾𝑠= 0.001, 0.003,

0.005, 0.008, 0.1 min-1 respectively. Figure 4.22(d) shows that these assumptions may be

used to correctly model the differences in the membrane potential response at different

stages of biofilm growth.

123

Figure 4.22. (a) Three-dimensional hyperpolarisation curve shows the membrane potential

response to 405 nm light of biofilm cells predicted by our model for a range of input

irradiances (85 µW/cm2 - 575 µW/cm2) as a function of time. (b) ThT fluorescence as a

function of dose in response to 405 nm light at an irradiance of 85 µW/cm2 (black), 185

µW/cm2 (green), 285 µW/cm2 (red) and 385 µW/cm2 (blue) produced by our model. (c) ThT

fluorescence as a function of dose produced by our model in response to 405 nm light at an

irradiance of 120 µW/cm2 which is switched on and off. The black curve represents the dose

124

response simulated when the laser is on constantly, the blue curve represents the response

observed when the laser is switched on for 45 sec then off for 1 min then back on, and the

red curve represents the response observed when the laser is switch on for 1 min then off

for 1 min then back on. (d), (e) and (f) Hyperpolarisation curves show the membrane

potential with time (or equivalently dose) in response to 120 µW/cm2 of 405 nm light

predicted by our Hodgkin-Huxley style model with corresponding sigmoidal fits to equation

(4.11). (d) Shows simulations of the five stages of biofilm growth based on the assumption

that the rate of ROS production and decay were dependent on the metabolic state of cells

and the stage of biofilm growth. (e) and (f) show simulations of our original and adapted

model in the presence (black) and absence (red) of ROS scavengers, based on the

assumption that the addition of scavengers leads to an increase in the decay rate of ROS.

The differences in the changes to the membrane potential response caused by the

addition of ROS scavengers, combined with the differences in the cumulative hazard

function of P. aeruginosa and B. subtilis, suggest a different response mechanism to

photooxidative stress. To test this hypothesis, the original model was adapted so that rather

than a Hill dependency, the channel opening rate (𝛼(𝑆)) was assumed to depend on the

ROS stress following a unit step response,

𝛼(𝑆) = 𝛼0 𝜃(𝑆 − 𝑆𝑡ℎ)

(4.21)

where 𝜃 = 0 when 𝑆 < 𝑆𝑡ℎ , 𝜃 = 1 when 𝑆 ≥ 𝑆𝑡ℎ and the closing rate (𝛽) was assumed to

be constant. The lag time described by the step function represents a delay, for example

in the internal gene circuit of the bacterium, such as a feedforward loop193. The rest of the

model and its assumptions remained the same. The parameters for this adapted model are

shown in Table 4.II.

ROS scavengers increase the decay rate of ROS, this directly translates to an

increase in the decay constant (𝛾𝑠). Figure 4.22(e) shows the change in the membrane

potential dose response produced by an increase in the ROS decay rate in our original

125

model with corresponding sigmoidal fits (equation (4.11)). The model predicted an increase

in the half-maximal dose constant (𝐷0) and a corresponding decrease in the steepness of

the response (1

𝑥). This confirms that our original model and its’ assumptions may be used to

explain the membrane potential hyperpolarisations of P. aeruginosa (Figure 4.11 and Figure

4.17). Figure 4.22(f) shows the change in the membrane potential dose response predicted

by the adapted model due to the addition of ROS scavengers with corresponding sigmoidal

fits. This adapted model (including equation (4.21)) produced an increase in the half-

maximal dose constant, but with no corresponding decrease in the steepness of the

response, successfully describing the observed membrane potential dose response of B.

subtilis (Figure 4.20).

4.5 Discussion

The response of both Gram-positive and Gram-negative bacteria to 405 nm light

was accompanied by membrane potential hyperpolarisations. Hyperpolarisations were

observed across a range of biofilm growth states. This suggests that this behaviour may be

universal. The hyperpolarisation response was delayed in the presence of ROS scavengers.

This implies a link between the hyperpolarisations and the ROS generated by 405 nm light.

Both P. aeruginosa and B. subtilis dispersed in response to 405 nm light. Hazard

functions were used to reveal differences in the dispersal of the two species. The dispersal

of B. subtilis was sudden, after a threshold dose of 405 nm light was reached, cells were

very likely to leave the surface (Figure 4.18). Whereas, the dispersal of P. aeruginosa was

gradual (Figure 4.7). There were also differences in the hyperpolarisation changes caused

by the addition of ROS scavengers. The hyperpolarisation half dose constant increased in

both species, but a large decrease in the steepness of the response was only observed in P.

126

aeruginosa. Differences in the experimental protocols used for the two species prevented

direct comparison, however, a Hodgkin-Huxley model was used to demonstrate that the

photooxidative stress dependency may explain these differences.

Recent studies have revealed that not just phototrophs respond to changes in light.

A range of photophysiological responses have been observed in range of different bacteria.

Photoreceptors have been shown to regulate the transition from the motile state to the

biofilm state and back again by a range of mechanisms196. Studies have revealed that 405

nm light induces photophysiological responses in a range of bacteria via different blue-light

receptor classes (LOV, BLUF and PYP)197,198. For example, 405 nm light has been shown to

induce responses, such as biofilm formation and motility in Acinetobacter baumannii via

the blue light using flavin (BLUF) protein BlsA199. Blue light was shown to activate σB and

the general stress of B. subtilis200, via the LOV protein, YtvA201. Providing a possible

mechanism via which B. subtilis responds to blue light.

It has recently emerged that P. aeruginosa responds to light via the photoreceptor

BphP202. It was found that biofilm formation and virulence were regulated by the

phosphorylation and activation of AlgB, by BphP, in response to far-red light. The

photosensory proteins LOV-HK and BphP1 form an integrated network that regulates

swarming motility in Pseudomonas syringae in response to multiple light wavelengths203. In

contrast the motility of Synechocystis sp. PCC 6803 is affected by 405 nm light via an cyclic

diguanylate (c-di-GMP) signal transduction system204. As c-di-GMP/cAMP levels have been

shown to regulate the stress response of P. aeruginosa205, I suggest that a similar

mechanism may be responsible for the observed 405 nm light induced dispersal.

The photophysical and the membrane potential response to 405 nm light was

dependant on biofilm growth. Intermediate stage biofilm cells (Stage II and Stage III)

showed the fastest and most significant physical changes in response to 405 nm light. These

127

cells also exhibited the fastest membrane potential changes. Mature biofilm cells (Stage V)

showed no physical response to 405 nm light and exhibited the slowest membrane

potential changes. This correlation between the physical and the membrane potential

responsiveness is interesting because membrane potentials have been shown to be crucial

in motility regulation70, mechanosensation206 and adhesion78. The changes in the sensitivity

of cells with biofilm growth are likely due to differences in the cells’ metabolic states and

due to the influence of the surrounding cells and the biofilm EPS. Our Hodgkin-Huxley style

model demonstrates how these differences may explain the changes in the membrane

potential response observed at different stages of biofilm growth.

Intracellular photosensitiser and antioxidant levels vary with the cells’ metabolic

state and have been shown to alter the magnitude of the ROS response30. The initial

colonising cells (Stage I) were at a stationary growth phase, following overnight planktonic

growth, which may explain why they were found to be less sensitive to 405 nm light than

the cells at Stage II and Stage III of biofilm growth. The steepness of the membrane potential

dose response (1

𝑥) increased by 67 ± 3 % from Stage I to Stage II and the hazard slope

constant (𝑏𝑡) increased by 41.1 ± 0.3 %. This implies that a cells’ metabolic state may greatly

affects its responsiveness to 405 nm light. During Stage III of biofilm growth, cells continued

to become even more responsive. This, combined with larger adhesive forces that

prevented cell dispersal, may explain the morphological changes (coccoid formation)

observed exclusively at this stage of biofilm growth.

The membrane potential dose response was 2.91 ± 0.02 times less steep in mature

biofilm cells (Stage V) than in initially adhered cells (Stage I). This combined with the lack of

dispersal suggests that biofilm growth may afford considerable protection for cells against

405 nm light. Initial cells that could not leave the surface hyperpolarised on a similar

timescale to the initial cells in the flow cells that were free to leave, but following

128

hyperpolarisation, these cells depolarised to levels below the initial resting potential and

showed physical damage, suggestive of cell death. This further suggested that the

differences in the responses of biofilm cells were due to additional protection afforded by

biofilm growth. This is an important consideration for antimicrobial therapies seeking to

use 405 nm light. Biofilm growth is associated with changes in metabolic activity, motility,

and adhesion. Higher levels of catalase, which significantly enhance the protection against

oxidative stress, have been detected in biofilm cells34. During biofilm growth, the role of

other external factors and surrounding cells becomes increasingly important. It has been

shown that alginate can protect P. aeruginosa biofilms by shielding them against UV35. The

biofilm matrix polysaccharides, cellulose and alginate, have also been shown to protect

against reactive oxygen species generated under stress36.

Additional experiments, in which the laser was turned on and off, were performed

to probe the mechanisms and timescales involved in the membrane potential response. It

was found that, above the threshold required to elicit a response, even if the laser was

switched off the response continued but at a slower rate. The ThT fluorescence slowed as

a function of time and sped up as a function of dose. When the laser was pulsed the

membrane potential still hyperpolarised, with a significantly faster dose response than

when the laser was on constantly. This is an important result and indicates that for long

term fluorescence microscopy experiments, it may not be valid to assume the dose is low

enough not to affect cells. Our non-linear Hodgkin-Huxley style model can be used to

explain these results, although the changes in the dose response produced by our model

were less drastic than observed experimentally. In the future, further experiments, on a

wider range of timescales may be used to inform a more complex model e.g. with a

different dose response function or additional time dependent biochemical processes e.g.

additional ion channels.

129

Studies have demonstrated the importance of membrane potentials in the stress

response of P. aeruginosa and B. subtilis. Starvation-induced dispersal in P. aeruginosa

operates through the intracellular second messenger cAMP and requires a membrane

potential207. One study also found that B. subtilis biofilms cells may communicate nutrient

stress via electrical signalling85. Our results show that membrane potential changes may

also play a role in the response of B. subtilis to photoinduced oxidative stress, although in

our experimental geometry the main determinant was the total dosage of light received by

each cell rather than communication between the cells.

I hypothesize that in the future further evidence will emerge connecting a range of

different bacterial stresses and responses with associated changes in membrane potential.

4.6 Conclusions

Membrane potential hyperpolarisations were seen in both the Gram-positive

bacterium B. subtilis and the Gram-negative bacterium P. aeruginosa. This is the first time

that membrane potential hyperpolarisations have been linked with photooxidative stress

in bacteria. The photophysical response of cells to 405 nm light included cell dispersal in

the early stages of biofilm growth, which is problematic for some treatments seeking to

implement this technique for widespread decontamination, but useful for others hoping to

prevent biofilm surface formation. The cell dispersal and the cell membrane potential dose

response were both dependant on the stage of biofilm growth. The membrane potential

dose response was 2.91 ± 0.02 times less steep in mature biofilm cells than in initially

adhered cells. This suggests that biofilm growth affords considerable protection for bacteria

against 405 nm light. A Hodgkin-Huxley model was able to describe the variations in the

experimental membrane potential dose response observed at different stages of biofilm

growth and due to the addition of ROS scavengers. Residence probabilities provided a

130

robust statistical tool to quantify the motile response of the bacteria to irradiation. These

results provide new insight into the role of membrane potentials in the bacterial stress

response and could be used in the development of 405 nm light-based biofilm treatments.

131

CHAPTER

FIVE

5 Measuring c-di-GMP levels and the membrane potential

response of Pseudomonas aeruginosa exposed to oxidative

stress

5.1 Overview

The secondary messenger cyclic di-guanosine monophosphate (c-di-GMP) is a

crucial regulator in biofilm growth in a range of bacterial species. Its role in regulating the

photooxidative stress response of P. aeruginosa to 405 nm light was investigated. More

broadly the oxidative stress response of P. aeruginosa and the connection between c-di-

GMP levels and membrane potentials was studied. The c-di-GMP levels were monitored

using a fluorescence-based GFP reporter, it was, therefore, necessary to assess the

suitability of using GFPs to study oxidative stress in bacteria. Photobleaching was used as a

tool to probe the mechanisms affecting GFP fluorescence.

5.2 Introduction

The threat to health posed by bacterial biofilms is globally recognised. Biofilms

protect bacteria from a wide range of external stresses5,20. They also represent a

coordinated form of growth which can adapt to a broad range of environmental

conditions28,36.

C-di-GMP has received increasing attention as one of the most important bacterial

secondary messengers. Its role in regulating the transition from the planktonic to the sessile

state and back again has been extensively studied, especially in the model organism P.

132

aeruginosa42–44. These studies have shown that c-di-GMP is involved in the regulation of a

broad range of biofilm associated characteristics from adhesion to exopolysaccharide

production. Low levels of intracellular c-di-GMP are associated with the planktonic state

and motility, whereas high levels of intercellular c-di-GMP promote the formation of

biofilms25,46,208.

Figure 5.1. Physiological functions of the intracellular secondary messenger c-di-GMP. C-di-

GMP is synthesised from 2 GTPs via diguanylate cyclases and is degraded into pGpG/GMP

via phosphodiesterases. Extracellular signals control the activity of these proteins and

therefore ultimately regulate the levels of intracellular c-di-GMP. Low levels of c-di-GMP are

associated with the promotion of planktonic behaviour (e.g. motility and acute virulence),

whereas high levels of c-di-GMP are associated with biofilm growth.

The levels of c-di-GMP in a cell are modified by the rate of its synthesis and

degradation. C-di-GMP is synthesized from two molecules of GTP by enzymes called

diguanylate cyclases (DGCs) and is degraded into 5′-phosphoguanylyl-(3′-5′)-guanosine

(pGpG) and/or GMP by phosphodiesterases (PDEs)43 (Figure 5.1). Environmental signals and

stresses modulate the activity of such proteins and therefore ultimately control the

intercellular c-di-GMP level. The balance between synthesis and degradation is complex

with a large number of c-di-GMP pathways thought to influence the c-di-GMP levels46. The

133

sensory input for most DCGs and PDEs remains unknown. Ultimately it is hoped that a

better understanding of c-di-GMP signalling pathways, from the sensory inputs to the

effector functions, will lead to the development of new methods to control biofilm growth.

Intracellular c-di-GMP levels have been shown to regulate a variety of responses,

to a broad range of stimuli, across a large array of different bacteria. This includes

influencing the resistance to and regulation of oxidative stress209–211. C-di-GMP levels have

also been shown to be altered in response to oxidative stress209. As detailed in Chapter 4

(Section 4.2), biofilm growth can provide significant protection against photooxidative

stress. At noncytotoxic levels, reactive oxygen stress can stimulate biofilm formation, via an

increase in c-di-GMP209. In contrast, at higher levels, reactive oxygen stress can inhibit

biofilm growth or induce biofilm dispersal, via lower c-di-GMP levels. For example, it has

been shown that at concentrations that do not supress growth, H202 stimulates biofilm

formation in P. aeruginosa, while at higher levels H202 inhibits biofilm formation212. Reactive

oxygen species have also been shown to drive the evolution of pro-biofilm variants in P.

aeruginosa through the modulation of c-di-GMP levels45.

Dispersal and membrane potential hyperpolarisations were observed in P.

aeruginosa in response to photooxidative stress induced by 405 nm light (Chapter 4). C-di-

GMP levels have been shown to regulate the motility response of P. aeruginosa205. This

combined with the role c-di-GMP levels play in regulating oxidative stress led us to

investigate whether c-di-GMP levels regulate the photooxidative stress induced by 405 nm

light. More generally the aim was to test the possible connection between membrane

potentials and c-di-GMP.

Most studies of c-di-GMP levels in bacteria use quantitative assays such as

LC/MS213. These techniques can be highly sensitive and successful; however, they require

preparation of bacterial extracts and so cannot be used to conduct non-destructive,

134

dynamic measurements. This inspired the development of fluorescence-based c-di-GMP

biosensors which, although only semiquantitive, can be used to study and compare

variations in c-di-GMP levels, in real-time, in-vivo214.

To monitor and compare c-di-GMP levels the fluorescence-based reporter P.

aeruginosa PA01 pCdrA::gfpc was used. The green fluorescent protein (GFP) is widely used

across the life sciences and has proved to be an excellent tool due to its heritability,

specificity and stability99,100. Since the original discovery of wild type GFP, many different

variants of GFPs have been engineered to accommodate the evolving needs of

researchers99,100,215. As with other fluorophores, GFPs undergo photobleaching following

light exposure (see Fluorescence microscopy (2.3) for more information). The dynamics of

photobleaching are highly dependent on the environmental conditions95,216. As well as

having a direct effect on fluorescence through photobleaching, light can also impact GFP

fluorescence by inducing other environmental changes90.

5.3 Materials and methods

5.3.1 Cell culture and growth

Experiments were conducted using P. aeruginosa PA01, P. aeruginosa PA01::gfp, P.

aeruginosa PA01 pCdrA::gfpc and E. coli DH5α pCdrA::gfpc (Table 5.I). E. coli was grown on

LB agar and in LB medium (Table 5.II). During plasmid transformations, P. aeruginosa PA01

was cultured on LB agar and in LB medium. For both c-di-GMP plate reader assays and

microscopy experiments, P. aeruginosa PA01 was either grown on ABTC agar or on TSB agar

and was cultured in ABTG+casA, TSB or 1 % TSB medium. Media recipes are shown in Table

5.II and antibiotics were supplied where necessary according to the quantities shown in

Table 5.I. The ABTG+casA medium and ABTC agar were made following the recipes of

135

Rybtke214. Briefly, the AB medium of Clark and Maaløe was prepared using medium A and

medium B, mixing them after autoclaving by pouring A into B. Medium A was prepared by

dissolving 2 g (NH4)2SO4, 6 g Na2HPO4 and 3 g KH2PO4 3 g NaCl in 200 ml of water. Solution B

was made by dissolving 0.010 g CaCl2, 0.1 g MgCl2 and 0.0005 g/l FeCl3 in 800 ml water. The

ABTG+casA medium was made by supplementing the AB medium with 2.5 mg thiamine

liter−1, 0.5 % [wt/vol] glucose and 0. 5% [wt/vol] Casamino Acids. The ABTC agar was

prepared by supplementing the AB medium with 2.5 mg thiamine liter−1 and 10 mM citrate.

Table 5.I. Descriptions and details for the bacterial strains used in this chapter.

Strain Description Antibiotic resistance (µg/ml)

Source/reference

E. coli DH5α pCdrA::gfpc

Growth Strain for the pCdrA::gfpc plasmid

Gm15 Rybtke et al. (2012)214

P. aeruginosa PA01

Common lab strain

- Lab collection

P. aeruginosa PA01::gfp

Control strain which expresses GFP constitutively

Gm60 Koch et al. (2001)217

P. aeruginosa PA01 pCdrA::gfpc

Fluorescence-based reporter of the level of the nucleotide secondary messenger cyclic di-GMP in P. aeruginosa PA01.

Gm60 Rybtke et al. (2012)214

Table 5.II. Recipes and sources for the culture media used in this chapter.

Media recipe Source/reference

LB 10 g/l NaCl,5g/l Yeast extract, 10 g/l Tryptone, supplemented with antibiotics as required.

-

136

LB agar 10 g/l NaCl,5 g/l Yeast extract, 10 g/l Tryptone, 15 g/l agar, supplemented with antibiotics as required.

-

TSB 30 g/l Tryptone Soya Broth supplemented with antibiotics as required.

-

TSA 30 g/l Tryptone Soya Broth, 15 g/l agar, supplemented with antibiotics as required.

-

1 % TSB 10 g/l Tryptone Soya Broth supplemented with antibiotics as required.

-

ABTG+casA 2 g/l (NH4)2SO4, 6 g/l Na2HPO4, 3 g/l KH2PO4 3 g/l NaCl, 0.010 g/l CaCl2, 0.1 g/l MgCl2, 0.0005 g/l FeCl3, 0.0025 g/l thiamine, 5 g/l glucose and 5 g/l Casamino Acids, supplemented with antibiotics as required.

(AB media) Clark and

Maaløe218

(ABTG+casA)Rybtke214

ABTC agar 2 g/l (NH4)2SO4, 6 g/l Na2HPO4, 3 g/l KH2PO4 3 g/l NaCl, 0.010 g/l CaCl2, 0.1 g/l MgCl2, 0.0005 g/l FeCl3, 0.0025 g/l thiamine, 2g/l citrate, 15 g/l agar, supplemented with antibiotics as required.

(AB media) Clark and Maaløe218

(ABTC agar)Rybtke214

To prepare an overnight culture, cells were freshly streaked onto the relevant

media plates from glycerol stocks two days before the experiment and incubated at 37oC

overnight. The next day 10 ml of the relevant medium in a glass universal was inoculated

with a single colony. The inoculum was then incubated and shaken at 200 rpm at 37oC

overnight (until OD600 ≈ 2). The optical density of cells was measured using a

spectrophotometer (Labtech).

5.3.2 The c-di-GMP reporter PA01 pCdrA::gfpc and the GFP control

strain PA01:gfp

137

PCdrA-gfpc was a gift from Tim Tolker-Nielsen (Addgene plasmid #111614)214.

PCdrA::gfpc uses the cdrA promoter fused with the artificial optimized ribosomal binding

site (RBSII) and GFP as a reporter of c-di-GMP levels in P. aeruginosa (PcdrA-RBSII-

gfp(mut3)-T0-T1). This gene cassette was inserted into the shuttle vector pUCP22Not

(Figure 5.2).

PA01:gfp was used as a control strain. It expresses GFP constitutively (it is produced

in relatively constant amounts regardless of environmental conditions) from miniTn7-based

chromosomal insertion. The chromosomal insertion was driven via miniTn7 expressing

gentamycin and chloramphenicol resistance expressing GFP from the PA1//04/03 promoter.

Figure 5.2. Sequence map of pCdrA::gfpc. (a) Addgene full sequence map for pCdrA::gfpc

created with SnapGene. Shown on the map are: unique 6+ cutters, primers, features and

138

translations. (b) Schematic showing the horizontal cassette map for pCdrA::gfpc showing

the cdrA promoter fused with the artificial optimized ribosomal binding site (RBSII). The

transcriptional fusion is followed by two transcriptional terminators (T0 and T1).

5.3.3 Transformation of pCdrA::gfpc into P. aeruginosa PA01 by

electroporation

PCdrA::gfpc was gifted in E. coli DH5α as an agar stab (from Tom Tolker-Nielson).

The E. coli DH5α pCdrA::gfpc was streaked onto LB agar (containing Gm15) and incubated at

37oC overnight. The next day a single colony was inoculated into 10 ml of LB mixed with

Gm15, which was incubated at 37oC and shaken at 200 rpm overnight. The PCdrA::gfpcc

plasmid was extracted from DH5α using the standard QIAGEN Plasmid protocol (see

Appendices for full protocol).

The plasmid DNA concentration and purity were checked using a Nanodrop ND-

1000 ultraviolet (UV) spectrophotometer (Labtech). Having confirmed the plasmid

preparation was successful, it was transformed into competent P. aeruginosa PA01 by

electroporation. P. aeruginosa cells were made electrocompetent using a sucrose

microcentrifuge-based procedure, following the protocol of Choi et al. 219. An overnight

culture of P. aeruginosa was grown in 6 ml of LB as described previously. This culture was

then separated into 4 different 1.5 ml Eppendorf tubes before the cells were harvested by

centrifugation at 22°C for 2 mins at 16,000 g. The cell pellet in each tube was washed twice

with 1 ml of 300 mM sucrose and the 4 cell pellets were then resuspended in a combined

total of 100 µl of 300 mM sucrose.

For electroporation, 500 ng of the prepared pCdrA::gfpc plasmid was mixed with

100 µl of electrocompetent cells and this mixture was transferred to a 2 mm gap width

electroporation cuvette (Bio-rad). The electroporation was performed at: 25 AF; 200 V; 2.5

kV, using a Bio-Rad GenePulserXcellk. Following electroporation, 1 ml of 22°C LB medium

139

was added at once and cells were transferred to a 1.5 ml Eppendorf. The cells were

incubated at 37°C and shaken at 200 rpm for 1.5 hrs. The cells were then harvested, by

centrifugation at 22°C for 2 min at 16,000 g and 900 µl of the supernatant was discarded.

The cell pellet was then resuspended in the remaining medium. The entire mixture was

then plated on LB with Gm60 plates. The plates were then incubated at 37°C overnight.

Controls included cells pulsed without the added plasmid. The next day a single colony was

inoculated into 10 ml of LB, which was incubated at 37oC and shaken at 200 rpm overnight.

In order to confirm successful transformation of pCdrA::gfpc into P. aeruginosa PA01, the

plasmid was extracted from P. aeruginosa PA01 using the modified QIAGEN Plasmid

protocol (detailed in the Appendices) and the concentration and purity was checked using

a Nanodrop ND-1000 ultraviolet (UV) spectrophotometer (Labtech).

The whole and digested extracted plasmid were analysed using agarose gel

electrophoresis. The plasmid was digested with 1 µl Xba1 and 1 µl HindIII restriction

enzymes (BioLabs) added to 7 µl of plasmid, 2 µl of 10x CutSmart Buffer (BioLabs) and 9 µl

of Milli-Q water and incubated at 37°C for 1.5 hrs. A 1 % w/v agarose gel was made by

dissolving 0.5 g agarose into 50 ml TAE buffer (40 mM Tris-acetate pH 7.7, 1 mM EDTA). The

gel was added to 0.5 µl of ethidium bromide (Sigma) and added to the gel holder before it

set. The holder was then placed in the gel electrophoresis system (ThermoFisher). The

samples were loaded and then run alongside a 1Kb HyperLadder (Bioline) at 110 V DC. The

samples were compared to the ladder using a UV transilluminator (UVIpro Silver, UVItec)

to determine their size. Finally, P. aeruginosa PA01 pCdrA::gfpc was tested using the SNP

fluorescence plate reader assay.

5.3.4 Plate reader assay

140

Microtiter plate-based assays were carried out following the protocol of Rybtke214.

Experiments testing P. aeruginosa PAO1 pCdrA::gfpc as a reporter of c-di-GMP levels using

SNP (Sodium nitroprusside) cells used ABTC agar and ABTG+casA medium. Other

experiments measuring the effect of H202 on c-di-GMP levels cells were conducted using

TSA and 1 % TSB medium.

Overnight inoculums were diluted to OD600 ≈ 0.03 in 30 ml of fresh medium. For

H202 assays, 250 μl of the culture was immediately added to the wells of the microtiter. For

SNP assays the growth was measured manually until OD600 ≈ 0.4, at which point 250 μl of

the cell culture was added to the microtiter wells. After the black-welled, clear bottomed,

96-well Corning microtiter plate was filled, it was incubated in a BioTek Synery HT plate

reader heated to 37°C and set up to measure OD450/OD600 and green fluorescence (arbitrary

GFP units) every 30 min. The plates were shaken in an orbital pattern (3-mm diameter) at

normal speed for 10 min before and after each round of measurements to optimize growth.

The SNP (Sigma) was prepared in 2-fold serial dilutions starting from 250 μM SNP

from a 50 mM stock of SNP dissolved in Milli-Q water. The H202 (Sigma) was prepared in

serial dilutions from a 1 mM stock dissolved in Milli-Q water.

5.3.5 Cell preparation for single cell fluorescence microscopy

Following overnight culture in TSB, as previously described, fresh 1 % TSB medium

with relevant antibiotics was reinoculated 1:10 with the overnight culture and grown with

shaking at 200 rpm at 37°C until the cells reached mid-exponential phase (OD600 ≈ 0.5). This

cell culture was then diluted to OD600 ≈ 0.05 using fresh 1 % TSB medium. Agarose

microscope slides were prepared following an adapted version of the protocol of Jong et al.

141

(2011)105 as described previously in Chapter 4 (Section 4.3.2). These slides were prewarmed

for at least an hour prior to imaging to avoid issues arising from drifting.

5.3.6 Microscope set-up

The microscope slides were mounted on an Olympus IX-71 inverted microscope

with an Olympus UPAON 100XOTIRFM (NA 1.49) immersion oil and TIRF objective lens.

Brightfield images were acquired using an LED (525 nm) which was focused onto the sample

using a condenser lens. For fluorescence imaging, the sample was illuminated with an OBIS

405LX or OBIS 488LX laser. Time lapse images were recorded using an ORCA-Flash4.0 LT

PLUS Digital CMOS camera (C11440-22CU).

5.3.7 Fluorescent dyes and GFPs

ThT is commonly used to stain amyloid fibres96, however, its positive charge also

allowed it to be used as a Nernstian voltage indicator85–87. ThT was supplied by Sigma-

Aldrich. Fresh 2 mM stocks were made up on the day of the experiment and added to the

cells and media at a final working concentration of 10 µM. ThT can be excited at 405 nm,

allowing it to be used in conjunction with 405 nm light treatment. Both the fluorescent

reporter (PCdrA::gfpc) and the constitutively firing GFP control strain (PA01::gfp) contained

the mut3 variation of GFP which was excited using the 488 nm laser.

5.3.8 405 nm light treatment and photobleaching using 488 nm light

142

To treat cells with 405 nm light an OBIS 405LX laser with an irradiance of 120

µW/cm2 was focused on to the surface of the microscope slide. Time lapse images were

acquired every 1 secs.

Cells were photobleached using an OBIS 488LX laser at an irradiance of 120 µW/cm2

and time lapse images were acquired every 6 secs.

5.3.9 Irradiance/dose measurements

The optical power was measured at the sample using a Thorlabs PM121D digital

power meter. The power across the area in which measurements were conducted was

uniform and so the irradiance was given by,

𝐼 =𝑃

𝐴 , (5.1)

where 𝐼 is the irradiance, 𝑃 is the power at the sample and 𝐴 is the illumination area and

the dose (D) in J/cm2 was given by,

𝐷 =𝐼𝑡

1000

(5.2)

where I is the irradiance in mw/cm2 and t is the time of illumination in s.

5.3.10 Data analysis

The data from the plate reader assays was exported to Matlab (MathWorks), where

custom made scripts were used to analyse it. This included the subtraction of the signal due

to background media, averaging over repeats and calculating ratios. All experiments

included at least three repeats with three biological triplicates.

143

The image analysis for the microscopy experiments was conducted using Image J

and Matlab (MathWorks). Background subtraction was done using the ImageJ (National

Institutes of Health) ‘rolling ball’ background plugin with a radius of 4 – 8 µm depending on

the experimental conditions. To obtain fluorescence curves the fluorescence was measured

using the ‘plot Z axis’ function on the ImageJ image analysis toolbox. Signal processing was

performed in Matlab, including signal smoothing using a cubic weighted Savitzky-Golay

filter, which removed additional noise. The average fluorescence per cell was calculated by

averaging the mean fluorescence per cell from at least 20 cells across at least five different

experiments. Origin was used to fit fluorescence curves, to perform the statistical analysis

and for graph plotting. All the errors presented are standard errors.

5.4 Results

5.4.1 Confirming the suitability of P. aeruginosa PAO1 pCdrA::gfpc as a

reporter of c-di-GMP levels

The fluorescence-based reporter P. aeruginosa PA01 pCdrA::gfpc was used to

monitor c-di-GMP levels. This biosensor was originally developed by Rybtke et al. (2012)

who demonstrated its success in detecting an increase in c-di-GMP levels in response to

SNP in P. aeruginosa PAO1 ΔwspFΔpelΔpsl/pCdrA::gfpc 214. This strain was used due to the

higher levels of fluorescence observed, but it is not suitable for biofilm studies, as

wspF/pel/psl are required for biofilm growth. Subsequent studies have used the wild type

P. aeruginosa PAO1 pCdrA::gfpc strain to gauge the levels of c-di-GMP in planktonic and in

biofilm cells220,221.

144

The original experiments of Rybtke were repeated and their results for PAO1

ΔwspFΔpelΔpsl/pCdrA::gfpc were reproduced for PAO1 pCdrA::gfpc (Figure 5.3), confirming

that PAO1 pCdrA::gfpc could be used to successfully gauge the levels of c-di-GMP.

Figure 5.3. Treatment of P. aeruginosa PAO1 pCdrA::gfpc with SNP at concentration of 0

µM, 62.5 µM and 125 µM. (a) Growth measurements given by the OD450. (b) Fluorescence

GFP measurements.

5.4.2 Changes in c-di-GMP levels in response to 405 nm light

145

Fluorescence microscopy was used to dynamically monitor the intracellular c-di-

GMP levels of individual cells treated with 405 nm light. Cells were grown on agarose pad

microscope slides (Section 5.4.5). GFP is photobleached by 405 nm light. It was therefore

not possible to take GFP measurements immediately following 405 nm light treatment, but

it was possible to measure the GFP following recovery. Figure 5.4 shows that after one hour

of recovery that the GFP fluorescence of PA01 pCdrA::gfpc increased, but not back to the

original intensity. Following the same treatment the GFP fluorescence of the constitutively

firing PAO1::gfp strain showed no recovery. This indicates that the continued decreases in

GFP fluorescence following recovery were not reflective of decreases in c-di-GMP levels,

but were instead caused by decreases in the promoter/translator capabilities of GFP or due

to GFP quenching. To probe the mechanism/s responsible for the decreased fluorescence,

I photobleached the GFP. The GFP photobleaching (𝐼(𝑡)) by 488 nm light (at a set irradiance

of 120 ± 2 µW/cm2) was well described by a monoexponential of the form,

𝐼(𝑡) = 𝐼0𝑒−𝑡𝜏

(5.3)

where 𝐼0 is the unbleached GFP fluorescence intensity and 𝜏 is the photobleaching decay

time constant.

The value of 𝐼0 decreased for both PA01 pCdrA::gfpc and PAO1::gfp, but 𝜏 did not

change significantly (Figure 5.4(b)). Before treatment with 405 nm light the GFP

photobleaching time constant of PA01 pCdrA::gfpc was 0.58 ± 0.07 min and after treatment,

it was 0.55 ± 0.03 min. The photobleaching time constant of PAO1::gfp was 0.75 ± 0.07 min

before treatment and 0.71 ± 0.05 min after treatment. The observed decrease in

fluorescence, but maintenance of photobleaching kinetics, is the expected consequence of

photobleaching alone. It is possible that a static quencher may also be acting, but the

maintenance of 𝜏 rules out dynamic quenching following treatment with 3.6 mJ/cm2 of 405

nm light90,92.

146

Figure 5.4. Treatment of P. aeruginosa PAO1 pCdrA::gfpc and P. aeruginosa PAO1::gfp with

405 nm light. (a) Normalised GFP fluorescence per cell before, following and an hour after

treatment of cells with 3.6 mJ/cm2 of 405 nm light. (b) Photobleaching curves for P.

aeruginosa PAO1 pCdrA::gfpc exposed to 488 nm light at an irradiance of 120 ± 2 µW/cm2

(black) before and (red) after treatment with 3.6 mJ/cm2 of 405 nm light, with exponential

fits given by Equation 5.3.

5.4.3 Changes in c-di-GMP levels in response to H202

As it was not possible to measure the c-di-GMP changes in response to 405 nm

light, the changes accompanying H202 oxidative stress were measured. Figure 5.5 shows the

results of plate reader assays in which P. aeruginosa PA01 pCdrA::gfpc , P. aeruginosa

PAO1::gfp and P. aeruginosa PAO1 were grown in the presence of 1 mM H202. At this

concentration the growth of PA01 pCdrA::gfpc and PAO1::gfp were affected, while the

growth of PA01 WT was not (Figure 5.5(b)). The growth of PAO1::gfp was slower than PA01

pCdrA::gfpc and the addition of H202 had a larger impact on the growth of P. aeruginosa

PAO1::gfp.

147

Figure 5.5. Treatment of P. aeruginosa with 1 mM H202. (a) Ratios of the average GFP

fluorescence to OD600 as a function of time since inoculation for: P. aeruginosa PAO1

pCdrA::gfpc (▲), P. aeruginosa PAO1::gfp (■) and P. aeruginosa PAO1 (●), without H202

(black) and with 1 mM H202 (red). (b) Growth curves of: P. aeruginosa PAO1 pCdrA::gfpc

(▲), P. aeruginosa PAO1::gfp (■) and P. aeruginosa PAO1 (●), without H202 (black) and with

1 mM H202 (red). (c) Ratio of average GFP fluorescence to OD600 , at mid-exponential growth

phase (OD600 ≈ 0.5), without H202 (blue) and with 1 mM H202 (green) for: P. aeruginosa PAO1

148

pCdrA::gfpc (▲), P. aeruginosa PAO1::gfp (■) and P. aeruginosa PAO1 (●). The presented

errors are standard errors.

Figure 5.5(a) shows the ratio of GFP fluorescence to growth (given by the optical

density (OD600)). At all stages of growth the GFP/ OD600 ratio of PA01 pCdrA::gfpc was larger

when grown in the presence of 1 mM H202. The GFP/ OD600 of PA01 pCdrA::gfpc increased

over 10 hrs of growth, with comparable increases observed when grown in the presence of

H202. The ratios of the background GFP fluorescence and autofluorescence to OD600 (given

by P. aeruginosa PAO1::gfp and PA01 WT respectively) decreased over 10 hrs of growth,

with both ratios showing comparable changes when grown in the presence of H202. Because

of the differences in the growth rates, the GFP/ OD600 values at mid-exponential growth

phase (OD600 ≈ 0.5) were compared (Figure 5.5(b)). The GFP/ OD600 of PA01 pCdrA::gfpc was

larger when 1 mM of H202 was added. In contrast, the ratio was lower in both P. aeruginosa

PAO1::gfp and PA01 WT upon addition of H202. This indicates that the increases seen in

PA01 pCdrA::gfpc were caused by an increase in the c-di-GMP levels, due to the addition of

low levels of H202.

It was not possible to conduct ThT measurements in the plate reader and so in

order to conduct parallel membrane potential and c-di-GMP measurements fluorescence

microscopy was used. Sessile cells are known to have a higher tolerance to H202 than

planktonic cells. We, therefore measured the response of cells on the agar microscopy

slides to a range of H202 concentrations between 1 μm – 10 mM. Figure 5.6(a) shows that

addition of 1 μm – 1 mM H202 did not significantly affect the average GFP fluorescence of

PA01 pCdrA::gfpc cells at the 5 % significance level. However, addition of >1 mM H202

significantly decreased the fluorescence. These results were mirrored in the control strain

PAO1::gfp, suggesting that these results were not caused by changes in the levels of c-di-

GMP. Photobleaching of the GFP using a 488 nm laser was faster (smaller 𝜏 in Equation 5.3)

149

at the higher H202 concentrations (Figure 5.6(b)). Faster bleaching was observed from 1 mM

H202 in PA01 pCdrA::gfpc and after 1 mM in PAO1::gfp.

Figure 5.6. Treatment of P. aeruginosa PA01 pCdrA::gfpc and P. aeruginosa PA01::gfp with

H202. (a) Average cell GFP fluorescence of P. aeruginosa PA01 pCdrA::gfpc and P. aeruginosa

PA01::gfp, for a range of H202 concentrations (1 µM, 10 µM, 1 mM and 10 mM), normalised

to the average cell GFP fluorescence without H202. (b) The normalised average cell decay

time constant of photobleaching by 120 ± 2 µW/cm2 488 nm light (Equation 5.3) as a

function of H202 concentration. The inset shows the decrease in cell fluorescence due to

addition of 10 mM H202 in fluorescence microscopy images. The presented errors are

standard errors.

5.4.4 Membrane potential response to H202

Previous ThT experiments (detailed in Chapter 4) were repeated in the presence of

H202. The average cell ThT fluorescence (averaged over at least 20 cells across at least five

different experiments) was significantly higher in the presence of 10 mM H202 (Figure

5.7(a)). Treatment with 405 nm light produced previously described membrane

hyperpolarisations, but these occurred at a faster rate in the presence of H202 (Figure

5.7(b)). Overall this suggests that P. aeruginosa hyperpolarises in response to H202 and that

150

the combination of H202 and 405 nm light is synergistic causing a larger oxidative stress and

so a faster hyperpolarisation response.

Figure 5.7. Membrane potential response of P. aeruginosa treated with H202. (a) Change in

average ThT fluorescence per cell caused by addition of 10 mM H202. (b) Average normalised

ThT (membrane potential) dose response without H202 (black) and with 10 mM H202 (red) in

response to 120 µW/cm2 405 nm light. The presented errors are standard errors.

5.5 Discussion

Biofilm growth relies on the coordination of behaviour between its constituent

bacteria; this is achieved via a complex network of signalling molecules and genetic cues.

The secondary messenger c-di-GMP is a crucial regulator in biofilm growth in a range of

bacterial species.

Recent studies have revealed that 405 nm light induces photophysiological

responses in a range of bacteria via different blue-light receptor classes (LOV, BLUF and

PYP)197,198 as described in the previous chapter (Section 4.5) The motility of Synechocystis

sp. PCC 6803 is affected by 405 nm light via an cyclic diguanylate (c-di-GMP) signal

transduction system204. P. aeruginosa has been shown to regulate behaviour in response to

151

light202 and regulates stress via c-di-GMP/cAMP levels205. The majority of diguanylate

cyclases (DGCs) and phosphodiesterases (PDEs) that have been identified in the genome of

P. aeruginosa still remain uncharacterized43. One hypothesis is that 405 nm light stimulates

the production of one or more of these PDEs, ultimately resulting in lower levels of c-di-

GMP and cell dispersal.

Unfortunately, it was not possible to measure the c-di-GMP levels following 405 nm

light treatment and so were unable to fully test this hypothesis. This was partly due to GFP

photobleaching, but was mainly because the GFP fluorescence was fundamentally affected

by the 405 nm light. The levels of c-di-GMP following recovery after photobleaching were

measured and it was found that the control strain which produced GFP constitutively did

not recover. In order to investigate what caused the GFP fluorescence decreases following

405 nm light treatment, the sample was photobleached using 488 nm light. The observed

decrease in fluorescence was not accompanied by a change in photobleaching kinetics. This

is the expected consequence of photobleaching alone. It is possible that a static quencher

may also be acting, but the maintenance of photobleaching kinetics rules out dynamic

quenching following treatment with 3.6 mJ/cm2 of 405 nm light90,92. Treatment of cells with

H202 affected the kinetics of photobleaching. We therefore suggest that the decreases in

fluorescence induced by 405 nm light were not caused by the quenching of GFP

fluorescence via ROS, or more specifically H202 (the main ROS produced by bacteria in

response to 405 nm light171). It is, however, possible that the photooxidative stress induced

by 405 nm light affected the promoter/translator capabilities of GFP. In the future, it may

be possible to test this hypothesis by using a fluorescence-based reporter which uses an

alternative fluorescent protein. The reporter strain PA01 pCdrA::gfpc contains GFPmut3,

which is more sensitive to photobleaching than some other fluorescent proteins216.

However, other fluorescent proteins may be unsuitable for alternative reasons, for

example, eGFP can induce high levels of oxidative stress in cells222. For future work the

152

development of a c-di-GMP reporter strain with a red fluorescent protein would be

useful223.

The levels of H202 required to elicit a response from adhered P. aeruginosa PA01

cells altered the GFPmut3 fluorescence and so our reporter could not be used. The decrease

in GFP fluorescence observed in response to > 1 mM of H202 was dynamic and decreased in

a time dependant manner when the sample was photobleached by 488 nm light. Exposure

to light can induce ROS production and the presence of a fluorescent protein that absorbs

the excitation light enhances ROS production224. If ROS was causing the fluorescence

quenching of GFP, the production of additional ROS could explain the time dependant

changes in fluorescence. This indicates that fluorescence imaging of GFPs in the presence

of ROS may induce significant ROS production, which may not only render the GFP

ineffective as a reporter of specific gene transcription but could also induces changes in cell

behaviour. Recent studies have shown that fluorescent proteins, such as eGFP, can induce

catalytic oxidative stress in biological systems222.

Also, of concern was the effect of GFP production on oxidative stress resilience. Our

control strain (P. aeruginosa PA01::gfp) had a larger GFP fluorescence than our reporter

strain (P. aeruginosa PA01 pCdrA::gfpc) and it was more sensitive to oxidative stress. Its

recovery was slower following 405 nm light treatment and its growth was more greatly

affected by H202. The growth of both the strains expressing GFP were slowed by the

addition of H202, whereas the wild type PA01 remained unaffected.

Taken together, these results demonstrate that GFP reporters may not be suitable

for systems with high levels of ROS. This is an important result for future studies and

highlights the importance of using a constitutively firing GFP strain as a control. Previous

studies have not always used such controls, but our results demonstrate the need to do so.

Excitation at longer wavelengths generally incurs less ROS production and damage than

153

excitation at equivalent irradiances with a shorter wavelength224,225. It would, therefore, be

interesting to develop a reporter strain which is based on a fluorescent protein that is

excited at a longer wavelength (e.g. red) and to test whether it performs better in the

presence of ROS.

Planktonic P. aeruginosa PA01 cells have a lower tolerance to H202 than biofilm

cells226,227. It was therefore possible to use a plate reader assay to study the effect of ROS

(caused by H202) on the c-di-GMP levels of cells. At lower levels of H202, where growth was

only minimally affected, the H202 stimulated an increase in the levels of intracellular c-di-

GMP. It is not possible to conduct membrane potential (ThT) measurements in the plate

reader assay and so it was not possible to directly connect these results with changes in the

membrane potential. However, via fluorescence microscopy, it was possible to study how

ROS affected the membrane potentials of surface adhered P. aeruginosa PA01 cells. The

addition of 10 mM H202 caused membrane potential hyperpolarisations and the addition of

H202 increased the rate of the previously described 405 nm light induced membrane

potential hyperpolarisations.

In summary, it was not possible to test the hypothesis that the cell dispersal and

membrane potential hyperpolarisations that were observed in response to photooxidative

stress were associated with changes in the levels of intracellular c-di-GMP levels. However,

it was possible to show that at low levels oxidative stress can induce increases in the levels

of c-di-GMP. This indicates that P. aeruginosa does regulate ROS stress via this intracellular

messenger.

154

5.6 Conclusions

The GFP-based fluorescent reporter P. aeruginosa PA01 pCdrA::gfpc was used to

investigate the role c-di-GMP levels play in the regulation of the photooxidative stress

response of P. aeruginosa to 405 nm light. The photooxidative stress induced changes in

the GFP and therefore PA01 pCdrA::gfpc could not be reliably used to measure the c-di-

GMP levels. Photobleaching kinetics were used to understand the mechanisms causing the

reduction in GFP fluorescence. It was found that the decreases were caused by decreases

in the expression of the GFP (due to different promoter/translator capabilities) and/or due

to static GFP quenching. Oxidative stress, in the form of H202, also directly affected the GFP

fluorescence. Our results suggest that 488 nm light (the wavelength used for excitation and

imaging of GFP) added to the oxidative stress.

Our results also suggest that strains which produce a larger quantity of GFP are more

sensitive to oxidative stress. The two key requirements of a successful fluorescence-based

biosensor are that the levels of fluorescence are reflective of the gene expression of interest

and that the sensing mechanism (in this case GFP expression) does not affect the cells. It

was found that both these requirements were not met by the GFPmut3-based reporter in

environments with oxidative stress > 1 mM H202.

At lower levels of oxidative stress the c-di-GMP levels were faithfully reported by

PA01 pCdrA::gfpc and it was possible to show that c-di-GMP levels were raised in response

to H202. Therefore, the oxidative stress response of P. aeruginosa PA01 was regulated by c-

di-GMP. At higher levels, it has been shown that H202 can induce dispersal of P. aeruginosa

via c-di-GMP. I therefore hypothesize that c-di-GMP regulates the observed dispersal

response of PA01 to 405 nm light and more broadly that intracellular c-di-GMP levels are

connected to the cell’s membrane potential. In the future, it may be possible to investigate

155

this using a fluorescence-based c-di-GMP reporter which uses an alternative fluorescent

protein.

156

6 Conclusions

Fluorescence microscopy was combined with mathematical techniques to

investigate the role membrane potentials play in the stress response of bacteria and in

bacterial communication. It recently emerged that bacteria in biofilms use electrical signals

to communicate85. Such universal mechanisms of communication offer particularly

attractive targets to develop treatments, as they may offer a widespread solution.

In order to understand the new class of active excitable matter that is bacterial

biofilms, it may be possible to learn from the well-established field of eukaryotic excitable

matter. We have demonstrated how new methods of data analysis and agent-based

modelling may be used to investigate electrical signalling in bacterial biofilms. These

methods were used to describe newly observed phenomena in the electrical signals of

circular B. subtilis biofilms. Centripetal electrical wavefronts, which travelled inward from

the edge of a circular biofilm were described for the first time. Moments’ analysis was used

to rigorously quantify these signals, as well as previously described centrifugal signals. This

new method of analysis revealed that, contrary to previous belief, the electrical signals did

not propagate constantly through the biofilm. The centrifugal wavefront travelled slower

than the centripetal wavefront, demonstrating that the curvature of the biofilms affected

the signal propagation, in agreement with theoretical predictions165. Also, in contradiction

to prior belief, it was found that the fluorescence energy density and the amplitude of

wavefronts decreased with distance from the biofilm centre.

An agent-based fire-diffuse-fire (ABFDF) model was used to show that the

arrangement of cells and curvature alone were enough to explain the key characteristics of

electrical wave propagation. However, some of the subtler characteristics of signal

propagation were not predicted by our ABFDF model. In the future it may be possible to

also explain these characteristics by adding more specific details of biofilm growth (e.g.

157

growth of bacteria in chains) to the current model. In nature, biofilms form complex, three-

dimensional structures so another extension to this project would involve studying

electrical signalling in three-dimensional biofilms. This is expected to be significantly more

challenging. The exact origin of the centripetal wavefronts remains elusive. This could be

investigated by studying B. subtilis deletion strains (e.g. yugO deletion) under varying

environmental conditions.

Charge transfer through a biofilm defines its electrical properties and is thus an

important consideration when studying electrical fluctuations and potentials. The diffusion

of charged and uncharged molecules through a biofilm is dependent on the structure and

composition of the EPS. For example, Geobacter sulfurreducens, an electricigen, produces

protein filaments which increase the efficiency of long-range electron transport228. The

composition and structure of a B. subtilis biofilm varies depending on the experimental

conditions20,51 (Section 1.2). The structure and composition of the circular B. subtilis biofilms

studied has not yet be quantified. In the future the biofilms’ fine structure may be resolved

via super-resolution microscopy, which can overcome the diffraction limit of traditional

techniques while still maintaining the native structure. In order to build an initial picture of

the biofilms’ structure, experiments which label the key components of B. subtilis biofilms

should be possible following the protocols summarised in the review by Schlafer & Meyer

(2015)229. For example, TasA protein fibres can be labelled using established protocols

which use Thioflavin-T or Congo red57.

New photodynamical therapies seek to use 405 nm light due to its intrinsic

antimicrobial effect171,199,230. The role membrane potentials play in the response of bacteria

to 405 nm light was investigated. Membrane potential hyperpolarisations were observed

in both the Gram-negative bacterium P. aeruginosa and the Gram-positive bacterium B.

subtilis. This was the first time that photo-induced membrane potential hyperpolarisations

were observed in bacteria. At the early stages of biofilm growth cells dispersed and/or were

158

physically altered by the 405 nm light. Residence probabilities provided a robust statistical

tool to quantify cell dispersal.

The photophysical and the membrane potential dose response were both

dependant on the stage of biofilm growth. The membrane potential dose response was

2.91 ± 0.02 times less steep in mature biofilm cells than in initially adhered cells. A non-

linear Hodgkin-Huxley model was used to explore differences in the observed responses.

These results provide new insight into the involvement of membrane potentials in the

photoresponse of bacteria.

One possible extension to this work would involve measuring the levels of reactive

oxygen species using either a ROS dye or a fluorescence-based reporter strain. Two

fluorescence-based reporter dyes, 6-carboxy-2',7'-dichlorodihydrofluorescein diacetate

(carboxy-H2DCFDA) and hydroxyphenylfluorescein (HPF), are commonly used to directly

measure ROS production in bacteria231,232. The reliability of these dyes has been called in to

question231,233,234. These dyes also cannot be used to dynamically measure ROS levels in

biofilms as the preparation technique requires cells to be washed before a single

measurement is taken. Therefore, these dyes cannot be used to conduct ROS

measurements alongside the current measurements and a fluorescence-based reporter

would be more suitable. We conducted initial experiments using the protein-based sensors

HyPer and SypHer. HyPer can be used to measure intracellular H202 levels and SypHer can

be used as a HyPer-control and pH sensor235. Unfortunately, HyPer and SypHer were both

photobleached by the 405 nm light and so it was not possible to use them to measure the

effect of the light. In the future it may be possible to develop a fluorescence-based reporter

which operates at different wavelengths so that such experiments can be carried out.

Another possible extension would involve studying the response of bacteria treated

with a 405 nm light-emitting diode (LED) array171,230,236. This would allow comparison

between laser and LED light, which initial studies show have comparable antimicrobial

159

efficacies237, as well as allowing comparison between planktonic and biofilm growth. LED

array experiments are also compatible with ROS dyes. In the future it may be possible to

follow previous experiments in which an LED array is used to treat cells grown in a multi-

well plate171. In this study cells were removed from the wells and washed before the ROS

was measured using the fluorescent dye carboxy-H2DCFDA. The fluorescence was

measured using a spectrofluorophotometer. Similar experiments could be performed

alongside measurements of membrane potentials using ThT, to confirm the association

between 405 nm light treatment, ROS and membrane potential hyperpolarisations.

However, the measurement of fluorescence via this method is not as sensitive or accurate

as the methods we have used. It is also possible that the incubation step required for

carboxy-H2DCFDA may introduce a delay long enough to cause hyperpolarisations to be

missed.

Recent studies have revealed that bacteria respond to 405 nm light via designated

blue-light receptors. The motility of Synechocystis sp. PCC 6803 is affected by 405 nm light

via the cyclic diguanylate (c-di-GMP) signal transduction system204. P. aeruginosa has been

shown to regulate behaviour in response to light202 and regulates stress via c-di-

GMP/cAMP levels205. The majority of diguanylate cyclases (DGCs) and phosphodiesterases

(PDEs) that have been identified in the genome of P. aeruginosa still remain

uncharacterized43. One hypothesis is that 405 nm light stimulates the production of one or

more of these PDEs, ultimately resulting in lower levels of c-di-GMP and cell dispersal. To

test this hypothesis the GFP-based fluorescent reporter P. aeruginosa pCdrA::gfpc was used

to measure c-di-GMP levels. Unfortunately the GFP fluorescence was altered by 405 nm

light so pCdrA::gfpc could not be reliably used to measure c-di-GMP. Oxidative stress, in the

form of H202, also directly affected the GFP fluorescence. Our results suggest that 488 nm

light (the wavelength used for excitation and imaging of GFP) as well as GFP production

added to the oxidative stress. These results suggest that in the presence of oxidative stress

160

(H202 > 1 mM) GFP is not a faithful reporter and that the imaging of GFP may induce further

oxidative stress. This is of interest to a wide range of studies which use GFPs99,100 and our

results demonstrate the importance of using a constitutively firing GFP strain as a control.

At lower levels of oxidative stress the c-di-GMP levels were faithfully reported by PA01

pCdrA::gfpc and it was found that c-di-GMP levels were raised in response to H202.

In the future, to test the hypothesis that c-di-GMP is involved in the response of P.

aeruginosa to 405 nm light, a reporter which faithfully reports c-di-GMP levels in the

presence in high levels of oxidative stress is required. It is possible that a reporter strain

based on a fluorescent protein which is excited at a longer wavelength (e.g. red) may be

used223,238. This would involve developing such a reporter and testing its performance

before using it. Once a suitable reporter has been developed it may be possible to

supplement traditional fluorescence microscopy techniques with fluorescence lifetime

imaging (FLIM)239, which produces an image based on the differences in the excited state

decay rate from the sample and can be used to produce more robust results than intensity

based methods.

Further experiments could also be performed using a 405 nm LED array as discussed

above. Following the work of Ramakrishnan et al (2016) a 405 nm LED array could be used

to treat cells in a multi-well plate171. This would allow the fluorescence plate reader

experiments performed for H202 to be conducted using 405 nm light as the stress. This

would allow our hypotheses to be tested further, as well as extending the study to include

different modes of growth.

161

7 Bibliography

1. Blee, J. A., Roberts, I. S. & Waigh, T. A. Spatial propagation of electrical signals in

circular biofilms: A combined experimental and agent-based fire-diffuse-fire study.

Phys. Rev. E 100, 52401 (2019).

2. Blee, J. A., Roberts, I. S. & Waigh, T. A. Membrane potentials, oxidative stress and

the dispersal response of bacterial biofilms to 405 nm light. Phys. Biol. 17, 036001

(2020).

3. Romeo, T. Bacterial biofilms. vol. 322 (Springer Berlin Heidelberg, 2008).

4. Flemming, H.-C., Neu, T. R. & Wozniak, D. J. The EPS matrix: the ‘house of biofilm

cells’. J. Bacteriol. 189, 7945–7947 (2007).

5. Wingender, J. et al. Biofilms: an emergent form of bacterial life. Nat. Rev.

Microbiol. 14, 563–575 (2016).

6. Jacobsen, S. M., Stickler, D. J., Mobley, H. L. T. & Shirtliff, M. E. Complicated

catheter-associated urinary tract infections due to Escherichia coli and Proteus

mirabilis. Clin. Microbiol. Rev. 21, 26–59 (2008).

7. Makhlouf, A. S. H. & Botello, M. A. Failure of the metallic structures due to

microbiologically induced corrosion and the techniques for protection. Handbook

of Materials Failure Analysis (eds. Makhlouf, A. S. H. & Aliofkhazraei, M. B. T.-H. of

M. F. A.) 1–18 (Elsevier, 2018).

8. Høiby, N. et al. The clinical impact of bacterial biofilms. Int J Oral Sci Int. J. Oral Sci.

3, 55–65 (2011).

9. del Pozo, J. L., Rouse, M. S. & Patel, R. Bioelectric effect and bacterial biofilms. A

systematic review. Int. J. Artif. Organs 31, 786–795 (2008).

162

10. Schultz, M. P., Bendick, J. A., Holm, E. R. & Hertel, W. M. Economic impact of

biofouling on a naval surface ship. Biofouling 27, 87–98 (2011).

11. R Robertson, S. & JC McLean, R. Beneficial biofilms. AIMS Bioeng. 2, 437–448

(2015).

12. Yu, Y., Chen, H.-L., Yong, Y.-C., Kim, D. & Song, H. Conductive artificial biofilm

dramatically enhances bioelectricity production in Shewanella-inoculated microbial

fuel cells. Chem. Commun. 47, 12825–12827 (2011).

13. McLoon, A. L., Guttenplan, S. B., Kearns, D. B., Kolter, R. & Losick, R. Tracing the

Domestication of a Biofilm-Forming Bacterium. J. Bacteriol. 193, 2027–2034

(2011).

14. Beyenal, H., Donovan, C., Lewandowski, Z. & Harkin, G. Three-dimensional biofilm

structure quantification. J. Microbiol. Methods 59, 395–413 (2004).

15. Di Martino, P. Extracellular polymeric substances, a key element in understanding

biofilm phenotype. AIMS Microbiol. 4, 274–288 (2018).

16. Hall, C. W. & Mah, T.-F. Molecular mechanisms of biofilm-based antibiotic

resistance and tolerance in pathogenic bacteria. FEMS Microbiol. Rev. 41, 276–301

(2017).

17. Anderl, J. N., Franklin, M. J. & Stewart, P. S. Role of antibiotic penetration limitation

in Klebsiella pneumoniae biofilm resistance to ampicillin and ciprofloxacin.

Antimicrob. Agents Chemother. 44, 1818–1824 (2000).

18. Chich, J.-F. A mini review: proteomic analysis, a post-genomic approach. Lait 81,

13–18 (2001).

19. Fallis, A. . BslA is a self-assembling bacterial hydrophobin that coats the Bacillus

163

subtilis biofilm. J. Chem. Inf. Model. 53, 1689–1699 (2013).

20. Vlamakis, H., Chai, Y., Beauregard, P., Losick, R. & Kolter, R. Sticking together:

building a biofilm the Bacillus subtilis way. Nat. Rev. Microbiol. 11, 157–168 (2013).

21. O’Toole, G., Kaplan, H. B. & Kolter, R. Biofilm formation as microbial development.

Annu. Rev. Microbiol. 54, 49–79 (2000).

22. Garrett, T. R., Bhakoo, M. & Zhang, Z. Bacterial adhesion and biofilms on surfaces.

Prog. Nat. Sci. 18, 1049–1056 (2008).

23. Klemm, P. & Schembri, M. A. Bacterial adhesins: function and structure. Int. J. Med.

Microbiol. 290, 27–35 (2000).

24. Kaplan, J. B. Biofilm dispersal: mechanisms, clinical implications, and potential

therapeutic uses. J. Dent. Res. 89, 205–218 (2010).

25. Wolska, K. I., Grudniak, A. M., Rudnicka, Z. & Markowska, K. Genetic control of

bacterial biofilms. J. Appl. Genet. 57, 225–238 (2016).

26. Hughes, D. T. & Sperandio, V. Inter-kingdom signalling: communication between

bacteria and their hosts. Nat. Rev. Microbiol. 6, 111–120 (2008).

27. Chun, C. K., Ozer, E. A., Welsh, M. J., Zabner, J. & Greenberg, E. P. Inactivation of a

Pseudomonas aeruginosa quorum-sensing signal by human airway epithelia. Proc.

Natl. Acad. Sci. 101, 3587–3590 (2004).

28. Stewart, P. S. & Franklin, M. J. Physiological heterogeneity in biofilms. Nat. Rev.

Microbiol. 6, 199–210 (2008).

29. Keren, I., Kaldalu, N., Spoering, A., Wang, Y. & Lewis, K. Persister cells and

tolerance to antimicrobials. FEMS Microbiol. Lett. 230, 13–18 (2004).

30. Dowds, B. C. A. The oxidative stress response in Bacillus subtilis. FEMS Microbiol.

164

Lett. 124, 255–263 (1994).

31. Madsen, J. S., Burmølle, M., Hansen, L. H. & Sørensen, S. J. The interconnection

between biofilm formation and horizontal gene transfer. FEMS Immunology and

Medical Microbiology vol. 65 183–195 (2012).

32. Orlandi, V. T., Bolognese, F., Chiodaroli, L., Tolker-Nielsen, T. & Barbieri, P.

Pigments influence the tolerance of pseudomonas aeruginosa PAO1 to

photodynamically induced oxidative stress. Microbiol. (United Kingdom) 161,

2298–2309 (2015).

33. Setlow, P. Spores of Bacillus subtilis: Their resistance to and killing by radiation,

heat and chemicals. J. Appl. Microbiol. 101, 514–525 (2006).

34. Jakubowski, W. & Walkowiak, B. Resistance of oxidative stress in biofilm and

planktonic cells. Brazilian Arch. Biol. Technol. 58, 300–308 (2015).

35. Miller, R. V. Study of the Response of a Biofilm Bacterial Community to UV

Radiation. Appl. Environ. Microbiol. 65, 2025–2031 (1999).

36. Svenningsen, N. B., Martínez-García, E., Nicolaisen, M. H., de Lorenzo, V. & Nybroe,

O. The biofilm matrix polysaccharides cellulose and alginate both protect

Pseudomonas putida mt-2 against reactive oxygen species generated under matric

stress and copper exposure. Microbiol. (United Kingdom) 164, 883–888 (2018).

37. Dai, T. et al. Blue light rescues mice from potentially fatal pseudomonas aeruginosa

burn infection: Efficacy, safety, and mechanism of action. Antimicrob. Agents

Chemother. 57, 1238–1245 (2013).

38. Wagner, V. E. & Iglewski, B. H. P. aeruginosa biofilms in CF infection. Clin. Rev.

Allergy Immunol. 35, 124–134 (2008).

165

39. Yu, C., Wu, J., Contreras, A. E. & Li, Q. Control of nanofiltration membrane

biofouling by Pseudomonas aeruginosa using d-tyrosine. J. Memb. Sci. 423–424,

487–494 (2012).

40. Altaf, M. M., Ahmad, I., Khan, M. S. A. & Grohmann, E. Bacillus biofilms and their

role in plant health. Biofilms in Plant and Soil Health 55–67 (John Wiley & Sons,

Ltd, 2017).

41. Benamara, H. et al. Characterization of Membrane Lipidome Changes in

Pseudomonas aeruginosa during Biofilm Growth on Glass Wool. PLoS One 9,

e108478 (2014).

42. Ha, D.-G. & O’Toole, G. A. C-di-GMP and its effects on biofilm formation and

dispersion: a Pseudomonas Aeruginosa review. Microbiol. Spectr. 3, 1–12 (2015).

43. Valentini, M. & Filloux, A. Biofilms and cyclic di-GMP (c-di-GMP) signaling: lessons

from Pseudomonas aeruginosa and other bacteria. J. Biol. Chem. 291, 12547–

12555 (2016).

44. Boyd, C. D. & O’Toole, G. A. Second messenger regulation of biofilm formation:

Breakthroughs in understanding c-di-GMP effector systems. Annu. Rev. Cell Dev.

Biol. 28, 439–462 (2012).

45. Chua, S. L. et al. C-di-GMP regulates Pseudomonas aeruginosa stress response to

tellurite during both planktonic and biofilm modes of growth. Nat. Publ. Gr. 1–13

(2015).

46. Römling, U., Galperin, M. Y. & Gomelsky, M. Cyclic di-GMP: the first 25 years of a

universal bacterial second messenger. Microbiol. Mol. Biol. Rev. 77, 1–52 (2013).

47. Hay, I. D., Remminghorst, U. & Rehm, B. H. A. MucR, a novel membrane-associated

regulator of alginate biosynthesis in Pseudomonas aeruginosa. Appl. Environ.

166

Microbiol. 75, 1110–1120 (2009).

48. Malone, J. G. et al. The structure–function relationship of WspR, a Pseudomonas

fluorescens response regulator with a GGDEF output domain. Microbiology 153,

980–994 (2007).

49. De, N., Navarro, M. V. A. S., Wang, Q., Krasteva, P. V & Sondermann, H. B. T.-M. in

E. Biophysical Assays for Protein Interactions in the Wsp Sensory System and

Biofilm Formation. Methods in Enzymology: Two-Component Signaling Systems,

Part C vol. 471 161–184 (Academic Press, 2010).

50. De, N. et al. Phosphorylation-independent regulation of the diguanylate cyclase

WspR. PLOS Biol. 6, e67 (2008).

51. Marvasi, M., Visscher, P. T. & Casillas Martinez, L. Exopolymeric substances (EPS)

from Bacillus subtilis: Polymers and genes encoding their synthesis. FEMS

Microbiology Letters vol. 313 1–9 (2010).

52. Lundberg, M. E., Becker, E. C. & Choe, S. MstX and a putative potassium channel

facilitate biofilm formation in Bacillus subtilis. PLoS One 8, e60993 (2013).

53. Badri, D. V. et al. Galactose metabolism plays a crucial role in biofilm formation by

Bacillus subtilis. MBio 3, 1–10 (2013).

54. Dogsa, I., Brloznik, M., Stopar, D. & Mandic-Mulec, I. Exopolymer diversity and the

role of levan in Bacillus subtilis biofilms. PLoS One 8, 2–11 (2013).

55. Ghafoor, A., Hay, I. D. & Rehm, B. H. A. Role of exopolysaccharides in Pseudomonas

aeruginosa biofilm formation and architecture. Appl. Environ. Microbiol. 77, 5238–

5246 (2011).

56. Cairns, L. S., Hobley, L. & Stanley-Wall, N. R. Biofilm formation by Bacillus subtilis:

167

New insights into regulatory strategies and assembly mechanisms. Mol. Microbiol.

93, 587–598 (2014).

57. Romero, D., Aguilar, C., Losick, R. & Kolter, R. Amyloid fibers provide structural

integrity to Bacillus subtilis biofilms. Proc. Natl. Acad. Sci. 107, 2230–2234 (2010).

58. Terra, R., Stanley-Wall, N. R., Cao, G. & Lazazzera, B. A. Identification of bacillus

subtilis sipw as a bifunctional signal peptidase that controls surface-adhered

biofilm formation. J. Bacteriol. 194, 2781–2790 (2012).

59. Hentzer, M. et al. Alginate overproduction affects Pseudomonas aeruginosa

biofilm structure and function. J. Bacteriol. 183, 5395–5401 (2001).

60. Ma, L. et al. Assembly and development of the Pseudomonas aeruginosa biofilm

matrix. PLOS Pathog. 5, e1000354 (2009).

61. Friedman, L. & Kolter, R. Genes involved in matrix formation in Pseudomonas

aeruginosa PA14 biofilms. Mol. Microbiol. 51, 675–690 (2004).

62. Koprowski, P. & Kubalski, A. Bacterial ion channels and their eukaryotic

homologues. BioEssays vol. 23 1148–1158 (2001).

63. Saimi, Y., Loukin, S. H., Zhou, X. L., Martinac, B. & Kung, C. Ion channels in

microbes. Methods Enzymol. 294, 507–524 (1998).

64. Treptow, W. & Tarek, M. K+ conduction in the selectivity filter of potassium

channels is monitored by the charge distribution along their sequence. Biophys. J.

91, L81–L83 (2006).

65. Doyle, D. A. The structure of the potassium channel: molecular basis of K+

conduction and selectivity. Science (80-. ). 280, 69–77 (1998).

66. Payandeh, J., Scheuer, T., Zheng, N. & Catterall, W. A. The crystal structure of a

168

voltage-gated sodium channel. Nature 475, 353–358 (2011).

67. Hou, X., Pedi, L., Diver, M. M. & Long, S. B. Crystal structure of the calcium release-

activated Calcium channel Orai. Science (80-. ). 338, 1308–1313 (2012).

68. Sansom, M. S., Shrivastava, I. H., Ranatunga, K. M. & Smith, G. R. Simulations of ion

channels--watching ions and water move. Trends Biochem. Sci. 25, 368–74 (2000).

69. Kuang, Q., Purhonen, P. & Hebert, H. Structure of potassium channels. Cell. Mol.

Life Sci. 72, 3677–93 (2015).

70. Miller, J. B. & Koshland, D. E. Sensory electrophysiology of bacteria: relationship of

the membrane potential to motility and chemotaxis in Bacillus subtilis. Proc. Natl.

Acad. Sci. U. S. A. 74, 4752–6 (1977).

71. Reza, S. N. Mechanosensitive ion channels. Usp. Fiziol. Nauk 33, 29–37 (2002).

72. Edwards, M. D. et al. Characterization of three novel mechanosensitive channel

activities in Escherichia coli. Channels (Austin). 6, 272–281 (2012).

73. Sukharev, S. I., Blount, P., Martinac, B., Blattner, F. R. & Kung, C. A large-

conductance mechanosensitive channel in E. coli encoded by mscL alone. Nature

368, 265–268 (1994).

74. Wright, S. H. Generation of resting membrane potential. Adv. Physiol. Educ. 28,

139 LP-142 (2004).

75. Arevalo, A. & Pastor, G. Verification of the Nernst equation and determination of a

standard electrode potential. J. Chem. Educ. 62, 882 (1985).

76. Strahl, H. & Hamoen, L. W. Membrane potential is important for bacterial cell

division. Proc. Natl. Acad. Sci. U. S. A. 107, 12281–6 (2010).

77. Blackiston, D. J., McLaughlin, K. A. & Levin, M. Bioelectric controls of cell

169

proliferation: ion channels, membrane voltage and the cell cycle. Cell cycle

(Georgetown, Tex.) vol. 8 3519–3528 (2009).

78. Pizarro-Cerdá, J. & Cossart, P. Bacterial adhesion and entry into host cells. Cell vol.

124 715–727 (2006).

79. Sirec, T., Benarroch, J. M., Buffard, P., Garcia-Ojalvo, J. & Asally, M. Electrical

polarization enables integrative quality control during bacterial differentiation into

spores. iScience 16, 378–389 (2019).

80. Kralj, J., Hochbaum, D., Douglass, A. & Cohen, A. Electrical spiking in Escherichia

coli. Science (80-. ). 333, 345–348 (2011).

81. Stratford, J. P. et al. Electrically induced bacterial membrane-potential dynamics

correspond to cellular proliferation capacity. Proc. Natl. Acad. Sci. 116, 201901788

(2019).

82. Del Pozo, J. L. et al. The electricidal effect is active in an experimental model of

Staphylococcus epidermidis chronic foreign body osteomyelitis. Antimicrob. Agents

Chemother. 53, 4064–4068 (2009).

83. Shirtliff, M. E., Bargmeyer, A. & Camper, A. K. Assessment of the ability of the

bioelectric effect to eliminate mixed-species biofilms. Appl. Environ. Microbiol. 71,

6379–6382 (2005).

84. Masi, E. et al. Electrical spiking in bacterial biofilms. J. R. Soc. Interface 12,

20141036 (2015).

85. Prindle, A. et al. Ion channels enable electrical communication in bacterial

communities. Nature 527, 59–63 (2015).

86. Liu, J. et al. Coupling between distant biofilms and emergence of nutrient time-

170

sharing. Science (80-. ). 356, 638–642 (2017).

87. Humphries, J. et al. Species-independent attraction to biofilms through electrical

signaling article species-independent attraction to biofilms through electrical

signaling. Cell 168, 200–209 (2017).

88. Larkin, J. W. et al. Signal percolation within a bacterial community. Cell Syst. 7,

137–145.e3 (2018).

89. Renz, M. Fluorescence microscopy—A historical and technical perspective. Cytom.

Part A 83, 767–779 (2013).

90. Valeur, B. & Berberan-Santos, M. N. Molecular fluorescence. (Wiley-VCH Verlag

GmbH & Co. KGaA, 2012).

91. Diaspro, A. Optical fluorescence microscopy: from the spectral to the nano

dimension. (Springer Berlin Heidelberg, 2010).

92. Lakowicz, J. R. Principles of fluorescence spectroscopy. (Springer, 2006).

93. Mamontova, A. V, Grigoryev, A. P., Tsarkova, A. S., Lukyanov, K. A. & Bogdanov, A.

M. Struggle for photostability: Bleaching mechanisms of fluorescent proteins. Russ.

J. Bioorganic Chem. 43, 625–633 (2017).

94. Alnuami, A. A., Zeedi, B., Qadri, S. M. & Ashraf, S. S. Oxyradical-induced GFP

damage and loss of fluorescence. Int. J. Biol. Macromol. 43, 182–186 (2008).

95. Lichtman, J. W. & Conchello, J.-A. Fluorescence microscopy. Nat. Methods 2, 910–

919 (2005).

96. Biancalana, M. & Koide, S. Molecular mechanism of Thioflavin-T binding to amyloid

fibrils. Biochim. Biophys. Acta - Proteins Proteomics 1804, 1405–1412 (2010).

97. Lo, C.-J., Leake, M. C. & Berry, R. M. Fluorescence measurement of intracellular

171

sodium concentration in single Escherichia coli cells. Biophys. J. 90, 357–365

(2006).

98. Yuste, R. Fluorescence microscopy today. Nat. Methods 2, 902–904 (2005).

99. Remington, S. J. Green fluorescent protein: a perspective. Protein Sci. 20, 1509–

1519 (2011).

100. Zimmer, M. Green fluorescent protein (GFP): applications, structure, and related

photophysical behavior. Chem. Rev. 102, 759–782 (2002).

101. Axelrod, D., Koppel, D. E., Schlessinger, J., Elson, E. & Webb, W. W. Mobility

measurement by analysis of fluorescence photobleaching recovery kinetics.

Biophys. J. 16, 1055–1069 (1976).

102. Wüstner, D. et al. Quantitative fluorescence loss in photobleaching for analysis of

protein transport and aggregation. BMC Bioinformatics 13, 296 (2012).

103. Alhede, M. et al. Combination of microscopic techniques reveals a comprehensive

visual impression of biofilm structure and composition. FEMS Immunol. Med.

Microbiol. 65, 335–342 (2012).

104. Xu, J., Ma, H. & Liu, Y. Stochastic Optical Reconstruction Microscopy (STORM).

Curr. Protoc. Cytom. 81, 12.46.1-12.46.27 (2017).

105. de Jong, I. G., Beilharz, K., Kuipers, O. P. & Veening, J.-W. Live cell imaging of

Bacillus subtilis and Streptococcus pneumoniae using automated time-lapse

microscopy. J. Vis. Exp. e3145 (2011).

106. Elias, S. & Banin, E. Multi-species biofilms: living with friendly neighbors. FEMS

Microbiol. Rev. 36, 990–1004 (2012).

107. Bridier, A. et al. Biofilms of a Bacillus subtilis hospital isolate protect

172

Staphylococcus aureus from biocide action. PLoS One 7, e44506 (2012).

108. Sanchez-Vizuete, P. et al. Identification of ypqP as a new Bacillus subtilis biofilm

determinant that mediates the protection of Staphylococcus aureus against

antimicrobial agents in mixed-species communities. Appl. Environ. Microbiol. 81,

109–118 (2015).

109. Azeredo, J. et al. Critical review on biofilm methods. Crit. Rev. Microbiol. 43, 313–

351 (2017).

110. Davey, M. E. & O’toole, G. A. Microbial biofilms: from ecology to molecular

genetics. Microbiol. Mol. Biol. Rev. 64, 847–867 (2000).

111. Nivens, D. E., Palmer, R. J. & White, D. C. Continuous nondestructive monitoring of

microbial biofilms: A review of analytical techniques. J. Ind. Microbiol. 15, 263–276

(1995).

112. Whitesides, G. M. The origins and the future of microfluidics. Nature 442, 368–373

(2006).

113. Su, F., Chakrabarty, K. & Fair, R. B. Microfluidics-based biochips: Technology issues,

implementation platforms, and design-automation challenges. IEEE Trans. Comput.

Des. Integr. Circuits Syst. 25, 211–223 (2006).

114. Merkel, T. C., Bondar, V. I., Nagai, K., Freeman, B. D. & Pinnau, I. Gas sorption,

diffusion, and permeation in poly(dimethylsiloxane). J. Polym. Sci. Part B Polym.

Phys. 38, 415–434 (2000).

115. Łopacińska, J. M., Emnéus, J. & Dufva, M. Poly(Dimethylsiloxane) (PDMS) affects

gene expression in PC12 cells differentiating into neuronal-like cells. PLoS One 8,

e53107 (2013).

173

116. Regehr, K. J. et al. Biological implications of polydimethylsiloxane-based

microfluidic cell culture. Lab Chip 9, 2132 (2009).

117. Paguirigan, A. L. & Beebe, D. J. From the cellular perspective: exploring differences

in the cellular baseline in macroscale and microfluidic cultures. Integr. Biol. 1, 182

(2009).

118. Stone, H. A. & Kim, S. Microfluidics: Basic issues, applications, and challenges.

AIChE Journal vol. 47 1250–1254 (2001).

119. Ottino, J. M. & Wiggins, S. Introduction: mixing in microfluidics. Philos. Trans. R.

Soc. A Math. Phys. Eng. Sci. 362, 923–935 (2004).

120. Tarn, M. D. & Pamme, N. Microfluidics. Ref. Modul. Chem. Mol. Sci. Chem. Eng. 1–7

(2013).

121. Singh, N. et al. Dual bioresponsive antibiotic and quorum sensing inhibitor

combination nanoparticles for treatment of Pseudomonas aeruginosa biofilms in

vitro and ex vivo. Biomater. Sci. 7, 4099–4111 (2019).

122. Fensterseifer, I. C. M. et al. Selective antibacterial activity of the cationic peptide

PaDBS1R6 against Gram-negative bacteria. Biochim. Biophys. Acta - Biomembr.

1861, 1375–1387 (2019).

123. Rinzel, J. Excitation dynamics: insights from simplified membrane models.

Federation proceedings vol. 44 2944–2946 (1985).

124. Keener, J. & Sneyd, J. Wave propagation in excitable systems - mathematical

physiology: I: cellular physiology. (eds. Keener, J. & Sneyd, J.) 229–271 (Springer

New York, 2009).

125. FitzHugh, R. Impulses and physiological states in theoretical models of nerve

174

membrane. Biophys. J. 1, 445–466 (1961).

126. Vandenberg, J. I. & Waxman, S. G. Hodgkin and Huxley and the basis for electrical

signalling: A remarkable legacy still going strong. J. Physiol. 590, 2569–2570 (2012).

127. Hodgkin, A. L. & Huxley, A. Hodgkin – Huxley model. Medicine (Baltimore). 1–4

(1963).

128. Bashir, Y., Betts, T. R. & Rajappan, K. Introduction to cardiac electrophysiology–a

brief historical perspective. OSH Cardiac Electrophysiology and Catheter Ablation

2–6 (Oxford University Press, 2010).

129. FitzHugh, R. Thresholds and plateaus in the Hodgkin-Huxley nerve equations. J.

Gen. Physiol. 43, 867–96 (1960).

130. Fitzhugh, R. Mathematical models of excitation and propagation in nerve.

Biological Engineering vol. 34 1–85 (1969).

131. Hassard, B. Bifurcation of periodic solutions of the Hodgkin-Huxley model for the

squid giant axon. J. Theor. Biol. 71, 401–420 (1978).

132. Sabah, N. H. & Spangler, R. A. Repetitive response of the Hodgkin-Huxley model for

the squid giant axon. J. Theor. Biol. 29, 155–171 (1970).

133. Cross, M. & Greenside, H. Pattern formation and dynamics in nonequilibrium

systems. (Cambridge University Press, 2009).

134. Herzog, W. Skeletal muscle mechanics: questions, problems and possible solutions.

J. Neuroeng. Rehabil. 14, 98 (2017).

135. Saxena, R. K., Mathai, A. M. & Haubold, H. J. Solution of generalized fractional

reaction-diffusion equations. Astrophys. Space Sci. 305, 305 (2006).

136. Bates, P. W., Lu, K. & Wang, B. Random attractors for stochastic reaction–diffusion

175

equations on unbounded domains. J. Differ. Equ. 246, 845–869 (2009).

137. He, J.-H. Α review on some new recently developed nonlinear analytical

techniques. Int. J. Nonlinear Sci. Numer. Simul. 1, 51 (2000).

138. Ziegel, E., Press, W., Flannery, B., Teukolsky, S. & Vetterling, W. Numerical Recipes:

The Art of Scientific Computing. Technometrics vol. 29 (Cambridge University Press,

1987).

139. Xavier, J. de B., Picioreanu, C. & van Loosdrecht, M. C. M. A general description of

detachment for multidimensional modelling of biofilms. Biotechnol. Bioeng. 91,

651–669 (2005).

140. Allen, R. J. & Waclaw, B. Bacterial growth: a statistical physicist’s guide. Rep. Prog.

Phys. 82, 16601 (2019).

141. Kreft, J.-U., Picioreanu, C., van Loosdrecht, M. C. M. & Wimpenny, J. W. T.

Individual-based modelling of biofilms. Microbiology 147, 2897–2912 (2001).

142. Martinez-Corral, R., Liu, J., Prindle, A., Süel, G. M. & Garcia-Ojalvo, J. Metabolic

basis of brain-like electrical signalling in bacterial communities. Philos. Trans. R.

Soc. B Biol. Sci. 374, 20180382 (2019).

143. Mikami, T., Asally, M., Kano, T. & Ishiguro, A. A reaction-diffusion model for

simulating the oscillatory expansion of biofilms. in The 2019 Conference on

Artificial Life 218–219 (MIT Press, 2019).

144. Volfson, D., Cookson, S., Hasty, J. & Tsimring, L. S. Biomechanical ordering of dense

cell populations. Proc. Natl. Acad. Sci. 105, 15346–15351 (2008).

145. Martinez-Corral, R., Liu, J., Süel, G. M. & Garcia-Ojalvo, J. Bistable emergence of

oscillations in growing Bacillus subtilis biofilms. Proc. Natl. Acad. Sci. 115, E8333--

176

E8340 (2018).

146. Keener, J. & Sneyd, J. Calcium dynamics. Mathematical physiology: I: cellular

physiology 273–346 (Springer New York, 2010).

147. Gutiérrez, M. et al. A new improved and extended version of the multicell bacterial

simulator gro. ACS Synth. Biol. 6, 1496–1508 (2017).

148. Verkhratsky, A. & Parpura, V. History of Electrophysiology and the Patch Clamp.

Patch-Clamp Methods and Protocols (eds. Martina, M. & Taverna, S.) 1–19

(Springer New York, 2014).

149. Vagos, M. R. S. S. et al. Computational modeling of electrophysiology and

pharmacotherapy of atrial fibrillation: Recent advances and future challenges.

Front. Physiol. 9, 1–29 (2018).

150. Cao, Y. et al. Gating of the TrkH ion channel by its associated RCK protein TrkA.

Nature 496, 317–322 (2013).

151. Chen, G.-Q., Cul, C., MAyer, M. L. & Gouaux, E. Functional characterization of a

potassium-selective prokaryotic glutamate receptor. Nat. Biotechnol. 402, 817–821

(1999).

152. Beagle, S. D. & Lockless, S. W. Electrical signalling goes bacterial. Nature 527, 44–

45 (2015).

153. Filosa, J. A. et al. Local potassium signaling couples neuronal activity to

vasodilation in the brain. Nat. Neurosci. 9, 1397–1403 (2006).

154. Gerstner, W. & Kistler, W. M. Spiking neuron models. (Cambridge University Press,

2002).

155. Hodgkin, A. L. & Huxley, A. F. A quantitative description of membrane current and

177

its application to conduction and excitation in nerve. J. Physiol. 117, 500–544

(1952).

156. Glen, C. M., Kemp, M. L. & Voit, E. O. Agent-based modeling of morphogenetic

systems: Advantages and challenges. PLOS Comput. Biol. 15, e1006577 (2019).

157. Bora, Ş. & Emek, S. Agent-Based modeling and simulation of biological systems.

Modeling and Computer Simulation Ch. 3 (IntechOpen, 2019).

158. Ferrante, E., Turgut, A. E., Dorigo, M. & Huepe, C. Elasticity-based mechanism for

the collective motion of self-propelled particles with springlike interactions: a

model system for natural and artificial swarms. Phys. Rev. Lett. 111, 268302 (2013).

159. Weber, C. A., Bock, C. & Frey, E. Defect-mediated phase transitions in active soft

matter. Phys. Rev. Lett. 112, 168301 (2014).

160. Branda, S. S., González-Pastor, J. E., Ben-Yehuda, S., Losick, R. & Kolter, R. Fruiting

body formation by Bacillus subtilis. Proc. Natl. Acad. Sci. U. S. A. 98, 11621–6

(2001).

161. Mohammed, B. S., Fields, D. A., Mittendorfer, B., Coggan, A. R. & Klein, S. Are

peristaltic pumps as reliable as syringe pumps for metabolic research? assessment

of accuracy, precision, and metabolic kinetics. Metabolism. 53, 875–878 (2004).

162. Mirtich, B. & Canny, J. Impulse-based simulation of rigid bodies. in Proceedings of

the 1995 symposium on Interactive 3D graphics - SI3D ’95 181- (ACM Press, 1995).

163. Ericson, C. Real-time collision detection : Christer Ericson. (Elsevier : Morgan

Kaufmann, 2005).

164. Jafri, M. S. & Keizer, J. On the roles of Ca2+ diffusion, Ca2+ buffers, and the

endoplasmic reticulum in IP3-induced Ca2+ waves. Biophys. J. 69, 2139–2153

178

(1995).

165. Nagy‐Ungvarai, Z., Ungvarai, J., Müller, S. C. & Hess, B. The role of curvature and

pulse width for transition to unstable wave fronts in the Belousov–Zhabotinsky

reaction. J. Chem. Phys. 97, 1004–1009 (1992).

166. Herzog, W. Skeletal muscle mechanics: questions, problems and possible solutions.

J. Neuroeng. Rehabil. 14, 98 (2017).

167. Bers, D. M. Calcium and Cardiac Rhythms. Circ. Res. 90, 14–17 (2002).

168. Stewart, P. S. Diffusion in biofilms. Journal of Bacteriology vol. 185 1485–1491

(2003).

169. Klavins, E. A language for modeling and programming cooperative control systems.

in IEEE International Conference on Robotics and Automation, 2004. Proceedings.

ICRA ’04. 2004 vol. 4 3403–3410 Vol.4 (2004).

170. Toole, G. & Hurdal, M. K. Turing models of cortical folding on exponentially and

logistically growing domains. Comput. Math. with Appl. 66, 1627–1642 (2013).

171. Ramakrishnan, P., Maclean, M., MacGregor, S. J., Anderson, J. G. & Grant, M. H.

Cytotoxic responses to 405nm light exposure in mammalian and bacterial cells:

Involvement of reactive oxygen species. Toxicol. Vitr. 33, 54–62 (2016).

172. Maclean, M., Anderson, J. G., MacGregor, S. J., White, T. & Atreya, C. D. A new

proof of concept in bacterial reduction: antimicrobial action of violet-blue light

(405 nm) in Ex Vivo stored plasma. J. Blood Transfus. 2016, 1–11 (2016).

173. Dai, T., Vrahas, M. S., Murray, C. K. & Hamblin, M. R. Ultraviolet C irradiation: An

alternative antimicrobial approach to localized infections? Expert Rev. Anti. Infect.

Ther. 10, 185–195 (2012).

179

174. Peker, I., Akca, G., Sarikir, C., Toraman Alkurt, M. & Celik, I. Effectiveness of

alternative methods for toothbrush disinfection: An in vitro study. Sci. World J.

2014, 1–9 (2014).

175. Yuan, Y. et al. Photodynamic antimicrobial chemotherapy with the novel amino

acid-porphyrin conjugate 4I: In vitro and in vivo studies. PLoS One 12, e0176529

(2017).

176. Hamblin, M. R. et al. Helicobacter pylori. Society 49, 2822–2827 (2005).

177. MacLean, M., Murdoch, L. E., MacGregor, S. J. & Anderson, J. G. Sporicidal effects

of high-intensity 405 nm visible light on endospore-forming bacteria. Photochem.

Photobiol. 89, 120–126 (2013).

178. McKenzie, K. et al. The effects of 405 nm light on bacterial membrane integrity

determined by salt and bile tolerance assays, leakage of UV-absorbing material and

SYTOX green labelling. Microbiology 162, 1680–1688 (2016).

179. Chiaviello, A., Postiglione, I. & Palumbo, G. Targets and mechanisms of

photodynamic therapy in lung cancer cells: a brief overview. Cancers (Basel). 3,

1014–1041 (2011).

180. Pattison, D. I. & Davies, M. J. Actions of ultraviolet light on cellular structures.

Cancer: Cell Structures, Carcinogens and Genomic Instability. 131–157 (Birkhäuser

Basel, 2006).

181. Sharma, P. & Dubey, R. S. Drought induces oxidative stress and enhances the

activities of antioxidant enzymes in growing rice seedlings. Plant Growth Regul. 46,

209–221 (2005).

182. Meriga, B., Reddy, B. K., Rao, K. R., Reddy, L. A. & Kishor, P. B. K. Aluminium-

induced production of oxygen radicals, lipid peroxidation and DNA damage in

180

seedlings of rice (Oryza sativa). J. Plant Physiol. 161, 63–68 (2004).

183. Mittler, R. Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci. 7,

405–410 (2002).

184. Huynh, T. T. et al. Glucose starvation-induced dispersal of pseudomonas

aeruginosa biofilms is camp and energy dependent. PLoS One 7, e42874 (2012).

185. Moghoofei, M. et al. Morphological and bactericidal effects of amikacin,

meropenem and imipenem on Pseudomonas aeruginosa. Jundishapur J. Microbiol.

8, (2015).

186. Aalen, O. O., Borgan, Ø. & Gjessing, H. K. Survival and Event History Analysis - A

Process Point of View. Springer (2008).

187. Kaplan, E. L. & Meier, P. Nonparametric estimation from incomplete observations.

J. Am. Stat. Assoc. 53, 457–481 (1958).

188. Barraclough, H., Simms, L. & Govindan, R. Biostatistics primer: What a clinician

ought to know: Hazard ratios. J. Thorac. Oncol. 6, 978–982 (2011).

189. Cook, R. J. & Lawless, J. F. Event history analysis. International Encyclopedia of the

Social & Behavioral Sciences (Elsevier, 2015).

190. Schlesinger, D. et al. Equivalence of cell survival data for radiation dose and

thermal dose in ablative treatments: analysis applied to essential tremor

thalamotomy by focused ultrasound and gamma knife. Int. J. Hyperth. 33, 401–410

(2017).

191. Liu, J. et al. Metabolic co-dependence gives rise to collective oscillations within

biofilms. Nature 523, 550–554 (2015).

192. Locke, J. C. W., Young, J. W., Fontes, M., Jimenez, M. J. H. & Elowitz, M. B.

181

Stochastic pulse regulation in bacterial stress response. Science (80-. ). 334, 366–

369 (2011).

193. Alon, U. An introduction to systems biology: Design principles of biological circuits.

Chapman Hall/CRC (2007).

194. Elasri, M. O. & Miller, R. V. Study of the response of a biofilm bacterial community

to UV radiation. 65, 2025–2031 (1999).

195. Stover, C. K. et al. Complete genome sequence of Pseudomonas aeruginosa PAO1,

an opportunistic pathogen. Nature 406, 959–64 (2000).

196. Gomelsky, M. & Hoff, W. D. Light helps bacteria make important lifestyle decisions.

Trends Microbiol. 19, 441–448 (2011).

197. Braatsch, S. & Klug, G. Blue light perception in bacteria. Photosynth. Res. 79, 45–57

(2004).

198. Conrad, K. S., Manahan, C. C. & Crane, B. R. Photochemistry of flavoprotein light

sensors. Nat. Chem. Biol. 10, 801–809 (2014).

199. Dai, T. et al. Blue light for infectious diseases: Propionibacterium acnes,

Helicobacter pylori, and beyond? Drug Resist. Updat. 15, 233–236 (2012).

200. Hecker, M., Pané-Farré, J. & Uwe, V. SigB-dependent general stress response in

Bacillus subtilis and related gram-positive bacteria. Annu. Rev. Microbiol. 61, 215–

236 (2007).

201. Ávila-Pérez, M., Hellingwerf, K. J. & Kort, R. Blue light activates the σB dependent

stress response of Bacillus subtilis via YtvA. J. Bacteriol. 188, 6411 LP-6414 (2006).

202. Mukherjee, S., Jemielita, M., Stergioula, V., Tikhonov, M. & Bassler, B. L.

Photosensing and quorum sensing are integrated to control Pseudomonas

182

aeruginosa collective behaviors. PLOS Biol. 17, e3000579 (2019).

203. Wu, L., McGrane, R. S. & Beattie, G. A. Light regulation of swarming motility in

Pseudomonas syringae integrates signaling pathways mediated by a

bacteriophytochrome and a LOV protein. MBio 4, 1–9 (2013).

204. Terauchi, K. & Ohmori, M. Blue light stimulates cyanobacterial motility via a cAMP

signal transduction system. Mol. Microbiol. 52, 303–309 (2004).

205. Chua, S. L. et al. C-di-GMP regulates Pseudomonas aeruginosa stress response to

tellurite during both planktonic and biofilm modes of growth. Sci. Rep. 5, 10052

(2015).

206. Bruni, G. N., Weekley, R. A., Dodd, B. J. T. & Kralj, J. M. Voltage-gated calcium flux

mediates Escherichia coli mechanosensation. Proc. Natl. Acad. Sci. 114, 9445–9450

(2017).

207. Huynh, T. T. et al. Glucose starvation-induced dispersal of pseudomonas

aeruginosa biofilms is camp and energy dependent. PLoS One 7, (2012).

208. Lin Chua, S. et al. Reduced intracellular c-di-GMP content increases expression of

quorum sensing-Regulated genes in Pseudomonas aeruginosa. Front. Cell. Infect.

Microbiol. 7, 451 (2017).

209. Strempel, N., Nusser, M., Neidig, A., Brenner-Weiss, G. & Overhage, J. The

oxidative stress agent hypochlorite stimulates c-di-GMP synthesis and biofilm

formation in Pseudomonas aeruginosa. Front. Microbiol. 8, 2311 (2017).

210. Huang, C.-J., Wang, Z.-C., Huang, H.-Y., Huang, H.-D. & Peng, H.-L. YjcC, a c-di-GMP

phosphodiesterase protein, regulates the oxidative stress response and virulence

of Klebsiella pneumoniae CG43. PLoS One 8, e66740–e66740 (2013).

183

211. Fernandez, N. L. & Waters, C. M. Cyclic di-GMP increases catalase production and

hydrogen peroxide tolerance in Vibrio cholerae. Appl. Environ. Microbiol. 85,

e01043-19 (2019).

212. Plyuta, V. A., Andreenko, J. V, Kuznetsov, A. E. & Khmel’, I. A. Formation of

Pseudomonas aeruginosa PAO1 biofilms in the presence of hydrogen peroxide. The

effect of the aiiA gene. Mol. Genet. Microbiol. Virol. 28, 141–146 (2013).

213. Irie, Y. & Parsek, M. LC/MS/MS-based quantitative assay for the secondary

messenger molecule, c-di-GMP. Methods Mol. Biol. 1149, 271–279 (2014).

214. Rybtke, M. T. et al. Fluorescence-based reporter for gauging cyclic di-GMP Levels in

Pseudomonas aeruginosa. Appl. Environ. Microbiol. 78, 5060–5069 (2012).

215. Sinkeldam, R. W., Greco, N. J. & Tor, Y. Fluorescent analogs of biomolecular

building blocks: Design, properties, and applications. Chem. Rev. 110, 2579–2619

(2010).

216. Barbier, M. & Damron, F. H. Rainbow vectors for broad-range bacterial

fluorescence labeling. PLoS One 11, e0146827 (2016).

217. Koch, B., Jensen, L. E. & Nybroe, O. A panel of Tn7-based vectors for insertion of

the gfp marker gene or for delivery of cloned DNA into Gram-negative bacteria at a

neutral chromosomal site. J. Microbiol. Methods 45, 187–195 (2001).

218. Clark, D. J. & Maaløe, O. DNA replication and the division cycle in Escherichia coli. J.

Mol. Biol. 23, 99–112 (1967).

219. Choi, K.-H., Kumar, A. & Schweizer, H. P. A 10-min method for preparation of highly

electrocompetent Pseudomonas aeruginosa cells: Application for DNA fragment

transfer between chromosomes and plasmid transformation. J. Microbiol. Methods

64, 391–397 (2006).

184

220. Chua, S. L. et al. Reactive oxygen species drive evolution of pro-biofilm variants in

pathogens by modulating cyclic-di-GMP levels. Open Biol. 6, 160162 (2016).

221. Rodesney, C. A. et al. Mechanosensing of shear by Pseudomonas aeruginosa leads

to increased levels of the cyclic-di-GMP signal initiating biofilm development. Proc.

Natl. Acad. Sci. 114, 5906–5911 (2017).

222. Ganini, D. et al. Fluorescent proteins such as eGFP lead to catalytic oxidative stress

in cells. Redox Biol. 12, 462–468 (2017).

223. Piatkevich, K. D. & Verkhusha, V. V. Guide to red fluorescent proteins and

biosensors for flow cytometry. Methods Cell Biol. 102, 431–461 (2011).

224. Dixit, R. & Cyr, R. Cell damage and reactive oxygen species production induced by

fluorescence microscopy: effect on mitosis and guidelines for non-invasive

fluorescence microscopy. Plant J. 36, 280–290 (2003).

225. Różanowska, M. et al. Blue light-induced singlet oxygen generation by retinal

lipofuscin in non-polar media. Free Radic. Biol. Med. 24, 1107–1112 (1998).

226. Bjarnsholt, T. et al. Pseudomonas aeruginosa tolerance to tobramycin, hydrogen

peroxide and polymorphonuclear leukocytes is quorum-sensing dependent.

Microbiology 151, 373–383 (2005).

227. Elkins, J. G., Hassett, D. J., Stewart, P. S., Schweizer, H. P. & McDermott, T. R.

Protective role of catalase in Pseudomonas aeruginosa biofilm resistance to

hydrogen peroxide. Appl. Environ. Microbiol. 65, 4594–4600 (1999).

228. Bond, D. R. & Lovley, D. R. Electricity production by Geobacter sulfurreducens

attached to electrodes. Appl. Environ. Microbiol. 69, 1548–1555 (2003).

229. Schlafer, S. & Meyer, R. L. Confocal microscopy imaging of the biofilm matrix. J.

185

Microbiol. Methods 138, 50–59 (2017).

230. Murdoch, L. E., Maclean, M., Endarko, E., MacGregor, S. J. & Anderson, J. G.

Bactericidal effects of 405 nm light exposure demonstrated by inactivation of

Escherichia, Salmonella, Shigella, Listeria, and Mycobacterium species in liquid

suspensions and on exposed surfaces. Sci. World J. 2012, 1–8 (2012).

231. Kalyanaraman, B. et al. Measuring reactive oxygen and nitrogen species with

fluorescent probes: challenges and limitations. Free Radic. Biol. Med. 52, 1–6

(2012).

232. Renggli, S., Keck, W., Jenal, U. & Ritz, D. Role of autofluorescence in flow

cytometric analysis of escherichia coli treated with bactericidal antibiotics. J.

Bacteriol. 195, 4067–4073 (2013).

233. McBee, M. E. et al. Production of superoxide in bacteria is stress- and cell state-

dependent: A gating-optimized flow cytometry method that minimizes ROS

measurement artifacts with fluorescent dyes. Front. Microbiol. 8, 459 (2017).

234. Liu, Y. & Imlay, J. A. Cell death from antibiotics without the involvement of reactive

oxygen species. Science (80-. ). 339, 1210–1213 (2013).

235. Oparka, M. et al. Quantifying ROS levels using CM-H2DCFDA and HyPer. Methods

109, 3–11 (2016).

236. Seistrup, K. H., Strahl, H., Hamoen, L. W., Gray, D. A. & te Winkel, J. D. Analysis of

antimicrobial-triggered membrane depolarization using voltage sensitive dyes.

Front. Cell Dev. Biol. 4, 1–10 (2016).

237. Masson-Meyers, D., Bumah, V., Biener, G., Raicu, V. & Enwemeka, C. The relative

antimicrobial effect of blue 405 nm LED and blue 405 nm laser on methicillin-

resistant Staphylococcus aureus in vitro. Lasers Med. Sci. 30, (2015).

186

238. Ganini, D. et al. Fluorescent proteins such as eGFP lead to catalytic oxidative stress

in cells. Redox Biol. 12, 462–468 (2017).

239. Becker, W. Fluorescence lifetime imaging - techniques and applications. J. Microsc.

247, 119–136 (2012).