138
POLITECNICO DI MILANO Department of Physics Doctoral Programme in Physics Mechanisms of cellular photostimulation in hybrid interfaces based on organic semiconductors Supervisor: Dr. Maria Rosa ANTOGNAZZA Tutor: Prof. Guglielmo LANZANI PhD Programme Coordinator: Prof. Paola TARONI Doctoral Dissertation of: Nicola MARTINO 2015 PhD Cycle XXVI

Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

POLITECNICO DI MILANO

Department of Physics

Doctoral Programme in Physics

Mechanisms of cellular photostimulation in hybrid

interfaces based on organic semiconductors

Supervisor: Dr. Maria Rosa ANTOGNAZZA

Tutor: Prof. Guglielmo LANZANI

PhD Programme Coordinator: Prof. Paola TARONI

Doctoral Dissertation of:

Nicola MARTINO

2015 – PhD Cycle XXVI

Page 2: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

i

Abstract

Hybrid interfaces between organic semiconductors and living tissues represent a new tool for in

vitro and in vivo applications, bearing a huge potential, from basic researches to clinical

applications. In particular, light sensitive conjugated polymers can be exploited as a new approach

for optical modulation of cellular activity. This thesis is focused on the study of the functioning

mechanisms of these interfaces, both from the physical point of view and from their ability to

stimulate biological cells. In particular, we are interested in understanding how photoexcitation of

the active material in the device is able to modulate the membrane potential of cells. First, we

review the current strategies used for measuring and controlling bioelectrical activity, with a

particular attention paid to optical techniques, and we introduce the biophysical mechanisms behind

the instauration of a potential across the plasma membrane of cells. We present a thorough

experimental characterization of the hybrid polymer/electrolyte interfaces, in which their

spectroscopic, electrical and thermal properties are investigated, delineating the main phenomena

that occur at the device surface upon illumination. The possibility of growing HEK-293 cells on

these hybrid interfaces is the investigated, and we study the different effects that the device

photoexcitation has on the cell membrane potential via patch-clamp analysis. We conclude by

wrapping up the results in the context of existing techniques for cell stimulation and by pointing out

to future developments, towards the creation of a multi-functional platform for light-controlled cell

manipulation, with possible applications in different fields of neuroscience and medicine.

Page 3: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

ii

Table of contents

Chapter 1 - Bioelectricity ............................................................................................................... 1

1.1 Electrical stimulation and recording ....................................................................... 1

1.2 Optical techniques .................................................................................................. 4

1.2.1 Optical measurement of bioelectric activity ............................................................. 4

1.2.2 Direct optical stimulation ......................................................................................... 5

1.2.3 Molecular-based stimulation .................................................................................... 7

1.2.4 Nano-/micro-particle stimulation .............................................................................. 8

1.2.5 Device-based stimulation .......................................................................................... 9

1.3 Organic semiconductors for biological applications ............................................... 9

1.4 Photoactive bio-polymer interfaces ...................................................................... 11

1.4.1 The hybrid solid-liquid organic photovoltaic cell ................................................... 12

1.4.2 Poly(3-hexylthiophene) .......................................................................................... 13

1.4.3 Photostimulation of primary cells ........................................................................... 14

1.4.4 Ex-vivo experiments on blind retinas ...................................................................... 15

Chapter 2 – The plasma membrane ............................................................................................ 16

2.1 The structure of the plasma membrane ................................................................. 17

2.2 Ion channels ......................................................................................................... 19

2.2.1 Ion channel structure and selectivity ...................................................................... 20

2.2.2 Gating mechanisms ................................................................................................. 21

2.2.3 Channel conductance and temperature dependence ............................................... 22

2.3 The membrane potential ....................................................................................... 23

2.3.1 Electrochemical equilibrium in biological membranes .......................................... 24

2.3.2 Action potentials in excitable cells ......................................................................... 26

2.3.3 Maintenance of ion concentrations ......................................................................... 27

2.3.4 Electrical equivalent of a cell membrane ................................................................ 28

Chapter 3 – Hybrid interfaces characterization ........................................................................ 32

3.1 Standard organic photovoltaic devices .................................................................. 32

3.2 Hybrid devices structure ....................................................................................... 34

Page 4: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

iii

3.3 Spectroscopic characterization ............................................................................. 35

3.3.1 Absorption and fluorescence .................................................................................. 35

3.3.2 Pump and probe spectroscopies .............................................................................. 36

3.3.3 Femtosecond transient absorption spectroscopy ..................................................... 37

3.3.4 Nanosecond transient absorption spectroscopy ...................................................... 40

3.3.5 CW photoinduced absorption spectroscopy ........................................................... 41

3.4 Electrical characterization .................................................................................... 42

3.4.1 Photovoltage measurements ................................................................................... 43

3.4.2 Photocurrent measurements .................................................................................... 48

3.4.3 Surface potential measurements ............................................................................. 50

3.4.4 Measurements on P3HT:PCBM ............................................................................. 52

3.5 Thermal characterization ...................................................................................... 53

3.5.1 Local temperature measurements ........................................................................... 53

3.5.2 Numerical simulations ............................................................................................ 56

Chapter 4 – Coupling hybrid interfaces with cells .................................................................... 59

4.1 Human Embryonic Kidney (HEK) 293 cells ......................................................... 59

4.1.1 Cultures of HEK-293 cells on polymeric substrates ............................................... 61

4.1.2 Basic electrophysiology of HEK-293 cells ............................................................. 62

4.2 Measurements on different substrates ................................................................... 65

4.3 Analysis of thermal effects ................................................................................... 67

4.3.1 Transient depolarization ......................................................................................... 68

4.3.2 Gradual hyperpolarization ...................................................................................... 70

4.3.3 Time evolution of membrane properties ................................................................. 73

4.3.4 Numerical modeling ............................................................................................... 77

4.4 Considerations on capacitive charging .................................................................. 80

Chapter 5 – Discussion and perspectives .................................................................................... 87

5.1 Discussion ............................................................................................................ 87

5.1.1 Capacitive stimulation ............................................................................................ 88

5.1.2 Thermal stimulation ................................................................................................ 90

5.1.3 Comparison with previous works ........................................................................... 91

5.2 Perspectives ......................................................................................................... 93

Appendix A .................................................................................................................................... 97

A.1 Optical Measurements .............................................................................................. 97

Page 5: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

iv

A.1.1 Femtosecond spectroscopy ........................................................................................... 97

A.1.2 Nanosecond spectroscopy ............................................................................................. 98

A.1.3 CW Photoinduced Absorption ...................................................................................... 99

A.2 Electrical and thermal characterization ................................................................... 100

A.2.1 Photovoltage measurements ....................................................................................... 100

A.2.2 Photocurrent measurements ........................................................................................ 101

A.2.3 Surface potential and temperature measurements ....................................................... 101

A.3 Electrophysiology measurements ............................................................................ 103

A.3.1 Electrophysiology setup .............................................................................................. 103

A.3.2 Electrolytic solutions and cell growth medium .......................................................... 104

Appendix B .................................................................................................................................. 106

B.1 Astrocyte cultures and electrophysiological properties ............................................ 107

B.2 Photostimulation of astrocytes membrane conductances ......................................... 109

B.3 Experimental methods ............................................................................................. 112

Bibliography ................................................................................................................................ 115

Acknowledgements ..................................................................................................................... 130

List of publications ..................................................................................................................... 133

Page 6: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

1

Chapter 1 - Bioelectricity

Bioelectricity, i.e. the arising of electrical potentials and currents in living systems, is a fundamental

process at the basis of many biological functions.1 Central in the formation of such electrical

phenomena are the biomembranes2 that enclose the different compartments of cells.

Electrochemical gradients arise across these selectively permeable membranes due to asymmetric

ion distributions,3 leading to the instauration of potential differences usually in the range from few

to hundreds of millivolts. The prototypical example of a bioelectrical phenomenon is the action

potential,4 i.e. the rapid variation of the plasma membrane potential that is able to rapidly propagate

along neurons transporting information in the nervous system. The same mechanism is also

responsible for muscular contraction in myocytes5,6

and release of hormones in endocrine cells.7

However, transmembrane potentials are present in non-excitable cells too, and are for example

involved in regulating the diffusion of ions and metabolites inside cells and organelles,8 in driving

the production of ATP in mitochondria9 and in controlling the fertilization process of oocytes.

10

Moreover, electric fields and associated currents are present also at the tissue scale and it has been

demonstrated that they play a pivotal role for example in the process of wound healing and in

establishing the left-right organ asymmetry during embryonic development.11

Understanding the

origin of bioelectrical phenomena and the ability of monitoring and controlling them represent thus

a fundamental aspect of biological sciences, with immense implications in medicine.12,13

In this chapter, the main techniques currently available for modulating and recording electrical

signals in biological systems will be presented; in particular, after a first introduction on standard,

purely electrical methods (Section 1.1), different strategies based on optical stimulation and

recording will be discussed (Section 1.2). In Section 1.3 the class of organic semiconductors and

their application in bioelectronics will be presented. Finally, in Section 1.4 photoactive bio-polymer

interfaces, which are the main topic of this thesis, will be introduced.

1.1 Electrical stimulation and recording

The study and exploitation of bioelectricity actually dates back to ancient times and has gone in

parallel with the discovery of electricity by mankind. Historical records show that in ancient Egypt

and Greece discharges from electrical catfishes and eels were used as treatments for pain relief and

Page 7: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

2

to improve blood circulation.11

However, the foundation of bioelectricity as a science is usually

dated to the pioneering experiments of the Italian physicist Luigi Galvani in the late XVIII century.

Around 1780 he began to conduct a series of experiments to prove that electric discharges from

different sources applied to preparations of frog legs were able to induce muscle contractions.14

His

observations were collected in the essay “De Viribus Electricitatis in Motu Musculari

Commentarius” (Commentary on the Effect of Electricity on Muscular Motion) published in 1791.15

To describe these phenomena he coined the term “animal electricity”, indicating a form of energy,

similar but different from natural electricity, that was generated by the tissues themselves. Galvani’s

work was then carried on and clarified by other scientists of that period, including Alessandro

Volta, Galvani’s cousin Giovanni Aldini and the German naturalist Alexander von Humboldt.14

Fifty years later, the German scientists Emil Heinrich Du Bois-Reymond and Hermann von

Helmholtz were able for the first time to record with a galvanometer action potentials (which they

actually called “action currents”) in frog nerves and to measure their propagation velocity. Du Bois-

Reymond’s book “Untersuchungen über thierische Elektricität” (Researches on Animal Electricity)

of 184816

is actually considered the beginning of scientific electrophysiology. In the following

decades the nature and properties of the nervous signals were investigated by many scientists,

culminating in the work of Alan Hodgkin and Andrew Huxley,17,18

who in 1952 published their

theory on the propagation of action potentials, one of the earliest and most famous models in

computational biochemistry.

The next big advancement in electrophysiology came in the late ‘70s with the invention of the

patch-clamp technique by Erwin Neher and Bert Sakmann.19–21

Recordings of electrical activity in

cells and tissues can be performed extracellularly, by placing the electrodes in proximity of one or

more cells, or intracellularly by accessing the cytoplasm to record the internal potential.

Intracellular recordings have the great advantage of giving information on the actual variations of

the membrane potential and the currents flowing through the cell membrane, thus allowing a deeper

understanding of the biophysical properties of the cell behavior. Before the introduction of the

patch-clamp technique, however, intracellular recordings were performed by impaling the cell with

metal electrodes or thin glass pipettes filled with an electrolyte solution;11

in order not to stress

excessively the cell, such a pipette needed to be very small (with sub-micrometer tip dimension)

and thus possessed a very high electrical impedance, making recording of small currents very noisy.

In a patch-clamp experiment, instead, a bigger pipette is used (usually with 1-2 μm tips) and it is not

inserted into the cell, but just put in contact with the cell membrane; after the formation of the so-

called gigaseal, the interior of the cell becomes electrically accessible, with much smaller

impedance with respect to intracellular electrodes.21

With this technique Neher and Sakmann were

able to isolate and record the conductances of single ion channels in the cell membrane for the first

Page 8: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

3

time, opening the way to the study of their fundamental influence on cell physiology and

pathophysiology.22,23

Patch-clamp rapidly became the gold standard in electrophysiology and allowed to decipher the

mechanisms by which neurons compute information and communicate between each other.

However, it has some intrinsic limitations, especially the possibility of measuring just one (or at

best very few) cells at a time and the short time the patched cell remains viable (usually less than

one hour). Thus, while it is an invaluable tool to study the functioning of single cells, this technique

is not suited for the investigation of the complex interactions between cells in large ensembles like

neural circuits, whose understanding is regarded as one of the major challenge of modern science.24–

26 In contrast, extracellular recording and stimulation techniques allow accessing larger populations

of cells and for longer times (even months), since they usually do not damage the cell membrane,

enabling to investigate the long-term properties of plasticity and learning in neural circuits.

Extracellular recordings can be performed with single electrodes (insulated metal electrodes or glass

pipettes) that record the activity of the cells in their proximity or with multi-electrode systems that

allow adding spatial resolution to the neural activity investigation.27,28

For in vitro experiments,

Multi Electrode Arrays (MEAs) exceeding 10000 electrodes are currently available;27–29

in vivo,

polytrodes30

with more than 100 electrodes have been developed. The main limitation in these

systems is the increasing impedance for smaller electrodes, which degrades the signal-to-noise

ration in recording and can lead to excessive heating in stimulation. Another strategy used for

extracellular recording is based on the field effect transistor (FET) architecture;31

in this case, the

cell extracellular potential basically acts as the gate signal of the transistor, modulating the current

flowing from source to drain. Both for electrodes and for FETs, the adhesion of cells to the active

surface and the electrical properties of the thin cleft between the basal membrane of the cell and the

device interface are fundamental in determining an efficient electrical coupling.32,33

In any case, all extracellular recording methods do not have access to the actual membrane potential

of the cell, but measure the so-called “(local) field potential”, in which the different electrical

signals related to neural (and glial) activity (action potentials, synaptic currents, calcium waves, …)

are superimposed in a complex spatiotemporal-dependent manner.34

Thus, retrieving relevant

information from extracellular recordings usually requires a great computational power and a

detailed previous knowledge of the system under study. To overcome the different limitations of

these techniques, considerable effort is now being undertaken in the development of devices, both

for multi-electrode and for transistor architectures, with capabilities of intracellular recording. These

systems usually consist in arrays of nanostructured surfaces which are engulfed or even partly

internalized by the basal cell membrane, allowing a very good electrical coupling between the

device and the intracellular compartment of the cell.26,27,35,36

This approach offers the advantages of

Page 9: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

4

both intracellular and extracellular techniques, being able to spatially address multiple

stimulating/recording sites and, at the same time, offering sensitivities comparable to intracellular

recordings.

1.2 Optical techniques

The obvious advantage of using electrodes to measure and/or excite biological tissue is that they

deal with signal of the same nature of bioelectrical phenomena, namely electrical current and

potentials. However, they need physical contact with the cell or tissue and their geometry is fixed

by design, so it cannot be adapted in real-time to the actual morphology of the system under

investigation. Moreover, it is usually difficult to obtain inhibition of neural activity instead of

excitation with purely electrical stimulation methods. Tools complementary to electrical means

have thus been developed in order to address these and other limitations; optical techniques have in

particular attracted a lot of interest.37

Since light can, to a certain extent and depending on the

wavelength, propagate through different biological tissues, no physical contact is required, thus

decreasing the risk of mechanical stress and damage. Moreover, light can be readily shaped in

desired pattern that can be adjusted to address specific regions of the field under study, with spatial

resolutions on the subcellular scale and a flexibility that pre-fabricated electrodes cannot achieve.

However, to modulate and monitor bioelectrical signal with light, some kind of transduction

mechanism is required. One of the main drawbacks is that quantitative analysis is not

straightforward and careful calibrations need to be performed. Also, at the moment there is no

single optical technique that is able to provide at the same time both stimulation and recording

capabilities. For stimulation, many architectures have been developed, which can be classified

based on the nature of the transducer: intrinsic absorbers, molecular probes, nano/micro-particles,

solid-state devices. As for recordings, the great majority of the research has focused on fluorescent

probes, even if few different approaches have been proposed.38,39

1.2.1 Optical measurement of bioelectric activity

Fluorescence is a ubiquitous tool in life science research and a wide variety of fluorescent probes

are nowadays available to detect virtually every molecule or investigate many biological processes.

To optically detect and record membrane potentials, the main strategy is to use the so-called voltage

sensitive dyes (VSDs).40,41

Standard VSDs are fluorescent molecules that get incorporated into the

cell membrane and have a fluorescence yield modulated by the external electric field; variation in

the membrane potential are thus reflected in a modulation of their fluorescence intensities. The main

problem of standard VSDs is that they generally stain unspecifically every membrane in the cell,

Page 10: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

5

reducing the effect of variations in the plasma membrane potential on the total modulation of

fluorescence. A very promising approach is thus the development of genetically encoded voltage

indicators (GEVIs)42,43

gene that can be genetically targeted to subpopulations of cells and to

subcompartments of the cell itself (like the cell membrane); these systems are usually composed of

a fluorescent protein and a voltage sensitive element able to modulate its emission. Both VSDs and

GEVIs allow the simultaneous recording of electrical activity in large populations of cells, both in-

vitro and in-vivo, with the possibility of having subcellular resolution, in a manner that is not

achievable with electrical recordings. However, they still face some issues with respect to the

sensitivity they can reach, especially when single-shot measurements need to be taken without

averaging (for example, when investigating spontaneous activity).37

Cell electrical activity can be investigated with fluorescence-based methods also by detecting

indirectly its effects. As one of the major signaling molecules, calcium ions have long been used to

assess cell activity.41,44

Calcium-sensitive dyes exploit the high transmembrane gradient in calcium

ions concentration and have been demonstrated to be sensitive enough to detect the variations due to

the opening of a single calcium channel in a synaptic spine. Other strategies involve the detection of

neurotransmitters released by synapses during neural activity or the activation of transmitter-gated

channels.

Optical monitoring of membrane potential has also been achieved without the need of fluorescent

probes, by exploiting intrinsic variation in the optical properties of cells, like changes in light

scattering, in birefringence or in optical dichroism, mainly due to local variations in refractive index

near the membrane upon osmotic changes associated with ion fluxes.45,46

However, the small signals

involved in these processes usually require extensive averaging and obtaining high-resolution

imaging with single-cell precision is still a challenge.

1.2.2 Direct optical stimulation

Except for some notable exception, like the retina and photosynthetic units, light usually does not

specifically interact with biological systems. To optically modulate cellular activity are thus

generally employed some photosensitive transducers brought into the cell or in its close proximity.

However, some examples of direct stimulation, i.e. without the use of exogenous absorbers, of

biological systems with light are present in literature. Actually, already in 1891 the French scientist

Jacques-Arséne d’Arsonval reported the capability of exciting muscular fibers with light.47

Later,

Arvanitaki and Chalazonitis published, starting from the ‘40s, different works in which both

excitation and inhibition of neural activity could be obtained optically in different neural

preparations.48,49

While many of these preparations were actually stained with vital dyes to obtain

Page 11: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

6

photosensitivity, other systems, like some large neurons of the marine mollusk Aplysia Californica,

showed the presence of endogenous fluorophores. This intrinsic pigmentation was related to heme

or carotenoid molecules. Their work was expanded by Fork in his paper of 1971,50

where he showed

that blue or green laser light can be used to stimulate Aplysia neurons and map cellular

interconnections. In all these reports, however, the mechanism of transduction was never clearly

identified. In 2008, Reece et al. reported that inhibition of C1 neurons of Helix Aspersa following

irradiation by 532 nm laser light is mediated by the activation of chloride currents;51

also in this

case, however, the primary photoabsorber could not be identified.

Another strategy for direct optical stimulation of neurons is based on the use of ultrashort laser

pulses (in the femtosecond to nanosecond regime) of near-infrared light. Transduction of the optical

signal is based in this case on non-linear processes due to the high peak intensity of the pulses.52–54

Two regimes of stimulation have actually been identified, depending on the laser intensity. At low

intensities, production of reactive oxygen species is reported to mainly mediate the firing of action

potentials, probably due to two-photon absorption by some endogenous fluorophore; at higher

intensities, spiking activity is a consequence of membrane depolarization due to transient poration

of the plasma membrane.

A more interesting approach to optically modulate cellular activity without the use of external

sensitizers is Infrared Neural Stimulation (INS), which is based on water absorption of pulses of IR

light. It has been proposed by Wells et al. in 2005,55

when they reported the successful in vivo

stimulation of compound nerve and muscle potentials in frogs and rats. Following this first

demonstration, the biophysical mechanism of transduction was investigated by different groups.56–60

In 2012 Shapiro et al.58

demonstrated that the local rise in temperature following light absorption by

water results in a variation of the plasma membrane capacitance in different kinds of cells, which is

reflected in a transient depolarization of the cell with values compatible with the firing of action

potentials in neurons. However, Albert et al.57

showed that infrared laser-evoked stimulation of

sensory neurons is mediated by the opening of temperature sensitive ion channels (in particular the

Transient Receptor Potential Vanilloid channel 4,61–64

TRPV4, which is expressed in different

neuron families). Also, while short pulses (on the time scale of milliseconds) of IR light have been

demonstrated to promote neural activity, for longer illumination suppression of action potential

formation and the blocking of spikes transmission along nerve fibres have been observed.65

This

opposite behaviour is attributed to the effect of the increasing baseline temperature on the Hodgkin-

Huxley gating mechanism of action potentials.66

Page 12: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

7

1.2.3 Molecular-based stimulation

Although, as reported above, cells can exhibit some intrinsic sensitivity to light, the efficiency of

such processes is quite low, especially for visible illumination (where intensities up to several

W/mm2 can be necessary), and usually not fully controllable. Researchers have thus generally relied

on exogenous (i.e. not present in the system in natural conditions) transducers with more efficient

absorption of optical radiation. A first approach is based on photoactive molecules or

macromolecules, i.e. molecular systems that upon illumination are able to modulate some relevant

function in the cell.39

Photoisomerizable compounds (also known as photoswitches) exploit the variation in functionality

of a molecule between two different isomeric forms; upon light irradiation in a specific wavelength

range the original inert system, usually based on an azobenzene unit, undergoes a cis-trans

isomerization that results in a biologically active compound. The active form can be stable for hours

or days, and inactivation, i.e. the reverse isomerization, can be usually obtained by irradiation with a

different wavelength. This principle has been applied for light-mediated pharmacological control of

different targets, like enzymes, ion channels and G-protein-coupled receptors.67,68

In particular,

targeting of ion channels can be exploited to control cell membrane potential and thus promote or

inhibit activity in neurons. As an example, acrylamide-azobenzene-quaternary ammonium (AAQ) is

a photoswitch inert in its cis form, but acts as a potassium channels blocker in its trans form,

increasing neuron excitability.69,70

This molecule has been successfully employed to restore light

sensitivity in blind rat retinas both in vitro and in vivo with intensities of few tens of μw/mm2.

A similar strategy is that of photocleavable compounds, in which the functional molecule is

inactivated by blocking it inside a molecular cage; this cage presents photolabile bonds that are

broken upon illumination, releasing the trapped molecule, for example a neurotransmitter, a

secondary messenger or an enzyme.71–73

With these systems it is possible to control with unique

spatiotemporal resolution the release of a molecule in specific regions inside or outside the cell and

they represent an invaluable tool to investigate cellular physiology and patophysiology. However,

they still suffer from some limitations, especially the inability to reverse the activation process and

thus decrease the local concentration of the active molecule, which can also diffuse to other regions.

Moreover, uncaging usually requires illumination with UV light, which has a limited penetration

depth in tissues.

In 2005 Boyden et al.74

reported the possibility to express in mammalian neurons channelrhodopsin-

2, a light-sensitive ion channel, to control neural activity with millisecond precision. This work

opened the way to the optogenetic revolution in neuroscience;75

over the next decade a wide variety

of optically controlled systems to modulate cellular activity have been developed: ion channels for

Page 13: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

8

both excitation74

and inhibition76,77

of neural activity, G-protein-coupled receptors to control

biochemical signaling pathways,78

transcriptional effectors to influence gene expression.79

The main

advantage of optogenetics, apart for the high spatiotemporal resolution achievable with optical

stimulation, is the possibility to genetically target the expression of the light-sensitive proteins to

specific subpopulation of cells, allowing the investigation of specific functions of different types of

cells in complex biological tissues.

1.2.4 Nano-/micro-particle stimulation

The use of nanoparticles (NPs) in biology has developed so much in the last few decades that the

term “nanomedicine”80,81

has been introduced to describe this broad field of research. The interest in

these systems stems from the combination of several properties: (i) they can be used as scaffolds to

transport different functional molecules; (ii) they possess a great surface-to-volume ratio; (iii) they

are easily controlled in shape and dimension; (iv) they present unique optical properties. NPs have

been employed both for diagnostic and therapeutic means as contrast agents for functional imaging,

carriers for drug delivery and actuators for photodynamic and photothermal therapies.

Their use as transducers for direct modulation of cell membrane potential is a relatively new and

less explored field. In a first attempt, Winter et al.82

proposed in 2001 to bind semiconducting

quantum dots (QDs) to the plasma membrane of different cell types and to exploit their electrical

dipole upon illumination to optically stimulate the cell, but they could not obtain any reliable

photoactivation. A more successful approach has been that of using NPs in form of thin films, both

as a functional substrate for cell growth83

or as a coating for patch micropipettes.84,85

In these

reports, apart from the capability of inducing a local electric field upon illumination, also the

possibility to have net Faradaic currents due to charge transfer reactions has been investigated.

Recently, semiconducting nanoparticles have also been used as sensitizers for thermal stimulation

based on the principle of Infrared Neural Stimulation.86,87

The advantage of using exogenous

sensitizers with respect to rely on water absorption is in the possibility of using light in the near-IR

(around 800 nm), while water absorbs efficiently light at higher wavelengths. As for INS, also in

this case both excitation and inhibition of neural activity could be observed, based on the timescales

of the stimulation protocol. A similar approach has been followed by other researchers using bigger

particles, on the scale of few microns, of iron oxide59,88

or carbon, again as photoabsorbers both in

the visible and the near-IR.

Page 14: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

9

1.2.5 Device-based stimulation

Solid-state electronic devices have been largely investigated in the past as bidirectional platforms

for electrical interfaces with neural tissues, starting with the pioneering work with silicon-based

transistors of Fromherz and coworkers in the ’90s.89–91

Coupling of electronic devices with optical

excitation has been proposed in 2001 by Colicos et al.92

in order to overcome the problem of poor

spatial resolution of electrode-based stimulation. Their approach relies on the variation in

conductivity of silicon under illumination;93

a passivated silicon chip is used as a substrate for the

cell culture; while an electric bias is applied to the device, a pulse of light is shone on the desired

area, producing a current in the semiconductor that is capacitively coupled to the cell layer on top,

resulting in a local excitation.94,95

This method allows thus to have an electrical capacitive

stimulation but with a flexible geometry defined by the patterned illumination that can be varied

during the experiment. Beyond crystalline silicon, also thin-film devices based on amorphous

hydrogenated silicon or titanium dioxide have been demonstrated.96,97

Recently, Palanker and his collaborator have realized a wireless retinal prosthesis based on the

photovoltaic effect in silicon. In this case, however, an array of discrete photodiodes driving

discrete electrodes for cell stimulation has been employed.98

Light is used here not to improve

spatial resolution, but to avoid the need of wiring to power the electrode array in the eye.

1.3 Organic semiconductors for biological applications

The devices described in the previous section are all based on inorganic semiconductors and metals;

indeed, these materials have been widely applied in the field of bioelectronics, i.e. the coupling of

solid-state electronic devices with biological systems. The continuous evolution of silicon

electronics has lead in the past decades to the development of many technologies now ubiquitous in

medical research and practice, from stimulating devices like pacemakers and cochlear implants to

recording instruments such electrocardiographs and electroencephalographs. However, problems

like the rigid nature of inorganic crystalline materials and their purely electronic conduction

properties have always posed difficulties in realizing efficient direct interfacing with biological

tissues.

Organic semiconductors99

are materials based on conjugated carbon atoms with sp2 hybridization;

these materials show semiconducting properties since the π-electrons easily delocalize along the

conjugated system. In the last thirty years this class of materials has gained a great deal of attention

in the scientific community due to their unique properties: they can be chemically modified to fine

tune their optoelectronic properties or to add functional groups; they can be usually processed with

Page 15: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

10

solution-based technologies that allow relatively simple and cheap fabrication processes and fast

prototyping; they generally posses, especially conjugated polymers, a “soft” nature that permits the

realization of flexible and conformable devices; they can, in some cases, conduct both electronic

and ionic charges. These properties have led to the development, and in some cases

commercialization, of different technologies, especially organic light emitting diodes (OLEDs),

organic photovoltaic cells (OPVCs) and organic thin film transistors (OTFTs).

Organic semiconductors, and in particular conjugated polymers, have been extensively used as

coating materials for inorganic electrodes in bioelectronic applications.100

Since their “soft” nature

closely matches the mechanical properties of biological tissues and their carbon-based chemical

structure resembles that of basic compounds in living matter, these materials generally present

better biocompatibility properties with respect to standard inorganic metals and semiconductors.101

Moreover, the possibility to have both ionic and electronic conduction represents a bridge between

the typical transport mechanisms in biological systems, which is based on ionic species, and the

electron-based conduction in standard electronic devices.102

All these properties have allowed the

achievement of better interfacing between electrodes and biological matter, decreasing both

electrical impedance of the contact and inflammatory responses from the tissue.

In the past decade, however, organic semiconductors have started to emerge in life sciences not only

as passive elements in coating layers, but as active functional materials for novel technologies. This

field of research, named organic bioelectronics by Berggren and Dahlfors in a seminal review paper

of 2007,103

has been greatly expanding in the last years; it now encompasses a number of different

applications that exploit the peculiar optoelectronic and mechanical properties of organic

semiconductors,101

from sensing of biomolecules104,105

to functional substrates for cellular

growth.106,107

Among these, technologies for interacting with bioelectrical signals in living systems

have attracted considerable attention.

The preferred architecture proposed for sensing bioelectrical signals, and in particular neural

activity, is the transistor, where the local potential variation to be detected acts as a gate signal for

the device. The most common design for an organic transistor is the one based on the field effect

(OFET, organic field effect transistor), in which the electric field generated by the gate potential

modulates the conductivity of the semiconductor. Muccini and his group have realized in 2013 an

organic thin film transistor capable not only of recording, but also of stimulating and inhibiting

neural activity;108

interestingly, they showed in recording configuration a signal-to-noise ratio 6 to

16 times better than the one of standard commercially available MEAs devices. This increase in

performance was tentatively attributed to a more efficient capacitive coupling between the organic

interface and the biological tissue. This device was based on a perylene-based small molecule as the

Page 16: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

11

organic semiconductor, but similar OFET architectures were also proposed by Biscarini and

coworkers with pentacene molecules.109

Another example where the peculiar properties of organic semiconductors are fully exploited is the

organic electrochemical transistor (OECT).110

In these devices the electrolyte, acting as a gate

electrode, is in direct contact with the organic semiconductor and ions can migrate into the material

modulating its bulk conductivity. This 3D modulation of the semiconductor conduction, with

respect to a standard field effect transistor where charges are transported along a 2D channel at the

interface with the gate dielectric, gives these devices a very high sensitivity, at expense of the

switching speed. Based on the OECT architecture, Malliaras and coworkers have recently

developed a novel conformable neural interface array that can record local field potentials and

action potentials in vivo from the surface of the brain without the need of penetrating

electrodes.111,112

The same principle of interplay between ionic and electronic transport is used, in a specular manner,

in organic electronic ion pumps (OEIPs), mainly developed in the laboratories of Berggren and his

collaborators.113

These devices operate basically as transistor for ionic species, where their transport

is regulated by an electronic gate signal. They have been employed as platform for the precise

spatio-temporal release to biological preparations of simple ions (like calcium)114

but also signaling

molecules and drugs;115,116

in particular, by delivery of different neurotransmitters, OEIPs can be

used to modulate electrical activity in neural tissues.

Another interesting application is the exploitation of the superficial oxidation state of an organic

semiconductor, in particular conducting polymers, to control the ability of cells to grow and

proliferate on top of it. Indeed, different groups have demonstrated, for different types of cells, that

they preferentially grow on oxidized region of the material surface, probably because of a different

interaction with the substrate of the proteins mediating the adhesion process.107,117

Closely related to

these studies is the possibility to control the outgrowth of neurites of cells grown on conducting

polymers with the application of short electrical pulses to the substrate,118

demonstrating the

importance of bioelectrical signals not only at the level of action potential transmission but also in

modulating the cell fate.

1.4 Photoactive bio-polymer interfaces

Although much of the interest in organic semiconductor is related to their unique optical properties,

the possibility of using them as transducer of light stimuli into bioelectrical signals has only recently

started to emerge.39

The basic idea is to exploit the photovoltaic action in organic materials,119

and

Page 17: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

12

in particular in bulk heterojunctions120,121

of conjugated polymers with fullerene-based electron

acceptors, to generate electrical charges upon illumination; these charges should then be able to

modulate the membrane potential of a cell grown on top of the device.

1.4.1 The hybrid solid-liquid organic photovoltaic cell

The first step in this direction was to demonstrate that an organic photovoltaic cell could work in a

liquid environment like that of a cell culture. Indeed, organic solar cells are quite known to suffer

from degradation problems due to oxygen and humidity;122

these issues stem not only from the

intrinsic electrochemical stability properties of conjugated polymer,123

but also from the degradation

of the highly reactive low work function metals usually employed as cathodes in these kinds of

solar cells.124

This problem was solved by realizing that the electrolytic solution of the extracellular

medium is by itself a charge conductor and can be used directly as an electrode, without the need

for a metal one. The device architecture reported by Antognazza et al.125

was thus composed of a

glass substrate covered with a transparent conductive oxide (ITO, indium-tin oxide), on top of

which a thin film of the photoactive layer was deposited via spin-coating. The organic materials

used in this first report where the conjugated polymer poly[2-methoxy-5-(2’-ethylhexyloxy)-p-

phenylene vinylene] (MEH-PPV) as light absorbing material and hole conductor, and the fullerene

derivative C61-butyric acid methyl ester (PCBM) as electron acceptor. The device was then put in

an electrolytic solution resembling the extracellular medium (Krebs-Ringer’s solution, KRH) and

the circuit was closed with a gold wire put as a counter-electrode in the solution. In the work, the

authors were able to demonstrate that such a hybrid solid-liquid photovoltaic device was able to

support the generation of a photocurrent, with an action spectrum similar to the one of a standard

organic solar cell.

In a subsequent work Lanzarini et al.126

showed that also other conjugated polymer could work as

light harvesting material in these hybrid devices and proposed that, under continuous illumination,

hydrogen evolution could occur at the polymer/electrolyte interface. Among the semiconductors

used in this report, poly(3-hexylthiophene-2,5-diyl) (regioregular P3HT) was selected as the choice

material for the applications in biological interfaces. Guerrero et al.127

were also able to fabricate

organic photoelectrochemical cells (OPECs) based on P3HT:PCBM blends with quantitative

photocarrier conversion in which, by proper selection of the redox couple, the organic layer was

able to provide either holes or electrons, expanding the device applicability to the production of

different solar fuels.

Page 18: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

13

1.4.2 Poly(3-hexylthiophene)

P3HT is one of the prototypical conjugated polymers used in the field of organic electronics and it

has been by far the most investigated polymer in the understanding of the photophysics behind the

functioning of organic photovoltaic cells.128

It is composed of a backbone of thiophene rings that

determine its optoelectronic properties, functionalized with alkyl side chains to confer solubility in

organic solvents. In its regioregular version (rr-P3HT) the side chains are periodically ordered along

the backbone, allowing interdigitation with adjacent polymeric chains. In the solid state, P3HT

tends to form lamellar structures with interchain stacking of the thiophene rings, leading to an

increased delocalization of the electron density and increase in charge mobility. Depending on the

polymer molecular weight and the deposition process, films usually present a crystalline fraction of

lamellae embedded in an amorphous matrix.129,130

The material presents a broad absorption in the visible peaking in the green (500-550 nm), with a

side shoulder at 605 nm attributed to the formation of the ordered lamellar phase. As in other

organic semiconductor, light absorption leads to the formation of a bound electron-hole couple

(exciton), with a binding energy of about 0.7 eV;131

some driving force, usually given by the

presence of an electron acceptor like PCBM, is thus needed to break the exciton and obtain free

polarons. While P3HT is generally considered a p-type material (i.e. a hole conductor), careful

investigation of its transport properties have revealed similar mobilities for holes and electrons in its

intrinsic state;132

however doping due to exposure to air results in traps for electrons that drastically

reduce electron mobility.133,134

Different studies have investigated P3HT stability upon illumination. As in many conjugated

systems, exposure to UV light in presence of oxygen leads to irreversible degradation of the

material, due to oxidation by radical chain mechanisms that destroy the backbone conjugation with

a deterioration of optoelectronic properties.123

Upon illumination with visible light, instead, only a

reversible effect is observed, attributed to the formation of a charge-transfer complex with

molecular oxygen.135,136

This process occurs, at a much lower rate, also in dark and it should be

favored by the presence of humidity.137

Interestingly for biological application, upon illumination

degradation of films in direct contact with a saline solution is not enhanced with respect to films

kept in air.138

Due to the presence of the alkyl side chains, P3HT surfaces are usually quite hydrophobic. To allow

the attachment of cell cultures, films usually need to be pre-treated with some adhesion layer.

Scarpa et al.139

demonstrated the possibility of growing mouse fibroblasts on P3HT films by using

interlayers of proteins like fibronectin, polylysine or collagen. However, also a mild plasma

treatment, to oxidize the surface and increase hydrophilicity, was seen to be sufficient to promote

Page 19: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

14

cell adhesion and proliferation. Biocompatibility of P3HT and P3HT:PCBM films has been studied

up to four weeks in-vitro also for cultures of primary cells like neurons140,141

and astrocytes,142

demonstrating viability rates and electrophysiological properties similar to cells cultured on

standard control substrates.

1.4.3 Photostimulation of primary cells

In 2011 Ghezzi, Antognazza et al.140

published the first report about the photostimulation of

primary neurons via light absorption in a P3HT-based biointerface. The device consisted of a thin

film (≈ 150 nm) of a P3HT:PCBM blend (in relative weight ratio 1:1) deposited on an ITO-coated

glass substrate. Rat hippocampal neurons were cultured on top of the active material, pretreated

with a poly-L-lysine adhesion layer to promote cellular adhesion. Pulsed (20 to 50 ms) light

excitation was obtained with a 532 nm green laser, delivered to the preparation through the

microscope objective (in an upright configuration) with an intensity of about 10 mW/mm2.

Recording performed with standard patch-clamp techniques showed that, upon illumination on the

cell body of the neuron, action potential firing was clearly elicited, with a success rate higher than

85 %. Interestingly, moving the illumination spot outside of the cell body did not produce any

excitation of cell activity, demonstrating an intrinsic spatial selectivity of the stimulation

architecture. The actual coupling mechanism between the device and the neuron was not fully

elucidated, but the author proposed a capacitive charging of the polymer/electrolyte interface upon

charge generation in the active material. In a later paper, it was reported that also in devices with an

active layer composed of only P3HT neuronal excitation could be reliably obtained with similar

power intensities; this result indicated that, contrary to what happens in organic solar cells, charge

generation in the bulk of the semiconducting material is not the main physical phenomenon leading

to cellular stimulation in these hybrid interfaces.

In a subsequent work by Benfenati, Martino et al.142

the same device architecture was used to

investigate the effect of photostimulation on astrocytes (Appendix B). In this case, continuous

illumination was used (λ = 560 nm, intensities up to 13 mW/mm2). A progressive depolarization of

the cell was observed along with a modification of the rectification properties of the membrane.

Pharmacological experiments demonstrated that the photostimulation was causing the opening of

the chlorine channels ClC-2. Based on the known stimuli involved in modulation of ClC-2

conductances, it was proposed that a local acidification was occurring upon illumination; this could

be due either to a capacitive rearrangement of charges at the polymer/electrolyte interface or to the

occurrence of electrochemical reactions.

Page 20: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

15

1.4.4 Ex-vivo experiments on blind retinas

The capability of the polymeric interface described above to transduce optical signals into a

bioelectrical stimulation of neuronal cells makes these devices an interesting platform for

neuroscientific research, but has also important implications in the field of vision restoration to

blind people. Diseases that affect the photoreceptor layer in the retina impairing its light sensitivity,

like Retinitis Pigmentosa, age-related macular degenerations and Stargardt’s disease, are the main

cause of legal blindness in the western world. Restoring light sensitivity in a degenerated retina is

thus one of the main focuses of research in developing visual prostheses.

The first report of a successful interfacing of a blind retina with a photoactive polymeric-based

device came in 2013 by Ghezzi et al.141

Photoreceptor degeneration was induced by prolonged

exposure to intense light in albino rats; the retinas were then explanted and put in contact with a

P3HT-based device in a subretinal configuration, i.e. with the degenerated layer of photoreceptors

facing the polymer. Multi-unit activity and local field potentials were recorded with an extracellular

electrode from the ganglion cell layer upon illumination with pulses of light (λ = 532 nm) over a

wide range of intensities (from 10 nW/mm2 to 4 mW/mm

2). While for blind retinas on control glass

substrates only small responses to light stimulation could be measured and with a threshold at quite

high intensities (80 μW/mm2), in retinas placed on photoactive devices the response was

significantly increased and the activation threshold was reduced (0.3 μW/mm2).

Similar results were also obtained in 2014 by other groups. Gautam et al.143

used a blend of P3HT

and a n-type conjugated polymer, P(NDI2OD-T2), to successfully excite blind embryonic chick

retinas. The same experimental model of retinal degeneration was used by Bareket et al.144

to

demonstrate a device architecture based, instead of semiconducting polymers, on a mesh of carbon

nanotubes, used as electron acceptors and transporters, sensitized for visible-light absorption with

inorganic core-shell quantum dots. In both cases, excitation intensities of the same order of

magnitude as the ones used by Ghezzi et al. were reported (tens to hundreds of μW/mm2); again, a

capacitive charging of the device interface was reported as the excitation mechanism.

Page 21: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

16

Chapter 2 – The plasma membrane

Eukaryotic cells are complex systems organized in many specialized compartments and organelles

that have evolved to perform different tasks essential for the correct functioning of the cell.145

Many

of these organelles, like the nucleus, the Golgi apparatus and the endoplasmic reticulum, are defined

by a membrane that separates them from the rest of the cell. Moreover, each cell is separated from

the extracellular space by what is called the plasma membrane.146

The basic elements that constitute

all these systems are double layers of amphiphilic lipids that spontaneously tend to form

bidimensional structures to balance hydrophilic and hydrophobic forces. Biomembranes, however,

are not simply partitioning elements, but are usually an active part of the organelles, fundamental in

their functioning. Their functionality is mainly given by the presence of several proteins, embedded

in or bounded to the surface, involved in a variety of biological processes.

One of the fundamental characteristics of biological membranes is their selective permeability, i.e.

the ability to block the diffusion of some molecular species while letting other pass through. The

diffusion of charged species in particular, from simple ions to charged macromolecules, is tightly

regulated by the presence of very selective transporting proteins that span the membrane, called ion

channels.147

This precise control of ion concentrations leads in many cases to the establishing of a

transmembrane potential difference that is fundamental in controlling several of the membrane

functions.3 In particular, the plasma membrane of virtually all eukaryotic cells is characterized by a

potential difference between the inner and outer side; for animal cells in physiological conditions

this membrane potential is usually in the range between - 80 mV and - 40 mV and it has been

shown, for example, to be correlated with the phase of the cell cycle and the cell ability to

proliferate.148,149

The properties of plasma membrane potentials have been mainly studied in the

context of signals transmission in excitable cells and in particular in neural circuit.

Since one of the main goals of the hybrid polymer-based devices introduced in Chapter 1 is the

ability to modulate the electrical potential in cell membrane, a deep understanding on their

composition and functioning is of great importance. In this Chapter, we first introduce (Section 2.1)

the fundamental structure of a typical biomembrane, composed of a lipid bilayer with embedded

proteins. We then discuss a particular class of transmembrane protein, the ion channels, which

regulate the flowing of ion species through the membrane (Section 2.2) and how they are involved

in the arising of the membrane electric potential (Section 2.3). The discussion here is focused on the

Page 22: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

17

plasma membrane of the cell, even if several of the elements presented are common also to other

systems.

2.1 The structure of the plasma membrane

A biological membrane is an object that is able to separate two different aqueous compartments by

selectively control the flow of molecules between them. In particular, the plasma membrane defines

the cell boundary, dividing the extracellular space from the interior of the cell. It is thus the main

element that controls the uptake and release of substances, regulating the composition of the

cytoplasm. Membranes are mainly composed of three different kinds of molecules: lipids, proteins

and sugars.146

Lipids are a broad class of hydrophobic or amphiphilic small molecules with different biological

functions, from energy storage to signaling and structural properties. The basic constituents of the

membrane are a particular class of lipids, called phospholipids. They are amphiphilic molecules

composed by a hydrophilic head containing a phosphate group to which a hydrophobic tail is

attached, usually made of two fatty acid chains. The head of phospholipids found in membranes can

be either negatively charged or zwitterionic (i.e. presenting both a negative and a positive charge);

positively charged phospholipids are not found in nature, but can be synthesized in laboratory. To

balance the hydrophobic and hydrophilic forces, phospholipids in aqueous solutions usually tend to

aggregate in structures that screen their hydrophobic tails from interactions with water molecules.

One of the most common structures is the lamellar phase, in which phospholipids arrange

themselves in a bilayer with hydrophobic tails in the middle and hydrophilic heads towards the

solution. This two-dimensional structure is the basic component of a biological membrane. The

presence of a hydrophobic core renders lipid bilayers generally quite impermeable to polar and

especially charged molecules, while apolar molecules like O2, N2, CO2 and fats can usually rapidly

permeate through them. A notable exception is water, which can permeate through membranes

although being polar; the actual mechanism by which this happens is however still debated.

Proteins are large macromolecules formed by one or more chains of amino acids and are the main

functional elements in living systems. In membranes, proteins perform a variety of tasks, among

which is the formation of selective pores for the transport of molecules for which the lipid bilayer is

normally impermeable. Membrane proteins can be classified based on how they are associated with

the lipid bilayer. Integral proteins are molecules that are permanently bound to the membrane, either

spanning the entire lipid bilayer (polytopic or transmembrane proteins), or being attached only to

one side, usually the inner one (monotopyc proteins). Peripheral membrane proteins are instead only

Page 23: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

18

temporarily attached, either to an integral protein or to the lipid bilayer, via a combination of non-

covalent interactions.

Sugars in biological membranes can be found in complexes with other molecules, either with lipids

(glycolipids) or with proteins (glycoproteins). Since they can form many different structures in

relatively short chains, sugars are mainly used as distinguishing features that enable recognition

processes between biomolecules. For this reason, in membranes they usually occur on the outer

side, allowing specific cell-cell interactions and targeting of signaling molecules like hormones.

Figure 2.1 | Schematic representation of the composition of the plasma membrane.

The basic model that describes a cell membrane is the fluid mosaic model, introduced in 1972 by

Singer and Nichols.150

It treats the membrane as a two-dimensional fluid composed of a matrix of

lipids in which proteins are embedded. In this model, lipids and proteins can easily diffuse and no

long-range order is present in the membrane, which is seen as a homogeneous system. This model

has been refined in the following years to take into account the observation that both lipids and

proteins may distribute inhomogeneously, forming clusters and domains in the membrane.

Biological membranes actually contain a great variety of lipids that differ for both the length of the

hydrophobic tails and the nature of the hydrophilic head. Also, transmembrane proteins can have

different lengths of the hydrophobic surface embedded in the membrane; if this length does not

match the hydrophobic core of the bilayer, unfavorable interactions can arise. Lipids with longer or

shorter fatty acid tails tend thus to accumulate around proteins with different core length to

compensate this effect. In the mattress model151

of Mouritsen and Bloom (1984) this phenomenon,

called hydrophobic matching, is proposed to explain the accumulation of certain lipids around

different proteins and the attraction between proteins due to capillary forces. Similarly, the

interaction mismatch between different species of lipids can explain the formation of aggregates and

domains.

Page 24: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

19

It is thus quite clear that biological membranes are complex systems in which thermodynamics

plays a fundamental role. Lipid bilayers can exist in a variety of phases, from crystalline to gel to

fluids and indeed phase transitions are observed in membranes at physiologically relevant

temperatures.146

The actual properties of a membrane considerably depend on its particular

composition. Different types of cells possess membranes with quite different composition of lipids

and proteins. Interestingly, it has also been shown that the same cells can be characterized by

different compositions of the membrane if grown at different temperatures or pressures. Moreover,

also the two leaflets composing the bilayer have different distribution of lipids, which rarely

spontaneously flip from one side to the other.146,152

Indeed, specific proteins like flippases and

scramblases are present in the membrane to allow the translocation of lipids between the two

monolayers. This asymmetry153–155

is fundamental in determining the electrostatic properties of the

membrane and actually the inner leaflet is usually more negatively charged with respect to the outer

one.

2.2 Ion channels

While lipid bilayers by themselves are impermeable to many molecules, especially polar and

charged ones, transmembrane proteins allow the transport of molecules from one side to the other of

biological membranes. These proteins can be classified in two broad categories depending on

whether they work or not towards thermodynamic equilibrium:156

passive pores (or channels) that

facilitate the diffusion of particular chemical species through the membrane, or active transporters,

which utilize energy (for example in form of ATP) to move substances against the electrochemical

gradients between the intracellular and extracellular compartments.

Among these membrane transport proteins, ion channels are a particular class of passive pores that

mediate the passage of ions (mainly K+, Na

+, Cl

-, Ca

2+) present in the electrolytic solution

composing the physiological media. Ion channels are fundamental in determining the bioelectrical

properties of membranes, promoting the establishment of a transmembrane potential and controlling

its value upon the presence of proper stimuli. Being able to modulate the intracellular ion

concentrations, they can also regulate cell volume by driving osmotic flow of water through the

membrane. A vast variety of ion channels with diverse properties have been discovered;157

their

expression is extremely variable between different cells, and also in the same cell at different phases

of its development, conferring them diverse bioelectrical properties. In particular, ion channels

present two peculiar characteristics that make them extremely flexible tools in controlling the flow

of ions through the membrane:3 (i) they can have a very high selectivity for a particular ionic

Page 25: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

20

species, blocking the passage of all the others; (ii) they can be activated or inactivated by the

presence of external stimuli.

2.2.1 Ion channel structure and selectivity

As for all proteins, the basic functioning of ion channels is strictly dependent on their geometry.

While many characteristics of these proteins had been already inferred via indirect methods, the first

actual high-resolution crystallographic structure of an ion channel became available only in 1998.158

The core part of an ion channel is a transmembrane unit with a central pore that spans through the

entire width of the membrane. This region is generally made up by the symmetrical arrangement of

different subunits of the protein around a central axis, forming a channel that can be filled with

water molecules. Along this channel, specific structures made by amino acid residues are present to

form what is called the selectivity filter,159,160

i.e. a region where a particular ionic species may be

recognized and allowed to pass. To allow very fast transport of ions, in many cases the pore is

formed by a long passageway where ions can diffuse freely, interrupted by a quite small selectivity

filter. Indeed, ion channels can transport up to 108 ions per second.

3 Apart from this central part, ion

channels can also have domains in the extracellular or intracellular space to sense the presence of

different stimuli.

The mechanism of channels selectivity is the result of different competing phenomena:160

(i) the

geometrical width of the filter, which can block larger particles from passing; (ii) the hydration state

of the ions that increases their geometrical hindrance; (iii) the electrostatic properties of amino acid

residues present in the selectivity filter. To pass through the selectivity filter, hydrated ions need to

lose their shell of water molecules and temporarily bind to the amino acid residues, from which they

are subsequently released by thermal energy. The channel selectivity between anions and cations is

thus given by the actual charge on the side chains of the amino acid forming the selectivity filter.

The selectivity for different ions with the same charge is instead given by the fact that smaller ions

(like Na+) have a stronger interaction with water molecules than bigger ones (like K

+). When an ion

reaches the selectivity filter, it binds to it only if it is in a more energetically favorable situation with

respect to the hydrated state. A small pore with a high binding energy is thus able to strip the water

molecules from a Na+ ion and let it through, while a K

+ ion is too big to pass from the filter. Instead,

a larger pore with a lower affinity can still provide sufficient energy to dehydrate the K+ ion, while

the Na+ ion retains its water shell and cannot pass.

Page 26: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

21

Figure 2.2 | Schematic representation of an ion channel. To pass through the pore, a

hydrated ion (A) needs to lose its water shell to bind to specific residues of the selectivity

filter (B). Conformational changes in the channel structure due to external stimuli may

close the pore, a mechanism called gating. © The Nobel Foundation.161

2.2.2 Gating mechanisms

Another peculiar property of ion channels is that they usually exist in at least two relatively stable

different conformations.3 In particular, channels have always at least an open and a closed state, and

can go from one to the other in the presence of appropriate stimuli. Moreover, closed channels can

be in a resting state, in which they can be opened upon the occurrence of the stimulus, or a

refractory state, in which they are insensible to external signals. This mechanism is called gating,

and it is fundamental in controlling many bioelectrical properties of cells. The stimulus triggering

the gating of a channel may be of different natures, depending on the actual functionality of the

channel itself.159,162

Voltage-gated channels are controlled by a variation in the membrane potential. These

channels posses a subunit (the “voltage sensor”) that can sense the local electric field and

trigger a conformational change in the protein. Although they are usually activated by a

depolarization of the membrane (i.e. a variation of the potential towards more positive values),

examples of voltage-gated channels that open upon hyperpolarization have also been found.

This mechanism is essential in the formation and propagation of action potentials in neurons,

Page 27: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

22

based on the interplay between the rapid opening and closing of sodium and potassium

channels.

Ligand-gated channels switch from a conformation to another upon binding of a specific

chemical species in a selective pocket on either the intracellular or extracellular side of the

membrane. There is a variety of substances that can act as gating agents, from simple ions like

Ca2+

to signaling molecules like neurotransmitters.

Mechanically-gated channels respond to a mechanical deformation of the membrane or of the

cytoskeleton of the cell; they are the main transducer of sensory stimuli like touch and hearing,

but are also involved in cardiovascular regulation and osmotic homeostasis.

Light-gated channels contain an isomerizable chromophore (like retinal) that changes

conformation upon light absorption, triggering the opening and closing of the channel. Only one

class of natural light-gated channel, namely channelrhodopsin from unicellular green algae, is

currently known; however a lot of research efforts have been devoted to develop new light-

gated channels for optogenetics.

Temperature-gated channels are found in the transient receptor potential (TRP) group, in

particular in the TRPV subfamily, and are responsible for the sensation of heat and pain, but

also for the regulation of body temperature.

The gating mechanisms allow the cell to change the permeability of its membrane in response to

either external or internal stimuli, exploiting the related variations in membrane potential and ion

concentrations to trigger different biophysical processes.

2.2.3 Channel conductance and temperature dependence

The conductivity properties of ion channels vary greatly from type to type. The introduction at the

end of the ‘70s of the patch-clamp technique by Sakmann and Neher19

allowed the recording of

currents through single channels and opened the way to a precise understanding of their electrical

characteristics. While some channels have linear voltage-current characteristics, others have been

seen to present rectifying behavior because of asymmetries in the structure of the pore and of the

selectivity filter. Typical conductivity values of a single ion channel can vary between 0.1 to 100

pS.147

The membrane conductivity is then given by the sum of the single conductivities of all the

open channels at a certain moment.

Since the transport of ion in the pore of a channel is determined by a trap-and-release mechanism

from the selectivity filter, the actual speed at which this happens is greatly dependent on

temperature, that provide the thermal energy for the releasing of the ion. Indeed, channels

conductivities have been experimentally measured to vary with temperature. A common way to

Page 28: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

23

give an estimate of such dependence is via the Q10 temperature coefficient, which measures how the

conductivity changes upon a temperature variation of 10 °C and is expressed by the following

formula:163

where G2 and G1 are the channel conductivities at temperature T2 and T1 respectively. Typical

values for the Q10 coefficient of ion channels are in the range between 1.2 and 1.6, but particular

channels with a coefficient of about 5-6 have been demonstrated.164

It is important to highlight that this temperature dependence is a general mechanism that applies to

all channels modulating their conductance, and it’s given by the fact that the ion transport is a

thermally activated process. The temperature-gating properties of channels like TRPV are a

different mechanism, in which the channel completely changes its conformation when a certain

temperature threshold is reached, passing from an open to a closed state or vice versa.

2.3 The membrane potential

The instauration of a potential across the cell membrane is the result of an electrochemical

equilibrium due to the combination of two essential factors that characterize the plasma membrane:3

An asymmetric distribution of ions between the intracellular and extracellular space. The main

ions that compose physiological media are K+, Na

+, Cl

- and Ca

2+, plus large anions (A

-)

constituted by proteins with charged residues. Of these main cations, the cytosol, i.e. the

electrolytic solution permeating the cytoplasm, is rich in potassium, while the extracellular

medium is rich in sodium. In normal conditions, the concentration of calcium ions is maintained

at very low values in the cytoplasm by active pumping towards the extracellular space or

sequestering in internal stores; the release of Ca2+

is in fact an important signaling event that can

trigger a variety of biological processes in the cell. For the anion, the charged proteins are

mainly confined inside the cell, since they cannot easily cross the membrane, while chloride is

usually found in the extracellular medium. Table 2.1 summarizes typical values for the

concentrations of the different ionic species for mammalian neurons in physiological

conditions.

The selective permeability of the membrane, that allows some ions to pass easily than others.

As described in the previous section, the membrane permeability to ions is mainly given by the

presence of ion channels and many of these channels are very selective only for a particular ion

(2.1)

Page 29: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

24

species. The actual permeability properties of a membrane are thus determined by the level at

which each channel is expressed, and this distribution can greatly vary from cell to cell. In

particular, in most cases the plasma membrane in normal conditions is mainly permeable to

potassium ions, while the conductances for other ions are lower. Gating mechanisms are

exploited to change the membrane permeabilities and thus modify the fluxes of different ions.

Table 2.1 | Typical values for ion concentration in the intracellular and extracellular

media for mammalian neurons, reproduced from Ref. 165.

2.3.1 Electrochemical equilibrium in biological membranes

To understand how these two factors determine the formation of the membrane potential, let’s

consider initially a situation in which only potassium conductances are present. The potassium

concentration in the intracellular compartment is much higher than that in the outer region ([K+]i >

[K+]o). Because of this concentration gradient, potassium ions tend to diffuse out of the cell, while

the movement of all other ionic species is hindered by the membrane selectivity. As K+ ions exit the

cell, the outer side of the membrane becomes positively charged, while the opposite happens on the

intracellular side. This charge accumulation at the two sides leads to the formation of a potential

difference across the membrane and thus an electric field that starts to drive back potassium ions

into the cell. A dynamic equilibrium is reached when the diffusional and electrical forces exactly

balance each other; the transmembrane potential at which this happens can be calculated from the

concentration of the ionic species in the intracellular [X]in and extracellular space [X]out with the

Nernst equation:3

Where R is the ideal gas constant, T the absolute temperature, F the Faraday constant and z the

valence of the ionic species (+1 in the case of K+). In the case of potassium ions, for the

concentrations reported in Table 2.1, the Nernst equilibrium potential Veq,K ≈ -89 mV.

If a cell has a plasma membrane only permeable to a single ionic species, the Nernst equation is

sufficient to determine its membrane potential. This is the case for example of several types of glial

Intracellular Extracellular

Potassium (K+) 140 mM 5 mM

Sodium (Na+) 5-15 mM 145 mM

Chloride (Cl-) 4-30 mM 110 mM

Calcium (Ca2+

) 100 nM 1-2 mM

(2.2)

Page 30: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

25

cells in resting condition. However, while usually cell membranes are mainly permeable to

potassium, also other conductances are expressed that allow the flow of lower but anyway sizable

currents of different ions. In these cases, the membrane potential assumes a value intermediate

between the Nernst equilibrium for the different ions involved, with the more permeable ones

having a higher weight in determining the final result. The precise value can be calculated in the

case of physiologically most relevant monovalent ions with the Goldman-Hodgkin-Katz (GHK)

equation:166

where the Pi terms refer to the membrane permeabilities for the different ionic species. For neurons

at resting conditions, for example, the typical relative ratios of the different permeabilities are:3

From the values of Table 2.1, sodium and chloride have Nernst equilibrium potential of respectively

Veq,Na ≈ 60 mV and Veq,Cl ≈ -70 mV (for [Na+]in = 15 mM and [Cl

-]in = 8 mM). While the main

contribution to the membrane potential is still given by the potassium ions, the actual equilibrium is

shifted towards more positive values for the presence of these other conductances to Veq ≈ -68 mV.

In these conditions, the single ionic species are not in equilibrium, since the membrane potential is

different from their Nernst equilibrium. In particular, there is a net efflux of K+ ions out of the cell,

since its diffusional driving force is not completely balanced by the electrical gradient, and an influx

of Na+ ions, since both electrical and chemical gradients tend to drive them into the cell. However,

these two currents balance each other to give a zero net current through the membrane. As for

chloride, in many cases its concentrations are such that its electrochemical equilibrium is very close

to the membrane potential, so the driving force acting on it is negligible; the actual Cl- currents

flowing through the membrane are thus usually quite small in resting conditions, even if the

membrane has a substantial permeability to this ion.

The membrane potential of a quiescent cell that is not subject to any particular stimulus is usually

referred as the resting potential. Its value can be different from cell to cell as it is determined by the

distribution of ion channels expressed in the membrane that are actually in the open state in this

resting condition. However, the membrane potential can be modulated by different factors and its

variations are involved in determining the cell response to internal and/or external stimuli. The main

pathway by which a stimulus can influence the membrane potential is through the gating

mechanisms described in the previous section. When an ion channel opens, the membrane

permeability changes producing a variation in its equilibrium potential. In electrophysiology, a

variation of the membrane potential towards more positive values with respect to the resting one is

(2.3)

(2.4)

Page 31: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

26

termed depolarization, while a change towards more negative values is referred to as

hyperpolarization. From the GHK equation it’s easy to infer that in normal condition, i.e. a cell at

rest with physiological concentrations of ions, the opening of sodium channels leads to a

depolarization of the cell, since the equilibrium potential is pushed more towards Veq,Na; conversely,

the opening of potassium channel is reflected in a hyperpolarization.

2.3.2 Action potentials in excitable cells

The prototypical example of this response is the generation of action potentials (APs) in neurons.4

These cells posses a significant number of voltage-gated sodium (NaV) channels that are closed in

resting conditions, but open in response to a depolarization of the membrane. Upon an appropriate

stimulus (for example the release of neurotransmitters by an adjacent synapse), specific receptor in

the neuron membrane are activated, producing a depolarization of the membrane. This

depolarization triggers the opening of the NaV channels, increasing the membrane permeability to

sodium, which in turn leads to a further depolarization. If the initial depolarization is high enough,

this mechanism becomes self-sustaining and the membrane permeability to sodium increases

exponentially and becomes 10/20 times higher than the one of K+. In this way, the membrane

potential quickly spikes to values close to Veq,Na. The NaV channels however stay opened for a very

short time and quickly enter a refractory state in which they are not sensitive to the membrane

potential. In the meantime, voltage-gated potassium (KV) channels, which have a slower activation

time, opens in response to the membrane depolarization. The closing of NaV channels and

concomitant opening of KV channels leads to a repolarization of the membrane that goes back

towards negative values and actually hyperpolarize since there is an increase in potassium

permeability. This fast spiking is called action potential and is the manner by which neurons

respond to stimulation and conduct information in neural networks. The firing of an APs is an all-

or-none response, which is triggered if the neuron is subjected to a depolarization exceeding a

certain threshold value. Generally, this threshold is reached if the cell is depolarized to potentials

more positive than -50 mV, from normal resting values of about -70 mV to -60 mV. Once the

threshold is reached, the action potential is fired and it evolves in a manner that is basically

independent on the actual intensity of the initial stimulus. The nature and magnitude of the

stimulation can however influence the frequency at which action potentials are generated. For this

reason information in neural network cannot be generally encoded in the intensity of the

propagating signals, but is mainly transmitted as a modulation in the frequency of APs.

While neurons are considered the prototypical example of systems that support the formation of

action potentials, they can also be found in other cell types, generally termed excitable cells. In

cardiac cells,5 APs are involved in the coordination of the heart contractions (pacemaking); in these

Page 32: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

27

cells the action potentials have a different time course with respect to neurons, with a plateau of

several milliseconds in the depolarized states before repolarization, given by the action of slower

voltage-gated calcium channel. In muscle cells6 the propagation of action potentials leads to an

increase of calcium ions concentration in the cytoplasm, which acts as a second messenger for

triggering the contraction of muscle fibers. A conceptually similar mechanism can be observed also

in endocrine cells,7 where the intracellular calcium concentration controls the release of hormones

by driving the fusion of vesicles to the plasma membrane.

In non-excitable cells, the lack of specific voltage-gated channels does not allow the instauration of

action potentials. However, also in these cells the membrane potential can vary in response to

external stimulations or internal processes. In this case, the variation of the potential is not an all-or-

none response, but is proportional to the stimulus intensity. These progressive responses that do not

involve the self-sustained opening of voltage-gated channels are called graded potential.3 Also

stimulations of excitable cells that do not trigger the firing of action potentials fall in this category,

like the postsynaptic potentials (PSP) produced in dendrites of neurons upon reception of

neurotransmitters. These stimuli can travel along the cell membrane until a specific region of the

neuron, the axon hillock, where they are summed up; it is in this region rich of voltage-gated

channels that, if the threshold is reached, the action potentials are fired.

2.3.3 Maintenance of ion concentrations

The correct functioning of a cell is tightly linked to maintenance of correct concentrations of ionic

species in both the intracellular and extracellular space. If only passive transport of ions through

channels in the membrane was present the ionic concentrations would rapidly change since, as

described in the previous section, even at equilibrium there is a net efflux of potassium ions and an

influx of sodium in the cell. Moreover, in excitable cells, the firing of action potentials is associated

with a substantial displacement of ions through the membrane. For this reason, cells have a series of

active transport proteins in the membrane that maintain the ionic concentrations at physiological

levels. These proteins use energy from different sources to move ions against their electrochemical

gradients and can move one or more different ionic species at a time.156

The energy needed can be

supplied as chemical energy by a molecule like ATP or by the movement of one ionic species along

its electrochemical gradients to move another species in an unfavorable direction.

The prototypical example of these systems is the Na+/K

+-ATPase, or sodium-potassium pump,

167 a

protein present in the plasma membranes of all animal cells that is fundamental in keeping the

concentrations of Na+ and K

+ in physiological ranges. Being an ATPase, it works using the energy

obtained from the decomposition of an ATP molecule into ADP. The conformational changes

Page 33: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

28

caused by the binding of an ATP molecule and subsequent release of ADP lead to the expulsion of

three Na+ ions out of the cell and the concomitant influx of two K

+ ions. Since for each cycle there

is a net efflux of positive charge, this pump is actually electrogenic, in the sense that it has a net

effect on the cell equilibrium potential, driving it towards more negative values with respect to the

one determined by only the diffusion processes considered in the GHK equation.

Figure 2.3 | Schematic representation of the functioning of the Na+/K

+ pump. (1) Three

sodium ions bind to the channel from the intracellular space. (2) An ATP molecule is

hydrolyzed by the pump, which upon phosphorylation undergoes a conformational change

that exposes the Na+ ions to the extracellular space; the affinity to Na

+ of the pump in this

new conformation is lower and the ions are thus released. (3) The pump binds two K+ ions

from the extracellular space, causing a dephosphorylation of the protein. (4) Upon

releasing of the phosphate, the pump reverts back to its original conformation which has a

low affinity for potassium ions, which are thus released in the intracellular space.

It is interesting to notice that some cells do not have active transporters for chloride ions.3 In this

case, the concentration of Cl- in the cell is only determined by the passive flow through ion

channels; it thus tends to reach a value such that its Nernst potential equals the one of the membrane

equilibrium potential determined by sodium and potassium.

2.3.4 Electrical equivalent of a cell membrane

In the previous sections we have seen that the plasma membrane is not just a barrier that separates

the cytoplasm from the extracellular space, but a complex system that actively regulates the

bioelectrical properties of the cell. The study of its behavior in response to different stimuli is thus

fundamental in understanding how a cell functions. While the GHK equation introduced in the

previous section gives useful information on the equilibrium state of the system, it is not suitable to

Page 34: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

29

investigate the dynamic properties of the cell. A convenient way to represent the cell membrane in

order to study how its potential can evolve in time is by developing an equivalent electric circuit of

the system. The different elements that make up the plasma membrane can in fact be represented as

electrical components bundled in a single circuit that reproduces its functioning.

The lipid bilayer is an insulating barrier that separates two conducting media, the intracellular

and extracellular electrolytic solution. It thus behaves as a capacitor, with the lipids being the

dielectric medium between the two conductive plates. The membrane capacitance can be

estimated as that of a parallel-plate capacitor, with the distance between the plates as the bilayer

thickness, which is usually between dlip = 3-4 nm and a relative dielectric constant on the order

of εlip = 2.5. This simple model gives a value for the specific capacitance of Cm ≈ 0.7 μF/cm2.

While this parameter is usually considered a biological constant, actual variations are measured

between different cells and in different regions of the same cell. This variability is given by the

different composition in terms of proteins content, which can modify the effective thickness of

the membrane. Moreover, the ions accumulated on the two sides of the membrane are not

exactly concentrated at the dielectric interfaces, but form diffuse layers extending into the

solution for several nanometers. In any case, experimental values are usually in the order of 1-2

μF/cm2.168

Many of the different molecules that make up the plasma membrane are generally charged in

physiological conditions. As already described in Section 2.1, the two leaflets of the membrane

have thus a surface charge and in particular the inner side is generally more negative than the

outer one. Moreover, the different composition of the extracellular and intracellular

compartments produces a different distribution of the ions in the diffuse layers at the two sides

of the membrane. These asymmetries effectively generate an intrinsic potential difference

across the membrane even if there is no charge accumulated on it or, conversely, there is an

accumulation of charges on the membrane capacitance even if the transmembrane potential is

kept at zero. This intrinsic potential can be modeled as a voltage generator (Vσ) in series with

the membrane capacitance.

Ion channels are conducting pores that let charges flow through the membrane and they can

thus be represented as resistances (RX) in parallel to the membrane capacitance. Since currents

across the membrane are carried by different ionic species, the various channels can be grouped

in different resistors each representing the membrane permeability to a particular ion. These

resistances, however, are not fixed, but can vary in time due to the gating mechanisms triggered

by the different stimuli to which the cell is subjected. Moreover, due to asymmetries in the

channel structure or in the ion concentrations on the two side of the membrane, some channels

do not behave as simple resistors (i.e. do not have a linear voltage-current relationship), but can

Page 35: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

30

present rectifying characteristics. Also voltage-gating of the channel conformations can result in

a non-linear dependence on the transmembrane voltage.

The difference in the distributions of ionic species on the two sides of the membrane acts as an

electromotive force that drives charges along their concentration gradient. Each of the resistor

for the different species is thus in series with a voltage generator, whose value is the Nernst

equilibrium potential (VX) for that particular ion.

Combining together all these elements, it is possible to come up with an equivalent circuit like the

one depicted in Figure 2.4a. The upper and lower nodes of the circuit represent the intracellular and

extracellular space respectively and the difference in their electrical potential is the actual

membrane potential, which is conventionally defined as Vm = Vin - Vout.

In the following chapters a simplified version of this electrical model is going to be used (Figure

2.4b). In particular, the different ionic conductances can be grouped into an effective membrane

resistance (Rm), driven by a single voltage source that represent the cell equilibrium potential (Veq)

as determined by the GHK equation.

Figure 2.4 | (a) Equivalent electrical circuit of the cell membrane with the different ionic

conductances explicitly modeled. (b) Simplified membrane circuit with only an equivalent

membrane resistance and equilibrium potential; in the scheme also the series resistance of

a patch pipette for electrophysiological measurement is represented.

It is important to remember that all the parameters in this equivalent representation are not fixed,

but depend strongly on the actual conditions of the cell. Apart from the channel resistances that can

be greatly modified by gating mechanisms, one of the main factors that influence the actual values

of the electrical parameters of the circuit is temperature. It has been previously discussed in Section

2.2 that charge transport in ion channels is a thermally activated process and also in the equation for

the Nernst equilibrium potential there is a direct dependence on the absolute temperature of the

system. Moreover, variations in the capacitance of the plasma membrane have been recently

Page 36: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

31

observed upon light-induced heating of cells and have been proposed as a stimulation mechanism

for cellular activity.58

Page 37: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

32

Chapter 3 – Hybrid interfaces characterization

Photovoltaic devices based on organic semiconductors have been widely investigated in the past

two decades, with power conversion efficiencies now exceeding 10 %.169

Very few reports are

instead present in literature of hybrid devices where the photoactive layer is interfaced with an

electrolytic solution. Previous works from our group demonstrated that these systems are able to

generate charges upon illumination with visible light and produce a photocurrent with an action

spectrum similar in shape to that of standard photovoltaic devices based on the same materials.125,140

The precise mechanisms behind these effects, however, were not fully elucidated. Narayan and

coworkers170,171

have also carried out extensive electrical characterizations of similar architectures,

but their works focused mainly on devices with active layers significantly thicker than those used by

us; they also employed different semiconducting materials, especially for the electron acceptor. In

any case, different reports have demonstrated the successful interfacing of these devices with

biological tissues;140–143

however, the complete understanding of the mechanisms behind

transduction of light stimuli into bioelectrical signal cannot be achieved without an in-depth

knowledge of the functioning of the polymer/electrolyte interface by itself.

In this chapter, a thorough characterization of hybrid polymer/electrolyte interfaces based on the

conjugated polymer poly(3-hexylthiophene-2,5-diyl) (P3HT) as the photoabsorbing material is

carried out. In order to better contextualize the results later presented, a first introduction on the

operation principles of standard organic photovoltaic devices is given in Chapter 3.1. The hybrid

polymer/electrolyte structure is then introduced in Chapter 3.2, with details on the devices realized

and measured in the subsequent experiments. In Chapter 3.3 the photophysics of the active layer is

investigated by means of optical spectroscopies, in order to assess the effect of the electrolyte

presence on the dynamics of photoexcited species. A complete electrical characterization of the

device is then carried out in Chapter 3.4. Finally, the thermal phenomena occurring at the device

surface upon photoexcitation are analyzed in Chapter 3.5.

3.1 Standard organic photovoltaic devices

The standard architecture of an organic photovoltaic (OPV) solar cell,172,173

in the simpler version, is

composed of an active layer, where incident photons are transformed to electron and holes,

sandwiched between two electrodes for charge extraction.

Page 38: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

33

The active layer is generally composed of two different p- and n-type organic materials for electron

and hole conduction respectively. The p-type material is usually a conjugated polymer that also acts

as the main photoabsorber. For the n-type phase, a fullerene derivative like phenyl-C61-butyric acid

methyl ester (PCBM) is normally used. Upon light absorption an exciton, i.e. a bound electron-hole

pair, is formed in the polymer. These excitons have binding energies on the order of 0.5 eV131,172

and thus thermal energy is not enough to break them into separated charges (as instead it happens in

inorganic semiconductors). Excitons diffuse in the p-type phase until they recombine; usual

diffusion lengths in polymeric semiconductors are on the scale of 10 nm.174,175

In order to break

down the exciton a driving force is thus needed; in the case of an organic solar cell, this is normally

given by the presence of the n-type material that acts as an electron acceptor. When an exciton

reaches an interface between the p- and n-type phases, the difference in the energetic levels for

electrons in the two materials can be enough to overcome the coulomb attraction between the

electron-hole pair and to obtain free charges.173

The precise photophysical mechanism (or

mechanisms) by which this separation occurs is actually still under debate in the scientific

community.176

In any case, the result is to have a hole in the p-type phase and an electron in the n-

type phase of the active layer; these charges are then free to move towards the electrodes, provided

that a continuous path can be found in the relative phases.

The first devices based on this principle were planar heterojunctions of p- and n-type materials,

similar to standard inorganic devices.177

However, the typical absorption lengths in conjugated

polymers are on the order of 50-100 nm, while exciton diffusion lengths are usually not higher than

10 nm. Only the photons absorbed very close to the interface with the n-type material are thus able

to efficiently be converted into charges, while all the other excitons recombine without contributing

to the final current. To overcome this problem, the bulk heterojunction architecture was

introduced;120

in this case, the p- and n-type materials are mixed together in a single layer composed

of an interpenetrating structure of the two, with phase separation on the order of the exciton

diffusion length.

Once charges have been separated, some driving force is needed in order to promote their migration

to the respective electrodes. Electrodes with different work functions are thus usually employed for

hole and electron collection respectively, producing a built-in electric field in the active layer that

directs the drift of charges. At least one of the two electrodes needs to be transparent, in order for

light to reach the absorbing layer; one of the most used materials is indium tin oxide (ITO), a

conductive oxide that presents a very good trade-off between conductivity and transparency. ITO is

usually used as an anode (i.e. the electrode extracting holes) in standard device architectures, but it

has been shown that it can also act as an electron collector.178

Complete photovoltaic devices

Page 39: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

34

usually have also electron and hole selective interlayers between the active material and the

collecting electrodes to help charges being collected at the proper electrode.

3.2 Hybrid devices structure

The devices used in this work are hybrid polymer/electrolyte interfaces, where the metal contact of

a standard organic solar cell has been replaced with an aqueous electrolyte, the typical environment

in which cells are cultured.125

The substrate is a thin glass slab (170 μm) coated with an indium tin oxide (ITO) layer. This

small thickness of the substrate is necessary when the samples have to be analyzed in the

electrophysiology setup, which is based on an inverted microscope with an objective with short

working distance. For specific measurements, also bare glass substrates are employed.

The active material is a thin film of P3HT alone or blended to form a bulk heterojunction with

PCBM. The materials are dissolved in chlorobenzene and the solution is deposited directly on

the substrate via spin-coating. The films obtained, depending on the concentration of the

solution and the spin parameters, have thicknesses that can range from tens to few hundreds of

nanometers. After deposition, all the films are annealed in air at 120 °C for 2 hours; this step

improves crystallinity of P3HT, but it serves also as a sterilization process, mandatory in

experiments with cells. In the following, if not specified otherwise, we refer to the

polymer/electrolyte interface to describe in general a device with an active layer made either

with by pristine P3HT or a P3HT:PCBM blend.

For specific control measurements, another material has been used as the photoabsorbing layer,

namely the photoresist MicroPosit® S1813

®. This material can absorb light in the visible (in

particular in the blue), but has insulating properties and it does not support the generation of

charges when illuminated.

The electrolytes (el) used are different depending on the experiment. Basic electrical

characterization of the interface is generally performed with a solution of sodium chloride at a

concentration of 200 mM, resembling physiological ionic strengths. For experiments with cells,

appropriate solutions are used, like Krebs-Ringer-Henseleit buffer (KRH) or complete cell

growth media.

Page 40: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

35

3.3 Spectroscopic characterization

In this section, the characterization of the optical properties of P3HT based films is presented. In

particular, we investigated the dynamics of the photoexcited species in this material on different

timescales by means of pump and probe spectroscopies. While the photophysics of P3HT and

P3HT:PCBM blends has been broadly studied in the past twenty years,179,180

the possible

interactions of these materials with a saline solution are still an open problem. Our main goal was to

assess if the photoexcited species in the active layer could interact with the saline solution to give

photochemical reactions.

3.3.1 Absorption and fluorescence

P3HT is a conjugated polymer that absorbs in the blue-green region of the visible spectrum, while

PCBM absorption band is shifted in the UV, with a small tail in the blue. The fluorescence spectrum

of P3HT is red-shifted with respect to the absorption, peaking between 650 and 700 nm; upon

addition of PCBM, however, singlet excitons are efficiently quenched and luminescence

significantly drops. Typical absorption and fluorescence spectra for thin films (d ≈ 70 nm) of P3HT

and P3HT:PCBM are reported in Figure 3.1.

Figure 3.1 | Absorption and fluorescence (PL) spectra for P3HT and P3HT:PCBM thin

films. The PL spectra are collected for an excitation at 510 nm.

In the absorption spectra, especially for the pristine polymer, the vibronic replicas of the main

electronic transition are visible, consistently with the semicrystalline nature of the polymer. The

shoulder at 605 nm is usually attributed to the formation of a lamellar structure with a π-π stacking

of the thiophene rings.181

The vibronic structure is also visible in the luminescence spectrum of

P3HT, while, as expected, the signal from the blend is significantly quenched. Upon addition of

PCBM, its absorption in the UV below 400 nm becomes clearly visible. Also, the vibronic structure

Page 41: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

36

of the P3HT becomes less accentuated, especially the shoulder at 605 nm, because of the presence

of the acceptor molecules that can lower the formation of P3HT crystallites.

In a previous study from our group, the effect of light, oxygen and water on the absorption and

luminescence spectra of P3HT was investigated in depth. In particular, it was demonstrated that

while absorption properties of the polymer are only marginally modified by the presence of oxygen,

the luminescence yield is greatly reduced, especially when the effect of oxygen is combined with

prolonged illumination by visible light.123,133

This effect was attributed to a reversible doping of the

polymer, due to the formation of a charge-transfer complex between P3HT and oxygen molecules.

While this doping mechanism does not break the conjugation of the polymer, conserving its

absorption properties, the charge-transfer complexes act as quenching sites for singlet excitons, thus

diminishing the fluorescence yield of the material. Interestingly, in the same work it was shown that

the presence of water, instead of ambient air, does not introduces significant differences,138

but

actually slightly reduces this doping effect, probably because a lower concentration of oxygen

molecules dissolved in water with respect to air or a different electrostatic coupling with the

environment.

3.3.2 Pump and probe spectroscopies

Optical spectroscopies can be employed to study the evolution in time of excited states in a

material. In particular, transient absorption (TA) techniques have been widely applied for the

investigation of charge generation and recombination processes in organic photovoltaics.182,183

These techniques are based on the pump and probe concept. The basic idea is to photoexcite the

material with a first beam of light resonant with one of its absorption transitions; the generated

photoexcited species modify the electronic transition of the material. A second beam of light is

employed to monitor the evolution in time of these variations, from which the dynamics of the

different excited states can be inferred.

The variations in the absorption spectra due to presence of excited states are quite small (usually no

more than few percents). The relevant signal that is recorded is the normalized differential

transmission (ΔT/T),183,184

i.e. the difference in the probe transmission signal in presence and in

absence of the pump, divided by the ground state transmission. This signal is measured for different

wavelength of the probe beam (TA spectra) and at different delays from the pump excitation (TA

dynamics). Three different kinds of phenomena can contribute to the final TA signal:

Ground-state bleaching (GSB, ΔT/T > 0). Excited species have different electronic transitions

with respect to the ground state. The absorption of the material at the wavelengths resonant with

Page 42: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

37

the ground-state transitions thus decreases in the presence of the pump beam, leading to a

positive differential transmission.

Stimulated emission (SE, ΔT/T > 0). When molecules are in singlet excited states, the

impinging probe photons at resonant wavelengths with fluorescence transitions can promote

stimulated emission from the material. Since more photons come out from the material with

respect to the incident ones, the differential transmission is again positive.

Photoinduced absorption (PA, ΔT/T < 0). The presence of excited species in the material

generates new optical transitions not present in the ground-state. At wavelengths resonant with

these new transitions, the differential transmission is thus negative, since absorption is

increasing in the presence of the pump beam.

Once the different bands appearing in the TA spectra have been assigned to the relative excited

species, their populations can be followed in time by looking at the TA dynamics at the relative

wavelengths. However, due spectral congestion, i.e. overlapping of bands arising from different

signals, single-wavelength dynamics can be made up of contributions from more excited species; in

these cases, extracting the temporal evolution of the different populations is not straightforward.

In the spectroscopic measurements that follows, the samples used are thin films of P3HT and

P3HT:PCBM deposited on glass substrates. The films have similar thicknesses of about 70 nm. To

study the effect of the electrolytic solution, we firstly recorded the measurements in air (these data

are referred to as “dry” condition); after that, the sample were put in contact with a saline solution

(NaCl 0.2 M in ultrapure water) and the measurements were performed again (“wet” condition).

3.3.3 Femtosecond transient absorption spectroscopy

Charge generation processes in semiconductors usually occur on the timescale of

femto/picoseconds, but electronic instrumentation is too slow to follow such fast dynamics. To

reconstruct TA dynamics on such ultrafast timescales, mode-locked lasers producing femtosecond

pulses are employed. Both pump and probe are ultrafast pulses, with the probe that is delayed with

respect to the pump varying its optical path by means of a translation stage. The probe transmission

through the sample can thus be measured at different delays to reconstruct the TA dynamics. The

time resolution of the measurement is given by the convolution of the pump and probe durations

and in standard setups is on the order of 100 fs. The maximum delays that can be measured depend

basically on the length of the translation stage and are usually in the range from hundreds of

picoseconds to few nanoseconds. A detailed review of ultrafast transient absorption setups can be

found in Ref. 184.

Page 43: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

38

The photophysics of P3HT has been widely investigated in literature.180,185,186

A typical visible TA

spectrum of regioregular P3HT (Figure 3.2a at different probe delays) presents a broad GSB

positive ΔT/T signal between 500 nm and 620 nm and a negative PA2 band peaking at 650 nm. The

PA2 signal has been assigned to interchain polaron pairs (or charge-transfer states), i.e. an electron-

hole couple delocalized between adjacent polymeric chains. These polaron pairs cannot easily

separate into free charges and thus recombine to the ground state in the first 100 ps after

photoexcitation. There is however a very small fraction of photoexcitations that are able to

overcome coulomb attraction and separate into free charges; these are responsible for the two small

PA1 and PA3 bands that become visible at 450 nm and 690 nm especially at longer time delays.

Figure 3.2 | Ultrafast pump-probe spectra on P3HT films in “dry” (a) and “wet” (b)

conditions, taken at different times after photoexcitation (100 fs – 300 ps).

Figure 3.3 | Comparison between the dynamics of the pump-probe signals at 560 nm and

650 nm for the samples in “dry” (red) and “wet” (blue) conditions. The traces of the “wet”

sample have been rescaled to the peak of the “dry” one, in order to have a clearer

comparison of the dynamics.

The presence of saline solution (NaCl, 200 mM) in contact with the polymer does not significantly

change the photophysical picture (Figure 3.2b). Apart from a small variation in the signal intensity,

Page 44: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

39

related to smaller pump intensity due to increased losses in the presence of water on the optical

path, the shape of the spectra are the same. Also the dynamics at the different relevant wavelengths

do not show any variation in the decay evolution in time (Figure 3.3).

Figure 3.4 | Ultrafast pump-probe spectra on P3HT:PCBM films in “dry” (a) and “wet”

(b) conditions, taken at different times after photoexcitation (100 fs – 300 ps).

Figure 3.5 | Comparison between the dynamics of the pump-probe signals at 560 nm, 650

nm and 690 nm for the samples in “dry” (red) and “wet” (blue) conditions. The traces of

the “wet” sample have been rescaled to the peak of the “dry” one, in order to have a

clearer comparison of the dynamics.

Upon addition of PCBM, some differences arise in the transient absorption spectra of the films. The

GSB signal is still present for wavelengths below 620 nm, but there is an increase of the signal in

the blue region of the spectrum, consistent with the steady-state absorption spectra reported in

Figure 3.4a. The PA2 band at 650 nm attributed to interchain polarons is still present; more

interestingly, the PA3 signal at 690 nm is here much more visible with respect to the pristine

polymer case (the PA1 band is however hidden by the high GSB signal in the blue). This increase in

the PA3 band is consistent with a much higher generation of free charges in the blend due to the

presence of the electron acceptor. It has also to be highlighted that, in contrast to the pristine case,

Page 45: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

40

even after 300 ps there is still a significant fraction of the photoexcited population that has not

recombined back to the ground state. This long-lived population is related to the presence of free

holes and electrons in the material that are spatially separated in the two different phases of the

materials (P3HT and PCBM), thus having a much longer lifetime compared to bound states.

However, also in the case of the P3HT:PCBM blend, no significant differences can be observed

between the “dry” and “wet” conditions, as it can be seen by the shape of the spectra in Figure 3.4b

and the comparison of the dynamics in Figure 3.5.

3.3.4 Nanosecond transient absorption spectroscopy

Next, we analyzed the behavior of the photoexcited species in the ns-μs regime. This is again a

pump-probe experiment, but given the different timescales involved, the setup implementation is

different. Since the operation frequency of electronic measurement instruments is abundantly

beyond the GHz, the temporal dynamics can in this case be followed in real time with an

oscilloscope. For the probe beam is thus sufficient in this case an incoherent lamp. The pump signal

instead, which needs to be shorter than the desired temporal resolution, is given by a Q-switched

laser (pulse duration ≈ 7 ns).

As it has been demonstrated above with ultrafast spectroscopy, only charged species are left in

P3HT after the first few hundreds of ps upon photoexcitation (if intersystem crossing towards triplet

states can be considered negligible).185,186

The dynamics of photobleaching and charge absorption

signals are thus the same and only one of the two needs to be measured in order to have a complete

picture. In Figure 3.6 the normalized temporal traces at 570 nm (in the P3HT bleaching band) are

reported for both P3HT and P3HT:PCBM samples in “dry” and “wet” conditions.

Figure 3.6 | Comparison of the normalized dynamics (presented as the difference in

optical density, ΔOD, in presence and absence of the pump) for the photobleaching band

at 570 nm of P3HT (a) and P3HT:PCBM (b) samples in “dry” (red) and “wet” (blue)

conditions.

Page 46: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

41

As it can be expected, the lifetime of charged species is longer in the blend sample with respect to

the pristine polymer. More interestingly, in both cases a slight difference can be observed between

the measurements in “dry” and “wet” conditions. However, this is to be attributed just to an

enhanced thermal conductivity of water with respect to air (kwater = 0.6 W m-1

K-1

, versus kair = 0.024

W m-1

K-1

).187

Indeed, similar variations in the dynamics have already been reported in literature to

occur upon changing the thermal conductivity of the substrate. In particular, faster dynamics were

observed in samples with a more thermally conductive substrate (namely sapphire, with respect to

standard glass).188

In our case, the same explanation can be applied, considering that water has a

higher thermal conductivity with respect to air.

3.3.5 CW photoinduced absorption spectroscopy

Free charges recombine on the nano-microsecond timescale; after that, however, in the material

some trapped species still remain up to the millisecond range. Since their number is usually quite

small, the relative variations in the transmission spectra are on the order of 10-4

. To detect such

small signals, an oscilloscope is not sensible enough and lock-in detection is needed. In this case,

both the pump and the probe are continuous-wave beams; the pump is periodically modulated,

usually with a mechanical chopper, and the corresponding variations in probe transmission are

extracted by the lock-in amplifier. This is not a transient technique and actually only information on

steady-state populations can be detected. However, information on the lifetimes can be obtained

performing a frequency-domain analysis varying the pump modulation.186,189

In pristine films of P3HT, detrapping of the trap states due to thermal energy quickly favors the

recombination of these species. At room temperature it is thus very difficult to detect a CW-PIA

signal.189

To slow down this process, cryogenic measurements are usually performed; however, if

the contribution of direct contact with water is to be assessed, this strategy cannot be applied. For

this reason, we could perform only measurements on P3HT:PCBM blends. In this material, the

higher number of charges generated and the spatial segregation between electrons and holes gives

higher steady-state populations that can be reliably detected.

Visible and near-IR spectra for both the “wet” and “dry” case of a P3HT:PCBM thin film are

reported in Figure 3.7. Apart from the bleaching at wavelengths shorter than 620 nm, a broad

negative signal is visible between 650 nm and 1050 nm. This signal is actually made up by two

different bands, PA1 and PA2 peaking at 690 nm and 980 nm respectively. These bands have been

assigned to two different types of polarons: PA2 band is related to intrachain polarons in the

amorphous fraction of the material; PA1 signal is instead attributed to delocalized polarons in the

Page 47: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

42

lamellar structures. In any case, also on the timescales no significant variation can be observed

between the “wet” and “dry” case.

Figure 3.7 | CW-PIA spectra in for P3HT:PCBM films in “dry” (red) and “wet” (blue)

conditions upon excitation at 560 nm.

3.4 Electrical characterization

The spectroscopic measurements of the previous section have shown that charge transfer reactions

from the photoactive material to the ionic conductor are not occurring, at least in a sizable amount.

It is thus very likely that, in contrast to the metallic electrode of standard OPV devices, the liquid

electrolyte does not behave as an ohmic contact for charge extraction. This observation suggests

that the functioning of these hybrid devices can be quite different from a standard solar cell.

Previous measurements from our group have demonstrated that P3HT in presence of oxygen, and

especially upon illumination, gets quickly p-doped via the formation of a charge transfer complex

between the polymer backbone and the O2 molecule.138

In particular, concentrations of free holes in

the range of Np = 1017

-1018

cm-3

have been estimated from capacitance measurements of devices

kept in electrolytic solution under illumination. Given these levels of doping, at the polymer/ITO

interface electrons should be injected into the polymer from the contact, creating an interfacial

dipole. This dipole can screen the internal electric field in the active layer, which in standard OPVs

drives the photogenerated charges towards the electrodes. Upon photoexcitation of the active

material, the absence of a net electric field in the bulk implies that in this region holes and electrons

do not get efficiently separated and should thus tend to recombine, not contributing to the final

photocurrent. Only in the interfacial region with ITO the electric field is able to separate

photogenerated electrons and holes, and to produce a sizable current. In particular, it should be

expected that electrons are collected by the ITO.

Page 48: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

43

In order to support this picture and to understand the actual coupling of the electrical signals

generated upon illumination in the active material with the ionic conductor, we have investigated in

more details the electrical functioning of the hybrid polymer/electrolyte systems. To simplify the

analysis, we start by studying the case of the pure P3HT active layer. In particular, we first

investigate the charge generation and separation processes in the material by measuring the

photovoltage produced in the ITO/P3HT/el device; we then study the coupling mechanisms at the

polymer/electrolyte interface with photocurrent and surface potential measurements. Finally, we

briefly compare the results obtained in the case of using the P3HT:PCBM blend as the active

material.

The following study is aimed at understanding the behavior of our hybrid devices upon illumination

with short pulses of light (i.e. few tens of milliseconds), since these are the relevant timescales when

capacitive stimulation of cells is considered as the final goal. The picture we draw here cannot thus

be transposed to the case of prolonged illumination, in which other physico-chemical phenomena

can intervene. Upon continuous illumination electrochemical reactions are indeed activated in these

hybrid interfaces, as it has been shown in previous works of our group;126,127

the description of these

mechanisms is however beyond the scope of this investigation, which is focused on the processes

occurring over short timescales.

3.4.1 Photovoltage measurements

We start the investigation of electrical properties of the hybrid polymer/electrolyte interfaces by

presenting an in-depth analysis of transient photovoltage measurements performed on ITO/P3HT/el

devices. Since we are dealing with an electrochemical system, in which both electronic and ionic

conductions are present, particular attention must be paid in the measurements. In particular, we

used for the recordings a three-electrode configuration. In this scheme, the current flows between

the working electrode (WE), which is the ITO/P3HT device we want to measure, and a counter-

electrode (CE) made of an inert material (in our case a platinum wire); however, the potential of the

WE is measured against a third electrode, the reference electrode (RE), whose potential is well-

known and stable. For our measurements, an Ag/AgCl electrode in a saturated KCl solution was

employed. The photovoltage transient generated at the WE is measured as the potential necessary to

counterbalance the current flowing through the CE, i.e. to keep the system in open circuit condition.

The complete measurement apparatus is described in more details in Appendix A.

The measurements presented in this section were performed by recording the photovoltage of

ITO/P3HT/el samples upon illumination with a 50 ms pulse of light from a collimated white LED.

The samples had approximately an area of 200 mm2 and were entirely illuminated by the spot; the

Page 49: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

44

electrolyte was a water solution of NaCl at a concentration of 0.2 M. A typical recorded trace is

reported in Figure 3.8a.

Figure 3.8 | (a) Photovoltage response measured on a P3HT sample upon illumination

with a 50 ms light pulse (represented by the light blue box). (b) Typical photovoltage trace

measured in a standard all-solid ITO/P3HT/Al photovoltaic device.

Before switching on the light, a constant potential Vdark is present. Its value is related to the

electrochemical equilibrium condition for the system with respect to the Ag/AgCl potential and is

subject to variations from sample to sample, with usually positive values ranging up to about 200

mV. This variability is attributed to small differences in the polymeric films, in terms of

morphology, defects and doping due to previous, unavoidable exposure to light and ambient

oxygen. Upon illumination, the potential has a downward dynamic that saturates after few

milliseconds at a plateau value (in this case, about ΔV ≈ -100 mV), consistently with our hypothesis

of electrons being collected at the ITO electrode.190,191

As the light is switched off, the system tends

to return to its original equilibrium condition, but, interestingly, this process is characterized by a

markedly longer time constant with respect to the onset of the excitation. It has also to be noted that,

during the light pulse, the potential does not remain constant at the plateau value and a slight

variation towards positive values can be observed. This process is probably related to the

instauration of electrochemical reactions at the polymer/electrolyte interface;126,127

since however

their effect is minimal over short timescales, we will neglect this component in the following

investigation.

As a comparison, the photovoltage generated in a standard all-solid device (ITO/P3HT/Al), in

which the liquid electrode is replaced with an aluminium contact, is shown in Figure 3.8b. In this

case, the signal is positive, consistently with the internal electric field in the bulk of the device that

drives the electrons towards the aluminium, while the ITO is here collecting the holes.

Page 50: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

45

In order to corroborate our hypothesis of an efficient generation of charges only in proximity of the

ITO interfaces, we recorded transient photovoltage measurements on ITO/P3HT/el samples of

different thickness and by changing the side from which light impinge on the device (i.e. from the

ITO or from the electrolyte). All measurements were performed at the same light intensity of 267.5

μW/mm2. From the traces, reported in Figure 3.9, the initial offset in dark condition has been

subtracted, in order to highlight the actual variation in the ITO potential upon illumination between

the different cases.

Figure 3.9 | Photovoltage responses of P3HT films with increasing thicknesses upon

illumination with a 50 ms light pulse (light blue box) from the ITO side (a) and the

electrolyte side (b).

Figure 3.10 | Dependence on thickness of the voltage peak (a) and rise time (b) of the

photopotentials measured upon illumination of the P3HT device from the ITO (blue) or

electorlyte (red) side.

As it can be expected, the peak potentials reached during the light pulse are quite dependent on the

film thickness; however, it is interesting to notice that this dependence is markedly different for the

two cases with different side of illumination (Figure 3.10a). In the case of light coming from the

ITO side, the potentials show a monotonic increase with the thickness of the active layer, reaching

Page 51: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

46

saturation at the higher values (d > 150 nm). In the case of illumination from the electrolyte,

instead, the photopotentials have a maximum for an optimal thickness (around ≈ 40 nm), and then

start to decrease as the films get thicker. It can also be noticed that, while the onset of the

photovoltage has a similar time constant in all the traces for the ITO illumination, in the other case

the building up of the signal gets slower for increasing dimensions (Figure 3.10b).

The data here presented are in good agreement with the photophysical picture proposed before.

When light is impinging from the ITO side, it is mainly absorbed close to this interface. The local

electric field can efficiently separate the hole and electron pairs, with the electron being collected at

the ITO and the building up of the photopotential. The higher thicknesses of the devices initially (d

< 50 nm) are reflected in a higher number of absorbed photons. However, upon a further increase of

the dimensions this improvement is less important because of two concomitant factors. Firstly,

according the Lambert-Beer law the highest fraction of light is absorbed in the first few tens of

nanometres (the P3HT absorption length of about 100 nm); increasing further the device thickness

does not lead to a sizable gain in the number photons absorbed. Secondly, the photons absorbed

further from the ITO interface than the region where the electric field is present do not contribute to

the final signal in a significant manner. The situation is different when the illumination comes from

the electrolyte side. In this case, the majority of photons are absorbed close to the P3HT/el interface.

For thin samples (d < 50 nm) this is not an issue, because the absorbed photons are anyway also

close to the ITO/P3HT interface and the photoexcited electrons in the active materials can still be

efficiently collected at the electrode. However, as the films get thicker, the number of photons

absorbed in the region with a sizable electric field decreases and the generation of the photovoltage

accordingly becomes less efficient.

We also performed measurements in dependence of the light intensity impinging on the sample.

Figure 3.11 reports different photovoltage transients recorded with illumination (from the ITO side)

ranging from 4.7 μW/mm2 to 197.5 μW/mm

2. The peak potential, as expected, increases with higher

intensities; in particular, a logarithmic dependence of the voltage can be observed (Figure 3.12a).

Moreover, also in this case an interesting slowing down of the ITO charging is observed when the

number of photogenerated charges is lower (Figure 3.12b), with a time constant for the

photovoltage formation that quickly drops of more than an order of magnitude (from 34 to 2 ms).

The logarithmic increase of the voltage with light intensity is an indication that the system behaves

as a rectifying junction. Indeed, the contact between P3HT and different metals (among which also

ITO) has been studied in details and it has been seen to form a Schottky junction.192–195

The strong

dependence of the lifetimes on the light intensity can be explained by considering that the

photocurrent flowing through the junction is, at least in a first approximation, proportional to the

number of absorbed photons, while the junction potential increases only logarithmically. In

Page 52: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

47

particular, to increase the potential is necessary to charge the junction capacitance and thus to

accumulate a charge that is proportional to the photopotential. Also this charge has a logarithmic

dependence on the light intensity, but since the photocurrent increases linearly with it, at higher

intensities a shorter time is needed to charge the junction. The same reasoning can also be applied to

the case of the increasing time constant with device thickness for illumination from the electrolyte

side, since also in this case what is actually happening is that less photon are absorbed in the

junction region.

Figure 3.11 | (a) Time traces of photovoltage measurements performed on a P3HT sample

(d ≈ 65 nm) upon illumination with a 50 ms pulse at increasing light intensities.

Figure 3.12 | Intensity dependence of the photovoltage peak (b) and rise time (c).

The photovoltage measurements reported above give us information mainly on what happens at the

ITO/P3HT junction upon illumination. In the discussion of the data, we have not considered

eventual effects occurring at the P3HT/el interface. This assumption is supported by the following

measurements (Figure 3.13), in which the photovoltage has been measured upon changing the

concentration of the ion species in the electrolyte. In particular, the recordings have been performed

with NaCl at 200 mM, 20 mM and in ultrapure water. The peak potential reached in all cases

remains basically the same, suggesting that the ion species in the electrolyte are not involved in its

Page 53: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

48

formation. Interestingly, only at the higher concentration the photovoltage signal is seen to slightly

tend towards positive values as the illumination continues, confirming our initial hypothesis that

this process can be attributed to some phenomena occurring at the P3HT/el interface. However, for

what concerns the initial negative photovoltage peak, its independence from the ionic strength of the

solution is a strong indication that the P3HT/el interface does not give significant contributions to

its formation and it is thus mainly determined by the photophysics of the ITO/P3HT junction.

Figure 3.13 | Dependence of the photovoltage signal on the concentration of the

electrolyte solution, measured on a P3HT sample (d ≈ 65 nm) for illumination from the

ITO side.

3.4.2 Photocurrent measurements

While the photovoltage measurements presented in the previous section give an exhaustive picture

on the functioning of the ITO/P3HT interface, they do not provide direct information on the effects

(at short times) of this photoexcitation on the device surface, which is the region where cell are

eventually cultured. In this section we present photocurrent measurements on the ITO/P3HT/el

devices, in which the ITO contact is short-circuited with a platinum counter-electrode in solution

(details on the experimental setup are given in Appendix A). Figure 3.14 shows the traces recorded

for a ≈ 70 nm P3HT film upon illumination from the ITO side with pulses of light (λ = 530 nm) at

different light intensities, ranging from 15.4 to 384 μW/mm2.

The positive signal upon illumination is consistent with the current flowing from the ITO/P3HT

device to the counter-electrode through the solution, i.e. with electrons from the polymer being

collected at the ITO as demonstrated with photovoltage measurements. As expected, the recorded

signals increase with light intensity but, interestingly, the photocurrent decays back to zero quite

rapidly during the light pulse, with a decay time constant that is on the order of few milliseconds,

decreasing for higher illuminations. In a conventional solar cell the current produced upon

illumination is, at least ideally, constant over time; the fact that in the ITO/P3HT/el device it drops

to zero after few milliseconds is again an indication that this system behaves differently with respect

Page 54: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

49

to standard OPVs.196

In particular, this behavior is an indication that there is an element in the

device structure that tends to block the flowing of current at steady state. Consistently with our

previous observations that there are no significant redox reactions taking place, on the investigated

timescales, at the P3HT/el interface, it is thus reasonable to conclude that this interface behaves,

electrically, as a capacitance (Figure 3.15). In electrochemical terms, we can say that the

photocurrent measured in this system is purely capacitive, with no appreciable faradaic

contributions.

Figure 3.14 | Photocurrent measurements on a P3HT film upon illumination with a 50 ms

pulse of light (light blue box) at increasing light intensities.

This finding is actually quite relevant in the context of using such hybrid devices for the

photostimulation of biological systems: eventual redox reactions could in fact lead to the generation

of harmful chemical species for the cell like oxidizing agents. Moreover, irreversible chemical

reactions would compromise the device stability in time. On the downside, this capacitive behavior

of the interface limits the total quantity of charge that the device can displace upon illumination:

once the P3HT/el capacitance has been charged, the current stops flowing through the device even

if the light is still on.

It is thus clear that one of the most important parameters that determine the functioning of the

device is the P3HT/el capacitance. This capacitance is given by the double layer of charges

accumulated on one side in the polymer at its surface and on the other in the diffuse layer of ions in

the electrolytic solution.170,197

The typical capacitance value that we measured, by means of

impedance spectroscopy, for P3HT is around Cint = 2 μF/cm2.197

This value is actually about one

order of magnitude higher than those usually found for standard inorganic devices based on silicon

used for capacitive stimulation. Such an increase is mainly due to the fact the that polymer surface

is in direct contact with the electrolyte in our hybrid devices, while in the case of silicon, an oxide

layer is always present between the semiconductor and the aqueous medium.91,102

This oxide layer

can have a thickness ranging from tens to hundreds of nanometers and has thus a significant effect

Page 55: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

50

in reducing the interface capacitance (a 10 nm layer of SiO2 has a geometric capacitance of Cox =

0.34 μF/cm2).

Figure 3.15 | Schematic representation of the ITO/P3HT/el architecture. The green box

represent the illuminated region. Cint: interfacial capacitance; Rbulk: resistance of the

polymer bulk; Iph: photocurrent generated at the Schottky junction between P3HT and

ITO.

3.4.3 Surface potential measurements

The measurements of the previous section show that the ITO/P3HT/el device is indeed able to

generate a photocurrent, even though transient, in short-circuit conditions, i.e. with the ITO contact

connected to a counter-electrode in the electrolyte. However, when they are used as active

substrates for cell culture and stimulation, these devices work in a different configuration:140–142

(i)

the ITO electrode is not directly contacted with an external circuit; (ii) only a small fraction of the

active area is illuminated (ideally, it would be desirable to address single cells, thus using a spot of

about 10-20 μm of diameter).

In order to understand if the capacitive charging of the P3HT/el interface occurs also in these

conditions, we measured the local potential generated at the surface of the device in conditions

similar to those used in experiments with cells. The sample was put in a petri-dish and immersed in

the electrolytic solution (NaCl 0.2 M); the local surface potential was measured with a glass

micropipette electrode positioned in close proximity (≈ 2 μm) of the P3HT interface. The device

was illuminated from the bottom (i.e. the ITO side) through a 40x microscope objective; the light

spot had a diameter of 540 μm. A more detailed description of the experimental method is reported

in Appendix A. The recorded traces for illumination with a 50 ms light pulse (λ = 470 nm) at

different intensities are reported in Figure 3.16a.

These measurements show that, upon photoexcitation, there is a flow of charges in the device even

if the ITO electrode is not contacted. This current produces a variation of the potential at the

Page 56: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

51

P3HT/el surface. Also in this case, because of the capacitive nature of the interface, the current

drops to zero quite rapidly, in few tens of milliseconds. In order for this current to flow, the

electrical circuit between the ITO layer and the electrolyte needs to be closed. Since no external

contact is made in these measurements, there has to be a parasitic coupling between the two, as

depicted in Figure 3.17a. This coupling may be due to a parasitic capacitance between the ITO

electrode and the solution or to physical contact between the two at the borders of the device or

cracks and defects in the polymer layer.

Figure 3.16 | (a) Local potentials measured at the P3HT/el interface with a glass

micropipette electrode upon illumination with a 50 ms light pulse (light blue box) focused

on the active surface (A ≈ 0.23 mm2) at increasing light intensities. (b) Surface potential

measurement on a glass/P3HT/el samples without the ITO contact.

In order to confirm the role of the ITO layer in the establishment of this parasitic coupling, we

measured the dependence of the surface potential signal from the lateral dimension of the ITO

contact. We realized ITO patches of increasing areas (from 0.37 to 18 mm2, in any case bigger than

the illumination spot size) on a glass substrate by selective etching with HCl of the conductive

oxide before the deposition of the P3HT active layer. The plot of Figure 3.17b shows that indeed the

peak of the surface potential measured depends on the dimension of the ITO contact underneath,

with a saturation for areas larger than ≈ 4-5 mm2. This result supports the hypothesis of a parasitic

coupling that increases with larger ITO electrodes. For small ITO areas, the low value of this

parasitic capacitance limits the current that can flow in the circuit; when the electrode gets bigger,

however, the current is no more limited by this coupling, but by the actual capacitance of the

polymer/electrolyte interface, and thus the signals does not grow any larger.

It is thus clear that the ITO contact is fundamental for the electrical charging of the device, because

it promotes the formation of a photovoltage at the junction with the ITO but also because in non-

contacted conditions it is fundamental for having the parasitic coupling with the solution that closes

the circuit, allowing the formation of a surface potential at the device interface. Interestingly,

Page 57: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

52

surface potential measurements performed on devices in which the active layer is deposited on bare

glass substrates without the ITO layer did not give any signal, confirming that in this situation the

capacitive charging of the interface is hindered (Figure 3.16b).

Figure 3.17 | (a) Electrical scheme of the parasitic coupling between the ITO electrode

and the electrolytic solution in the surface potential measurements. Rel: resistance of the

electrolytic solution; Cpar: parasitic capacitance; RE: reference electrode. (b) Dependence

of the surface potential signal on the dimensions of the ITO electrode; the data represent

the peak voltage measured on the onset of the light pulse (λ = 470 nm, I = 235 μW/mm2).

3.4.4 Measurements on P3HT:PCBM

We conclude the characterization of the electrical properties of the hybrid interfaces by analyzing

their behavior when the active layer is replaced with a P3HT:PCBM blend. In this case, the charge

generation processes in the material are favored due to the presence of the electron acceptor, as also

shown by the spectroscopic data presented in Section 3.3. The photovoltage, photocurrents and

surface potential measurements for P3HT:PCBM films of ≈ 50 nm of thickness are reported in

Figure 3.18.

Figure 3.18 | (a) Photovoltage measurements of a P3HT:PCBM film upon illumination

with a 100 ms light pulse. (b) Photocurrents recordings for a P3HT:PCBM film

illuminated with a 20 ms pulse at increasing light intensities. (c) Surface potential elicited

at the P3HT:PCBM/el interface upon illumination with a 20 ms pulse of light at different

intensities.

Page 58: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

53

Qualitatively, the data recorded on the blend are similar to the case of the pristine P3HT active

layer, suggesting a similar functioning mechanism of the hybrid interface. However, it is evident

from the traces that the blend presents higher signals, consistently with the increased charge

generation efficiency. This increase is also accompanied by faster dynamics with respect to the

P3HT case, due to the fact that, with higher currents, the surface P3HT/el capacitance gets charged

more quickly. Also in the case of P3HT:PCBM, devices without the ITO contact did not give any

surface potential signals.

3.5 Thermal characterization

Upon illumination of the active material, the photogenerated excitons can recombine or be

dissociated into charges; these charges, however, are not extracted from the device. After an

equilibrium situation has been reached with the charging of the interface capacitance, all

photogenerated charges will recombine. Given the low fluorescence yield of P3HT, especially in the

doped state,138

recombination of excitons and charges occurs mainly via non-radiatively pathways.

Most of the energy of the photons absorbed by the material is thus transformed into vibrational

energy, i.e. into local heating of the active layer. This thermal energy is then dissipated through both

the substrate and the electrolyte, with a local increase in temperature.

3.5.1 Local temperature measurements

In order to assess the thermal effects of photoexcitation of the active layer, we have measured the

local temperature variations of the bath at the polymer/electrolyte upon light pulses of different

duration and intensity. The measurements were carried out with the technique of the calibrated

pipette.198

This technique exploits the temperature dependence of the resistance of a glass

micropipette filled with an electrolytic solution (Figure 3.19). If this relationship is known, the

temperature evolution in the close proximity of the micropipette tip can be known simply by

measuring the current flowing through the pipette for a fixed applied voltage.

The measurements were carried out in the standard electrophysiology setup described in Appendix

A. Aqueous solutions of sodium chloride at 200 mM were used as both the bath and the pipette

filling media, in order not to have liquid junction potentials arising at the pipette tip. To calibrate the

pipette resistance, we measured the current response flowing through the pipette for a potential step

of ΔV = 5 mV at different values of the bath temperature controlled with an external heater (Figure

3.20a). Plotting the measured currents against the temperature in an Arrhenius plot results in a

straight line relationship (Figure 3.20b), from which a constant activation energy Ea = 11.7 kJ/mol

Page 59: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

54

for the process can be extracted. The relationship between current and temperature can thus be

expressed in the following form:198

where I0 is the current flowing through the pipette at a base temperature T0 and R is the ideal gas

constant.

Figure 3.19 | Typical variation of the resistance of a pipette tip placed in proximity of a

photoactive interface upon illumination with a 200 ms pulse of light (light blue box).

Figure 3.20 | (a) Current responses of the pipette upon the application of a 5 mV potential

difference with respect to the reference electrode at different temperatures of the bath. (b)

Arrhenius plot for the temperature dependence of the current, with a linear fit to extract

the activation energy Ea.

We measured the temperature variations at the polymer/electrolyte interface upon illumination for

thin films deposited on glass substrates of different active materials: P3HT, P3HT:PCBM and

Photoresist. The deposition processes for the three samples were carefully controlled in order to

obtain films with similar light absorption coefficient at the selected excitation wavelengths (λ = 475

(3.1)

Page 60: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

55

nm for P3HT and P3HT:CBM, λ = 435 nm for Photoresist, see Figure 3.21). The pipette tip was

micromanipulated in close proximity of the interface (≈ 2 μm) and an offset potential was applied

with respect to the reference electrode in order to have a current of about I0 = 4 nA at the base

temperature of T0 = 23 °C. The time evolution of the current was then recorded upon illumination of

the substrate with pulses of 20 ms and 200 ms at different light intensities. The temperature profile

was finally extracted with the formula of Equation (3.1).

Figure 3.21 | Absorption spectra for the P3HT, P3HT:PCBM and Photoresist films used

to measure the local temperature increase upon illumination. The blue and light blue boxes

represent the two wavelengths used to excite Photoresist (λ = 430 nm) and P3HT-based (λ

= 475 nm) samples respectively. Different wavelengths were chosen in order to have

comparable absorption in the three samples.

The recorded traces (Figure 3.22) show similar dynamics for all the investigate samples, indicating

that they posses similar properties in terms of heat generation. When light is switched on,

temperature increases monotonically with a decreasing slope. Upon illumination, heat is generated

in the active material and transferred to the bath in proximity of the interface; as time passes,

however, this thermal energy starts to be dissipated away from the surface and the temperature

increase becomes slower. As the light pulse ends, temperature starts to decrease until the base value

is finally reached.

The measured increases in temperature scale linearly with light intensity; in particular, at the

maximum light intensity used, temperature variations of about 3 °C and 7 °C are measured at the

end of the 20 ms and 200 ms light pulses respectively.

Page 61: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

56

Figure 3.22 | Local temperature dynamics in close proximity of the device surfaces for

P3HT (left), P3HT:PCBM (center) and Photoresist (right) upon illumination with light

pulses (represented by the blue and light blue boxes) of 20 ms (top) and 200 ms (bottom)

at different intensities.

3.5.2 Numerical simulations

In order to corroborate the origin of the recorded signal as arising from local temperature variations,

we conducted a numerical simulation of heat generation and diffusion in the system. A classical

heat diffusion model was implemented with the geometry depicted in Figure 3.23, with the

following hypotheses:

1. All energy absorbed by the active layer is transformed into heat, consistently with the fact that

virtually no charges are extracted from the devices and luminescence yield is negligible.

2. The bottom surface of the substrate is thermally isolated from the sample holder.

3. The simulated domain of the bath is large enough that the temperature at its extremes can be

considered constant and equal to the base temperature.

Page 62: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

57

Figure 3.23 | Scheme of the geometry used for the numerical solution of the heat diffusion

problem to calculate the temperature dynamics at the polymer/electrolyte interface (the

drawing is not in scale).

Given the circular section of the light spot (radius 270 μm), a cylindrical symmetry was assumed in

solving the problem. Given the similar experimental results obtained for the three different

substrates, the case of P3HT was considered in the simulations. Actually, given the very small

thickness of the active material with respect to the other two domains (glass and water), no

significant effect on the results is expected by small variation of its thermal parameters. The

physical parameters for the different materials (glass, water and P3HT) were taken from literature

and are summarized in Table 3.1.

Table 3.1 | Values of the parameters used for the resolution of the numerical model of

heat diffusion. cp: specific heat at constant pressure; ρ: density; k: thermal conductivity; α:

absorption length; t: thickness.

The temporal traces of temperature variation in the bath in close proximity with the surface of the

device were extracted from the simulation and compared with the experimental results (respectively

open circles and solid lines in Figure 3.24). A very good agreement can be seen between the two,

validating the experimental calibrated pipette method used for measuring the local temperature.

Glass Water P3HT

cp [J kg-1

K-1

] 840 4181.3 1400 [199

]

ρ [kg m-3

] 2500 1000 1100 [200

]

k [W m-1

K-1

] 1 0.6 0.2 [201

]

α [cm-1

] - - 105

t [μm] 170 2000 0.15

Page 63: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

58

Figure 3.24 | Comparison between the experimentally measured temperature dynamics on

a P3HT substrates (open circles) and the results of the numerical simulations (solid lines)

upon stimulation with 20 ms (a) and 200 ms (b) light pulses (represented by the light blue

boxes) at different intensities.

The spatial distributions of temperature as obtained from the numerical simulation, at different

times during and after a 200 ms pulse, are reported in Figure 3.25. They show that the temperature

increase is an effect basically localized to the illuminated area of the device and its immediate

surroundings, quickly decaying in few hundreds of micrometers from the interface, with the bulk of

the electrolyte remaining at the base temperature during the light pulse.

Figure 3.25 | Distribution of the local temperature values in space (along the radial and

axial direction) at different times during and after a 200 ms pulse (switched on at t = 0 ms)

of 57 mW/mm2. The vertical light blue lines separate the light by the dark region, while

the horizontal dashed lines is located at the position of the thin-film of active material.

Page 64: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

59

Chapter 4 – Coupling hybrid interfaces with cells

In Chapter 3 we have characterized the functioning of the polymer/electrolyte interfaces in terms of

the physical phenomena occurring upon photoexcitation. In particular, we have observed that two

different effects can take place at the surface of these hybrid devices: (i) a capacitive charging of the

interface that leads to a superficial transient potential; (ii) a local heating of the electrolyte following

the thermalization of the photoexcited species in the absorbing layer.

In this chapter we try to understand if and how these two effects can be exploited to modulate the

bioelectrical activity of cells. In particular, we are interested in studying the variations of the

membrane potential of cells grown on the active surface upon illumination. Our previous

publications have demonstrated that such interfaces are actually able to stimulate electrical activity

in neurons,140,141

astrocytes142

and explanted retinas.141

However, they are quite complex biological

systems and it is difficult to assess the direct effects of the polymer-mediated stimulation and

decouple them from the active responses they trigger. We thus decided here to use simpler cellular

systems to understand the actual effects of photostimulation on the membrane properties of the cell.

The cells used for this investigation are introduced in Chapter 4.1, along with a brief introduction on

experimental techniques used for measuring their bioelectrical properties. In Chapter 4.2 we

conduct experiment of photostimulation on these cells and we observe that both thermal and

electrical effects can take place in modulation of their membrane potential. Chapter 4.3 is dedicated

to an in-depth analysis of the thermal mechanisms and their influence on the membrane electrical

properties. Finally, in Chapter 4.4 we discuss about how the capacitive charging of the interface can

affect the membrane potential.

4.1 Human Embryonic Kidney (HEK) 293 cells

The previous works published by our group on the subject of photostimulation of biological systems

with P3HT-based hybrid interfaces were mainly focused on the study of neurons.140,141

These cells

are the main active elements of the nervous system, where they transport and elaborate the electrical

signals that control virtually all the vital functions in the body. Neurons are thus the main target in

the development of brain-machine interfaces to sense and control the functioning of living

organisms.90,91

The response of a neuron to a stimulus is generally represented by an action potential

(see Chapter 2.3); this rapid variation of the membrane potential is however a codified response,

Page 65: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

60

which is basically the same independently on the origin or the intensity of the stimulus, as long as a

certain threshold is reached. It is thus difficult to directly correlate the response of a neuron to a

specific cause and usually pharmacological modulation of different electrophysiological properties

of the cell is needed to address the problem.

In order to have a clearer understanding of the effects of polymer-mediated photostimulation on the

electrical properties of cellular membranes we thus decided to employ a simpler biological system.

Although only neurons and few other excitable cells (like myocytes and endocrine cells) are able to

sustain the firing of action potentials, all cells (for simplicity’s sake, we refer here to animal cells)

possess a plasma membrane and maintain a potential difference across it. The cells we selected for

our investigation are Human Embryonic Kidney 293 (HEK-293) cells.202

As their name suggests,

these cells have been isolated from the kidney of a human embryo in 1973 in the laboratories of

prof. Alex van der Eb (Leiden, The Netherlands) and have been subsequently genetically modified

to yield an immortalized cellular line. These cells are widely used in cell biology due to their

easiness in handling and can be grown in adhering cultures on different substrates. In particular,

they have found a lot of applications in electrophysiology for different reasons:202

(i) they are easy

to transfect; (ii) they are non-excitable cells that express quite low intrinsic conductivities and thus

can be used as a platform to study the behavior of exogenously expressed ion channels; (iii) they

have a compact shape with few processes, reducing space-clamp artifacts in electrophysiological

measurements.

Apart for their easy handling and their suitability for electrophysiological measurements, we

decided to use this particular cell type as a simple model of a cellular plasma membrane with a

small intrinsic conductivity.203

Our goal is thus to understand the actual biophysical mechanism that

couples the thermal and capacitive effects occurring at the polymer/electrolyte interface to the

electrical properties of the cell membrane. In principle, other simpler, non-living systems could be

used for this purpose, for example suspended lipid bilayers204

or vesicles.89

However, these artificial

systems lack all the complex machinery of a real cell that mediate the adhesion to the substrate.

With these lipid-based systems we would thus probably have a quite different interaction with the

substrate, both in terms of distance and composition of the cleft (i.e. the space between the surface

of the device and the basal membrane of the cell). HEK-293 cells can be considered as a good

compromise between a complex biological cell like a neuron and a too simplistic artificial system

like a lipid vesicle.

The experiments reported in this chapter on the photostimulation of HEK-293 cells, while

interesting on their own, cannot be directly transposed to more complex systems like neurons and

retinas. Different cells have different compositions of the membrane, both in terms of the lipid

content and of the actual transmembrane proteins expressed (especially ion channels);147

the effects

Page 66: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

61

of photostimulation can thus be in principle different on different cellular types. However, the

understanding of the basic biophysical mechanisms occurring in HEK-293 cells can be used as a

foundation upon which building the investigation of more complex systems. We will comment

more on this important issue in Chapter 5.

4.1.1 Cultures of HEK-293 cells on polymeric substrates

Before starting with electrophysiological investigation of the photostimulation of HEK-293 cells, it

is necessary to assess if they can be cultured on our interfaces based on semiconducting polymers.

The biocompatibility of P3HT-based thin films has actually been already investigated in the past on

different cellular types like fibroblasts, neurons and astrocytes (see Chapter 1.4).

Given the hydrophobicity of P3HT films, cells cannot be grown directly on them. An interlayer is

thus usually needed in order to promote cellular adhesion. Among the different molecules that can

be used for this purpose, we opted to use fibronectin, a glycoprotein of the extracellular matrix that

had already been investigated by Scarpa et al.139

for adhesion of cells on P3HT. The devices on

which we cultured HEK-293 cells are the same described in Chapter 3.2 and are basically composed

of a thin glass substrate (170 μm) with or without an ITO coating, on top of which a thin-film of the

active material (P3HT, P3HT:PCBM or photoresist) is spin-coated. The procedure for cell culturing

is reported in the following.

1. The devices are first sterilized in an oven at 120 °C for 2 h. This step is crucial to eliminate

bacterial contaminations that can hamper the growth of animal cells. From this point on, all

handling of the substrates needs to be done in sterile conditions under a biological hood.

2. The devices are put in a petri-dish or a multiwell and a solution of fibronectin in PBS

(Phosphate Buffered Saline) at a concentration of 2 μg/ml is casted on its surface. In order to

keep the solution on the hydrophobic substrates, PDMS (polydimethylsiloxane) wells are used.

The devices are incubated at 37 °C for at least 30 minutes to promote adhesion of the

fibronectin to the surface of the polymeric film. The substrates are then rinsed with PBS.

3. After the deposition of the fibronectin layer, HEK-293 cells are seeded on the surface of the

devices at a desired density (number of cells per cm2) in complete growth medium (see

Appendix A). The cells are then incubated at 37 °C and 5 % CO2.

In normal conditions, HEK-293 cells adhere to the substrate in few hours after seeding. If the

seeding density is not too high, they can be initially found as single cells; however, they quickly

start to replicate until a continuous monolayer covering the entire surface is formed. Depending on

the initial density, this process occurs in few days up to about a week.

Page 67: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

62

In order to have more quantitative data on the possibility of culturing HEK-293 cells on P3HT-

based surfaces, we performed the MTT assay for cell viability. This test is based on the reduction of

the tetrazolium dye 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) to its

formazan form, which is insoluble and has a purple color. This reduction is operated in living cells

by enzymes involved in mitochondrial activity.205

The quantity of the MTT that is reduced to the

formazan form is thus dependent on the number of living cells on the substrates and can be

measured optically by collecting its absorption in at a selected wavelength (we used λ = 570 nm, see

typical absorption spectrum in Figure 4.1a). By repeating the assay at different times after seeding,

growth curves can be obtained to assess if a particular material is better or worse in promoting cell

culturing with respect to a control substrate. The growth curves measured on P3HT and

P3HT:PCBM films are reported in Figure 4.1b for 1, 2, 3, 4 and 7 DIV, compared to standard glass

substrates (also coated with fibronectin). The three curves show comparable growth rates,

confirming previous literature data on the biocompatibility of these materials.

Figure 4.1 | (a) Typical absorption spectrum of MTT in ethanol extracted from cells to

evaluate their vitality; the green bar (λ = 570 nm) represents the wavelength used to

evaluate the absorbance. (b) Growth curve up to 7 DIV for HEK-293 cells grown on

different substrates: glass (control), P3HT and P3HT:PCBM.

4.1.2 Basic electrophysiology of HEK-293 cells

After having assessed the viability of HEK-293 cells grown on the P3HT-based films, we started

with the investigation of their basic electrophysiological properties, to see if they were comparable

to the data present in literature.

As briefly introduced in Chapter 1.1, the gold standard for measuring bioelectrical activity of cells

is the patch-clamp technique, especially in the whole-cell configuration.21

A thorough description of

this technique and the experimental setup is given in Appendix A. Briefly, the substrates with HEK-

293 cells are put under the microscope in a petri dish with a Krebs-Ringer-Henseleit buffer (KRH,

extracellular solution). A glass micropipette filled with a solution resembling the ionic composition

Page 68: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

63

of the cytoplasm of the cell (intracellular solution) is then pressed against the plasma membrane. By

forming the so-called gigaseal, the patch of membrane inside the micropipette tip is removed and

the pipette enters in electrical contact with the inside of the cell. Having an Ag/AgCl electrode in

the pipette and a counter-electrode of the same material in the extracellular solution, it is possible to

measure the potential difference across the plasma membrane.

After the formation of the patch, different measurements can be performed on the cell, depending

on which properties need to be investigated. Some of the electrical parameters of the cell are readily

available to the experimenter: the resting potential V0, the membrane capacitance Cm and the patch

series resistance Rs. The resting potential is the potential difference measured across the membrane

at electrochemical equilibrium (Veq), i.e. no net current flowing through the membrane, when the

cell is not perturbed by an external stimulus; the value of Veq is regulated by the Goldmann-

Hodgkin-Katz (GHK) equation, as already introduced in Chapter 2.3:

In the following, we indicate with V0 the cell resting potential, i.e. the membrane equilibrium

potential Veq at the base temperature (T0) before the cell has been excited in any way. The

membrane capacitance and the series resistance can be read from the patch amplifier after the

capacitive transients present upon electrical stimulation have been corrected during the formation of

the patch.

One of the basic characterization that can be done to understand if the cell is behaving correctly,

and thus the patch is good, is to measure the I-V characteristic of the membrane. We performed this

measurement by applying a sequence of increasing voltage steps (with pulse duration of 500 ms)

from -100 mV to +50 mV (in steps of 10 mV) to the cell and measuring the correspondent current

elicited, as depicted in Figure 4.2a for a typical HEK-293 cell. Between the voltage steps, a holding

potential of Vhold = -40 mV was kept, a value close to the ideal resting potential of HEK-293 cells.

The period between each voltage step was 10 s. The I-V curve for the membrane was then

reconstructed by plotting the mean value of the current in the last 100 ms of the pulse against the

respective value of applied potential. Given that the actual current flowing through the membrane is

proportional to the cell area, it is usually convenient to normalize the measured current for the

capacitance of the cell, which is proportional to the plasma membrane area. The intersection of

these curves with the x-axis is the potential at which no current flow through the membrane, and

thus the equilibrium potential of the cell. The slope of the curve in this point is related to the

membrane resistance Rm; more precisely, it is the membrane conductance Gm, or the specific

(4.1)

Page 69: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

64

conductance spG if the current has been normalized by the cell capacitance. Figure 4.2b shows

recorded I-V traces on different HEK-293 cells on glass substrates.

Figure 4.2 | (a) recorded current traces on an HEK-293 cell upon the application of a

stepping potential protocol from -100 mV to 50 mV (in 10 mV steps). (b) I-V curves

extracted from the current measurements for 7 different HEK-293 cells cultured on control

glass substrates.

The curves reported above show a generally rectifying behavior of the HEK-293 plasma membrane,

with small inward currents at negative potentials and bigger outward currents at positive potentials.

This behavior is consistent with the existing literature on endogenous currents in HEK-293 cells.203

However, a great variability in the actual values of the currents is observed in the different cases,

implying a substantial variability in the actual ion channel content of the membrane of the different

patched cells. Similar measurements have been performed also on cells grown on P3HT and

P3HT:PCBM devices; while the variability in the membrane properties like specific conductance

and resting potential is still high, no substantial difference can be observed for the different

substrates, as shown in Figure 4.3.

Figure 4.3 | Distributions of the values of membrane potential (a) and resistance (b) of

HEK-293 cells in resting conditions measured on different substrates: P3HT (n = 48),

P3HT:PCBM (n = 14) and glass (n = 12).

Page 70: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

65

4.2 Measurements on different substrates

We now turn to the study of the interaction mechanisms between the polymer-based interfaces and

the cell membrane upon illumination. A direct way to investigate the effect of photoexcitation of the

active material on the bioelectrical properties of HEK-293 cells grown on top of it is to measure the

corresponding variation in membrane potential with the patch-clamp. Typical whole-cell recordings

for a cell cultured on an ITO/P3HT:PCBM substrate are shown in Figure 4.4, upon illumination

with 20 ms light pulses (λ = 475 nm) at different intensities ranging from 7.7 to 47 mW/mm2.

Figure 4.4 | Typical traces of membrane potential measurements on an HEK-293 cell

grown on an ITO/P3HT:PCBM sample upon illumination with 20 ms pulses of light

(represented by the light blue box) at increasing intensities (7.7, 15, 35, 47 mW/mm2).

Two different components of the signal (A and B) may be distinguished based on their

characteristic timescales.

Figure 4.5 | Membrane potential measurements on four different HEK-293 cells cultured

on bare glass substrates for light pulses (I = 57 mW/mm2) of 20 ms (a) and 200 ms (b).

Two main components can be detected in the elicited signal during illumination: (i) a fast spiking at

the onset of the light pulse (A); (ii) a slower transient depolarization evolving during the

illumination (B). When the light is switched off, a complementary behavior is observed, with a

Page 71: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

66

reversed spike at light offset and a hyperpolarization afterwards. Both components increase with

higher light intensities. In order to univocally correlate the recorded signal to an effect mediated by

the active material, control measurements of cells on bare glass substrates were also taken (Figure

4.5). Accordingly, in this case, no variation in the cell membrane potential could be recorded.

The spiking signals (A) on ITO/P3HT:PCBM substrates are reminiscent of the capacitive charging

of devices described in Chapter 3.4. The slower component (B) may be a consequence of such

electrical coupling, but could also be related to other effects mediated by the photoactive interface.

In order to disentangle the origin of these two signals, the same measurements were conducted also

on ITO/P3HT and ITO/photoresist substrates. In pristine P3HT, charge generation is still occurring,

but with a much lower efficiency with respect to the blend; in the photoresist, instead, no charge is

generated upon illumination. The intensity of the components in membrane potential variations

related to the capacitive charging of the interface should thus be different in the three cases. Indeed,

the intensity of the A-component at light onset and offset is clearly dependent on the substrate

electrical properties, as shown in Figure 4.6. However, the B-component of the recordings is

qualitatively the same in all the devices.

Figure 4.6 | Membrane potential recordings of HEK-203 cells on different photoabsorbing

materials deposited on ITO-coated substrates: (a) P3HT:PCBM , (b) P3HT, (c)

Photoresist. Magenta boxes highlight the presence of fast spikes on devices with charge

generation capabilities. The different traces refer to increasing same light intensities (7.7,

15, 35, 47 mW/mm2).

To demonstrate that the fast spikes are related to the capacitive charging of the interface and not to

other effects due to charges generated in the active layer, measurements on glass/P3HT:PCBM and

glass/P3HT substrates were also performed. As it has been demonstrated in Chapter 3.4, the

presence of the ITO layer is crucial for closing the circuit with the solution. Its absence thus

prevents the instauration of a surface potential at the polymer/electrolyte interface. Accordingly, the

fast spikes in membrane potential recordings are suppressed in this case for both substrates (Figure

4.7). Nonetheless, the slower component is still present with similar intensities with respect to the

previous cases.

Page 72: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

67

Figure 4.7 | Membrane potential recordings of HEK-203 cells on different photoabsorbing

materials deposited on glass substrates: (a) P3HT:PCBM , (b) P3HT. The different traces

refer to increasing light intensities (7.7, 15, 35, 47 mW/mm2).

The origin of the B-component seems not to be correlated to an electrical effect mediated by the

active layer upon generation of charges. It has been shown in Chapter 3.5 that, apart from electrical

activity, photoexcitation of the active materials used here leads to a local heating in the bath close to

the surface. Indeed, for substrates with comparable absorption coefficients, similar temperature

transients have been measured when the light intensities used are the same, consistently with the

qualitative similarity observed for the membrane potential signals recorded on the different devices.

Thermal effects can thus be involved in the observed depolarization/hyperpolarization response.

4.3 Analysis of thermal effects

We focus now our attention on the characteristics of the slow (B) component recorded in the

previous section, in order to understand its origin and if it is actually due to thermal effects

mediated by the active layer. Since it has been shown that this component is qualitatively the same

in all the different substrates, we continue the analysis on cells grown on glass/P3HT devices, in

which only the B-component is present. Recordings of the membrane potential variations in four

different cells at the same light intensity (57 mW/mm2) are reported in Figure 4.8 for both 20 ms

and 200 ms pulses.

The recordings for the short pulses are similar to the traces already presented in the previous

section; however, it is evident that there is a great variability on the recorded signals from cell to

cell, both in terms of the intensity of the signals recorded and of their temporal dynamics. Looking

at the traces for the long pulses (200 ms), we can actually identify three different parts: (i) an initial

transient depolarization after the light is switched off; (ii) a gradual hyperpolarization occurring

during the illumination that takes over the initial depolarization signal; (iii) a further

hyperpolarization following the end of the pulse. In order to avoid confusion between the two

Page 73: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

68

hyperpolarization signals, we named them hypon and hypoff, respectively. Given that the hypoff signal

seems to be just the reverse processes of the initial depolarization, we focus on the two components

occurring during the light pulse.

Figure 4.8 | Membrane potential measurements on four different HEK-293 cells cultured

on glass/P3HT devices for light pulses (I = 57 mW/mm2) of 20 ms (a) and 200 ms (b).

4.3.1 Transient depolarization

First, we investigate the transient depolarization that occurs in the first milliseconds after the light

onset. As already observed, the traces in Figure 4.8a indicate that there is a great variability of this

depolarization component from cell to cell. This variability is confirmed by the boxplots in Figure

4.9a, where the peak depolarization values (ΔVpeak) for different light intensities are reported for n =

51 cells. However, by carefully observing the traces of Figure 4.8a, it can be seen that the variability

of the peak depolarization values is not casual, but is clearly correlated with the time the signal

takes to reach the peak (time-to-peak, tpeak): the faster the peak is reached, the lower is the value of

the depolarization. If the ΔVpeak values for all the cells (n = 51) are plotted against tpeak, a clear

correlation can be see between the two (Figure 4.9b).

Figure 4.9 | (a) Distribution of peak depolarization values obtained upon 20 ms

illumination at different light intensities for HEK-293 cells cultured on glass/P3HT (n =

Page 74: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

69

48) and control glass (CTRL, n = 12) substrates; for the control case, only the maximum

intensity (57 mW/mm2) is reported.. (b) Scatter plot of the peak depolarization values (I =

57 mW/mm2) versus the time in which the peak is reached upon the onset of the light

pulse for the data on glass/P3HT.

There should thus be a parameter of the cell that determines the dynamics of its temporal response.

Since the cell membrane can be schematized as the parallel of a capacitance (Cm, given by the

charges accumulated across the lipid bilayer) and a resistance (Rm, given by the ion channel

conductances), an obvious candidate is the system time constant given by the product of the two (τm

= RmCm). The membrane capacitance of each cell is known from the compensation of the capacitive

transients performed after the gigaseal is established, while the membrane resistance can be

extrapolated from the cell I-V curves (in particular, from the slope of the curve at the crossing with

the x-axis). Figure 4.10 shows the statistical distributions for the measured Cm and Rm, as well as for

their product τm; it is clear that the different cells have quite different values for the membrane time

constant, as already observed in Section 4.1, which could explain the great variability observed in

the measured dynamics.

Figure 4.10 | Distribution of the membrane electrical properties for the HEK-293 cells

cultured on glass/P3HT (n = 48): (a) membrane capacitance, (b) membrane resistance, (c)

membrane time constant, i.e. the product τm = CmRm.

Indeed, plotting the values of peak depolarization ΔVpeak and time-to-peak tpeak versus the time

constant of the corresponding cell, it is evident the influence of τm in determining the behavior of the

cell dynamics (Figure 4.11).

A behavior similar to the depolarization observed here was reported by Shapiro et al.58

in their

investigation of the physical mechanism behind Infrared Neural Stimulation (see Chapter 1.2). INS

is a cell stimulation technique based on illumination of biological tissues with an IR pulse that is

absorbed by water causing a local increase in temperature. In that paper, the authors demonstrated

that the depolarization measured in different experimental systems (oocytes, HEK-293, lipid

bilayers) upon photoexcitation was related to an increase in the membrane capacitance with

temperature. Interestingly, in their model the cell depolarization was indeed dependent on the

Page 75: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

70

membrane time constant. As the membrane capacitance increases, the potential difference across it

decreases (ΔV = Q/C), i.e. the cell depolarizes. However, this variation in potential from the resting

value Vr leads the membrane out of electrochemical equilibrium and a net current is established

across the membrane to restore the equilibrium potential. This re-equilibration mechanism explains

why the observed depolarization during the illumination is only transient; moreover, the rapidity

with which the equilibrium is reached depends on how much currents can flow through the

membrane, and thus on its specific conductance (spG), i.e. the conductance of the membrane per

unit area. Since the area of the membrane is proportional to the capacitance, spG is usually

expressed in terms of the ratio Gm/Cm, which is just the inverse of the membrane time constant.

Figure 4.11 | Scatter plots showing the dependence of the properties of the membrane

depolarization signal, the peak depolarization (a) and the time to peak (b), from the

membrane time constant τm.

4.3.2 Gradual hyperpolarization

After the transient depolarization, the membrane potential does not come back to its resting value

during the light pulse, but a hyperpolarization of the cell is instead observed (hypon). This signal

increases in time and, at least qualitatively, seems to follow the local temperature dynamics

measured in Chapter 3.5. The boxplots of Figure 4.12a show the maximum hypon signals recorded

just before the end of the light pulse at different light intensities for the n = 51 cells measured on the

glass/P3HT substrates. Again, also for this signal a great variability can be observed. However, this

signal does not show any significant correlation with the membrane electrical parameter (Figure

4.12b,c), thus indicating that its origin is different from that of the initial depolarization.

Page 76: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

71

Figure 4.12 | (a) Distribution of maximum hyperpolarization reached during a 200 ms

illumination at different light intensities in HEK-293 cells cultured on glass/P3HT (n =

48) and control glass (CTRL, n = 12) substrates; for the control case, only the maximum

intensity (57 mW/mm2) is reported. (b,c) Scatter plots showing the independence of the

hyperpolarization signal measured at the end of a 200 ms illumination (I = 57 mW/mm2)

from membrane time constant (b) and membrane capacitance (c).

A mechanism that could explain a hyperpolarization proportional to the local temperature is based

on the variation with heating of the equilibrium potential of the membrane given by the GHK

formula. The equilibrium potential Veq can be extracted from the I-V characteristic of the cell as the

potential at which no current flows through the membrane. To see if the hyperpolarization is indeed

related to a variation of the equilibrium potential, we measured its value at the end of a 200 ms

pulse (57 mW/mm2). To have more precise data, we did not measure the I-V curve in the entire -100

mV / +50 mV range, but in a smaller interval close to the resting condition. In particular, the

membrane potential of the cell was clamped at a holding potential (Vhold) close to its resting

potential; a stepping protocol was applied to the cell with 800 ms pulses of potential from -5 mV to

+ 5 mV (with respect to Vhold) in 1 mV steps. During the voltage pulses, a 200 ms pulse of light was

then delivered to the cell. The equilibrium potential at each instant can be extracted as the x-axis

intercept in the corresponding I-V plot. A typical example of such a measurement is presented in

Figure 4.13.

The recorded I-V characteristics show that indeed a variation in the equilibrium potential towards

more negative values occurs upon photostimulation. Plotting this value against the corresponding

hyperpolarization measured in current-clamp recordings (n = 17 cells) clearly show a good match

between the two (Figure 4.14a), indicating that indeed the hyperpolarization is due to a variation in

the equilibrium potential in time.

Page 77: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

72

Figure 4.13 | (a) Current response of an HEK-293 cell to a stepping protocol with 1 mV

steps around the resting membrane potential (from -5 mV to +5 mV); the light blue box

represent the 200 ms pulse of light (I = 57 mW/mm2), while the grey and magenta boxes

identify the regions from which the current data for the dark and light conditions were

respectively extracted. (b) Comparison between the I-V curves extracted from the

measurements in panel (a) in dark and light conditions.

Figure 4.14 | (a) Correlation between the cell hyperpolarisation measured in current-

clamp experiments at the end of a 200 ms pulse of light and the variation in reversal

potential as measured from the protocol of Figure 4.13; the grey dashed line represents the

quadrant bisector. (b) Correlation between the variation in cell membrane resistance in the

dark and at the end of the light stimulus; the grey dashed line represent the quadrant

bisector, while the solid blue line represent the line best fitting the data, with a slope of

0.804 ± 0.017. Points in e,f represent data from individual cells.

This observation is consistent with the GHK equation, where it can be seen that Vr is directly

proportional to the temperature. If V0 the cell resting potential at temperature T0, the equilibrium

potential Veq upon heating can be expressed as:

(4.2)

Page 78: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

73

Since the resting potential V0 is negative, an increase in temperature leads to a negative variation in

the membrane potential (ΔV), i.e. a hyperpolarization, as it is observed. Moreover, a bigger

hyperpolarization should be obtained for cell starting from a more negative resting potential.

However, this behavior is not observed experimentally, as it can be seen by the scatter plot in Figure

4.15, where the maximum hyperpolarization values (at 57 mW/mm2) are plotted against the resting

membrane potential for the corresponding cell.

Figure 4.15 | Scatter plot of the hyperpolarization measured in a HEK-293 cell at the end

of a 200 ms light pulse (I = 57 mW/mm2) with respect to the resting potential of the same

cell (n = 48).

This result indicates that in the simple temperature dependence of Equation (4.2) something is

missing. In particular, has it has been discussed in Chapter 2.2, the permeability of ion channels is a

thermally activated process164

and thus the parameters P[x] in the GHK equation are actually

temperature dependent. The actual relationship between the equilibrium potential and temperature is

thus more complex and is dependent on the thermal properties of the single channels and their

density in the cell membrane. In any case, since ion channels permeability increases with heating,

the membrane resistance should decrease at higher temperatures. The membrane resistance can be

estimated from the slope of the I-V curves reported in Figure 4.13b; indeed, upon illumination a

steeper characteristic is obtained, which is consistent with a lower membrane resistance. In

particular, a decrease of about 20 % of the initial value can be estimated by the statistical analysis of

the data for all the n = 17 cells reported in Figure 4.14b.

4.3.3 Time evolution of membrane properties

To sum up, from the analysis of the membrane potential traces we hypothesize that the two

components of the signal during the illumination are given by two different mechanisms, both

triggered by the local heating. In particular, we associate the initial depolarization to an increase in

membrane capacitance that leads to a transient variation in the membrane potential before this effect

Page 79: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

74

is balanced by the current flowing through the membrane and the cell goes back to its equilibrium

potential. The subsequent hyperpolarization is instead due to an increase in the conductances of the

membrane ion channels that leads to a variation in the cell equilibrium potential.

In order to prove in a direct way the effect of temperature on the membrane electrical properties, i.e.

resistance and capacitance, we tried to measure their temporal dynamics during illumination. The

measurement was performed by analyzing the cell response to an oscillating voltage input21

during a

pulse of light (200 ms, 57 mW/mm2). The cell was patched in voltage-clamp mode and was hold at

its resting potential; on top of this potential, a double sinusoidal perturbation was applied:

In the following measurements, we used A = 10 mV and f = 195.3125 Hz (which corresponds to a

period of T = 5.12 ms). In complex notation, with ω = 2πf, this stimulus can be expressed as:

As explained in Section 2.3, in a simplified version a patch-clamp experiment can be modeled with

a with the parallel of a capacitance (Cm) and a resistance (Rm), making up the membrane, and a

series resistance (Rs), that takes into account the presence of the patch pipette (Figure 2.4b). In the

following analysis the voltage sources VX and Vσ are considered as DC signals. While in reality

these two parameters can vary in time due to the temperature gradient, their variation is slower

compared to the frequency here considered. The result of this approximation is that they thus do not

give any contribution to the final result in the following AC analysis. The complex impedance of

this system can be calculated as:

Upon the application of the input voltage of Equation (4.3), the theoretical output current is given

by the following relationships:

Figure 4.16 show an example of the output current simulated with this model for a cell with the

following electrical parameters: Rs = 15.5 MΩ, Rm = 715 MΩ, Cm = 13.5 pF. For comparison, an

actual current trace measured from a cell with the same electrical properties is also reported.

(4.3)

(4.4)

(4.5)

(4.6)

Page 80: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

75

Figure 4.16 | (a) Simulated current response (in red) of a cell with Cm = 13.5 pF, Rm = 715

MΩ, Rs = 15.5 MΩ upon the stimulation with the oscillating voltage protocol depicted in

blue. (b) Current traces actually recorded on a HEK-293 cell with the same electrical

parameter as panel (a).

In order to extrapolate the values of the electrical parameters in time from the measured currents,

the traces were divided in single periods of 5.12 ms. From each period, the instantaneous values of

Cm, Rm and Rs were estimated by fitting the measured current with the one expected from the

theoretical impedance of Equation (4.5). Repeating this process for each period in the measurement,

we reconstructed the temporal dynamics of the three parameters. The measurements were

performed on n = 39 cells; while different cells have different initial values of capacitance and

resistance (see Figure 4.10), the relative variations of these parameters were seen to be quite

reproducible as it can be seen from the statistical distributions of the signals reported in Figure

4.17.

Figure 4.17 | Variation in time of membrane electrical properties upon illumination with a

200 ms pulse (I = 57 mW/mm2): (a) membrane capacitance, (b) membrane resistance, (c)

series resistance.

The results here presented show that indeed all three parameters vary during the illumination pulse.

Interestingly, the traces have similar dynamics to the temperature variations measured in Chapter

3.5, supporting the influence of temperature on the membrane electrical parameters.

Page 81: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

76

The series resistance Rs is seen to vary of about 10 % during the 200 ms pulse. This resistance does

not describe a parameter of the cell itself, but it’s a necessary element to perform the measurement.

Since the series resistance is mainly determined by the access resistance of the pipette, it is not

surprising that it varies with temperature. Indeed, the temperature dependence of the resistance of a

micropipette was used in Chapter 3.5 to estimate the local heating at the interface. Also in that case,

a variation of about 10 % of the pipette resistance was measured upon illumination with a 200 ms

pulse at 57 mW/mm2.

The membrane capacitance Cm has a maximum variation of about 2 % at the end of the 200 ms

pulse. This increase in capacitance is consistent with the proposed explanation for the transient

depolarization signal. However, since this phenomenon is transient, only the increase of temperature

occurring in the first milliseconds is actually important in determining the depolarization value

reached. At the same time, however, the current flowing through the membrane is balancing the

depolarizing effect. If we consider a cell with a very small specific conductance spG, the effects of

this second phenomenon can be considered negligible in the first 20 ms; in this case we can

approximate the charge on the membrane as constant. The depolarization reached is thus given by

the following relationship:

In which V0 – Vσ is the potential difference across the membrane capacitance (see equivalent circuit

in Figure 2.4). Considering typical values as V0 = -30 mV and Vσ = 120 mV58

and a relative

variation in the capacitance of about ΔCm/Cm = 1 % in the first 20 ms, a depolarization of ΔV ≈ 1.5

mV is expected. This value is consistent with the maximum depolarization reached in the

experiments from Figure 4.9b, since a cell with a small specific conductance corresponds to the data

for high time constants τ. Moreover, comparing the temperature variations recorded in Chapter 3.5

with the data here, we can estimate a variation in membrane capacitance of about 0.3 % for each

degree of temperature increase. This value is consistent with the experiments reported for HEK-293

cells by Shapiro et al.58

upon stimulation with IR light.

The membrane resistance Rm traces show a decrease at the end of the 200 ms pulse of about 18 %.

This value is consistent with the one extrapolated in the previous section from the I-V curves of the

cells. The variation in membrane resistance has been attributed to the increase in ion channel

conductances with local heating. From the measured decrease of about 20 % upon a local heating of

7 °C, the temperature coefficient Q10 can be estimated as:

(4.7)

Page 82: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

77

Actually, it is reported in literature that different ion channels have different temperature

coefficients.164

It is thus difficult to correlate the value calculated here with literature data, not

knowing the exact types and distribution of ion channel in the membrane. This is especially true if

we consider the great variability of membrane resistances in the measured HEK-293 cells, which

means that each cell has a different composition in terms of ion channels expressed. However, a

rough comparison with literature date shows that typical values for the Q10 coefficient in different

families of ion channels usually range from 1.2 to 1.6,164

in accordance with the value calculated

here.

4.3.4 Numerical modeling

The measured evolution in time of the membrane properties (i.e. resistance and capacitance) seems

to follow quite closely the dynamics of the local temperature reported in Chapter 3.5. Here we try to

extrapolate a numerical relationship between the membrane electrical parameters and the

temperature, in order to directly link the membrane potential variations measured in HEK-293 cells

with the local heating. Applying Kirchhoff’s law to the equivalent circuit introduced in Chapter 2.3

(see Figure 2.4), the following differential equation is obtained:

Here, Vm is the membrane potential to be determined (which is equal to the potential inside the cell

if the outside is fixed for convention at zero), while Ip is the current flowing through the pipette.

This term is in principle dependent on the series resistance Rs, but since the measurements of the

membrane potential are taken in current-clamp mode, Ip is in this case it is equal to zero. The

equation can thus be readily solved numerically once the time evolutions of the membrane

parameters (Cm, Rm, Vr and Vσ) are known.

Based on the collected measurements and on literature models, for the membrane capacitance Cm,

we hypothesize a linear relationship of this parameter with the local temperature T, in the form of:

where T0 is the base temperature and αC the proportionality constant. As for the membrane

resistance Rm, it can be readily obtained from the Q10 formulation as:

(4.8)

(4.9)

(4.10)

Page 83: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

78

Starting from the temperature transients measured in Chapter 3.5 and setting the parameters Q10 =

1.32 and αc = 0.0032 K-1

, these two formulas can be used to closely reproduce the traces

experimentally found in the previous section, as reported by the solid lines in Figure 4.18.

Figure 4.18 | Simulated traces (blue solid lines) for the variation upon illumination of the

membrane resistance (a) and membrane capacitance (b), compared to the experimental

data from Figure 4.17 (grey shaded regions).

The modeling of the equilibrium potential Veq is a more complex problem. Theoretically, it is given

by the GHK equation. As reported above, however, this relationship has a non-trivial dependence

on temperature that involves the knowledge of the ion channels distribution and their specific

thermal responses. Since we cannot reach such level of insight with our measurements, we need to

use an empirical relationship to simplify the GHK equation:

This formula is a generalization of Equation (4.2), with the parameter αV used to model the complex

temperature dependence of the logarithm part of the GHK equation. The case αV = 1 corresponds to

an “ideal” situation in which all the ion channels in the cell membrane have the same Q10 coefficient

and thus their relative contributions to the total membrane current remains constant; in this case the

logarithm part does not vary during the illumination and the only temperature dependence is in the

prefactor, leading to the formulation of Equation (4.2). Values of αV > 1 are indicative of a tendency

of the cell to hyperpolarize more than the ideal case, i.e., on average, the increase in temperature

favors more the conduction of channels mediating outward currents. The opposite holds true for αV

< 1, and in the case of negative values the cell actually tends to depolarize with increasing

temperature. The last parameter of the model, Vσ, is necessary to take into account the charge

asymmetry on the two sides of the plasma membrane, which is due to the differences in ion

(4.11)

(4.12)

Page 84: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

79

concentrations but also to different charging of the inner and outer surfaces of the membrane, as

explained in Chapter 2.3. As a first approximation, we consider this parameter constant in the

temperature range here investigated.

Based on the relationships just introduced and the measured time traces of local temperature

(Section 3.5), Equation (4.9) can be solved to reproduce the experimental traces of the membrane

potential upon illumination. In particular, with this model we fitted the signals from the n = 51 cells

measured on P3HT/glass, both for 20 ms and 200 ms pulses. For each cell, we considered αc and αV

as free parameters in the fitting, while the values of Cm,0, Rm,0 and V0 were available from the

experimental data. For the remaining parameters, we chose common values for the all the cells

which gave the best fit, paying attention that they were compatible with literature data; in particular,

we used Q10 = 1.32164

and Vσ = 160 mV.58

The distributions of the obtained values for αc and αV are

reported in Figure 4.19.

Figure 4.19 | Distribution of the values for the αC (a) and αV (b) parameters obtained by

fitting the numerical model of the cell behavior to the experimental curves.

The results for αc show a quite narrow Gaussian distribution for this parameter, with a mean value

of 0.0031 K-1

and a standard deviation of 0.0004 K-1

. This value is consistent with the observed 0.3

% variation in capacitance for each degree of temperature and the narrow distribution implies that

the capacitance variation upon heating is basically the same in all the cells measured; the variability

of the actual depolarization peaks reached upon illumination is thus given by the differences in the

membrane time constants, as already explained in Section 4.3. The values for αV are instead much

more scattered, from close to 0 up to about 2; this variability is reflected in a great difference in

hyperpolarization dynamics between cells, consistently with what reported in Section 4.3.

The actual dynamics of the simulated membrane potential dynamics for a cell are compared to the

experimental data in Figure 4.20, both for 20 ms and 200 ms light pulses. The dashed lines in the

graphs represent the contributions to the final signal coming from the variation in capacitance (blue

Page 85: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

80

lines) and from the variation in the membrane equilibrium potential (green lines). As it was

proposed at the beginning, the first component shows that the transient depolarization upon

illumination and the corresponding hyperpolarization after the end of the pulse are given by the

variation in membrane capacitance with temperature. The second component instead confirms that

the gradual hyperpolarization occurring during the light pulse is related to the shift in the

equilibrium potential.

Figure 4.20 | Results of the numerical fitting (pink solid line) for an HEK-293 cell upon

illumination with a 20 ms (a) and 200 ms (b) light pulse (I = 57 mW/mm2), compared to

the experimental traces. In the two panels are reported also the single contribution to the

total signals from the variations in membrane capacitance (blue dashed lines) and

equilibrium potential (green dashed lines).

4.4 Considerations on capacitive charging

In the previous section we have characterized in depth the effects of local heating on the electrical

properties of the cell membrane in HEK-293 cells and the relative variations in membrane potential.

However, the measurements in Figure 4.6 showed that also an effect related to charge generation in

the active material can be present in samples deposited on ITO-coated substrates. In particular, we

observed very fast spikes in the recordings of membrane potential upon the onset and offset of the

light pulse. These signals were quite similar to the traces of the surface electrical potential due to

the capacitive charging of the polymer/electrolyte interface investigated in Chapter 3.4. For direct

comparison, we measured the surface potentials on the different samples used for HEK-293

experiments at the same light intensity used for cell stimulation (57 mW/mm2).

Page 86: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

81

Figure 4.21 | Surface potentials measured on the devices used for cell stimulation with a

light intensity of 57 mW/mm2: (a) P3HT, (b) P3HT:PCBM, (c) Photoresist. Both

architectures with (blue) and without (grey) the ITO electrode were tested.

As already seen in Chapter 3.4, devices without the ITO layer did not give any signals. In the

presence of the conductive contact, instead, surface potentials were visible both on P3HT and

P3HT:PCBM, but not on Photoresist, since it does not support the generation of charges. As

expected, the signal on the blend was significantly higher than the one measured on the pristine

polymer. These signals are perfectly comparable to those measured in the membrane potential of

HEK-293 cells upon photostimulation; however, it is not clear if such signals are an actual variation

in the membrane potential, if they are just an artifact due to the coupling of the surface signal to the

recording electrode, or a combination of both.

Let’s call Vcell the potential inside the cell and Vsurf the potential at the polymer/electrolyte interface

and Vm the transmembrane potential. If we consider the cell as an isolated system in proximity of

the device interface, as depicted in the electrical scheme of Figure 4.22, the potential on the outer

side of the membrane is always equal to Vsurf (we are considering the resistance of the electrolytic

solution negligible) and thus:

A variation in the surface potential does not modify the actual membrane potential. Instead, the

potential inside the cell changes in the same way as the outside potential changes, and thus the

potential difference across the membrane remains constant. This happens because, since every

portion of the membrane feels the same outer potential, no current can flow through the membrane

resistances (i.e. the ion channels) and thus the transmembrane potential cannot change. However, in

a patch-clamp experiment in current-clamp configuration what is measured is not the actual

transmembrane potential Vm (which is the biophysically relevant parameter), but the cell potential

Vcell (with respect to the counter-electrode one Vref, that can be considered equal to zero as a

convention). In this case, a variation in the measured potential is to be considered an artifact due to

the superposition of the Vsurf signal to the actual Vm in the recordings.

(4.13)

Page 87: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

82

Figure 4.22 | Simplistic modeling of the interface between a cell and the hybrid polymer-

based interface.

This schematization, however, is not complete. Capacitive charging as a tool for stimulating cells

has been largely investigate with devices based on inorganic technology. In particular, Fromherz

and coworkers91

have studied in depth the interfacing of cells with silicon-based devices (transistors

and capacitors). In these systems it is important to consider that the adhesion to the substrate brings

the basal membrane of the cell at a very small distance from the device surface, which is usually

less than 100 nm. Conduction in such a confined space, called cleft,32,33

is to some extent reduced.

To model this effect we can, at least as a first approximation, divide conceptually the cell membrane

in two different parts: the basal membrane and the lateral membrane. The basal membrane feels, on

its outer side, the potential in the thin cleft (Vcleft); the lateral membrane, i.e. the remaining part of

the plasma membrane in contact with the extracellular medium, feels on the outside the surface

potential Vsurf. The cleft region and the extracellular space are connected via a cleft resistance Rcleft,

as schematized in Figure 4.23. Upon charging of the device interface, the cleft and surface potential

can be quite different if Rcleft is high enough and thus the two portion of the membrane can have

different potentials, in principle different from the resting one; in this case, a stimulation of the cell

can be possible. If Rcleft is too small, instead, the model becomes equivalent to the isolated cell one

and no actual stimulation of the cell can be obtained with the capacitive charging of the surface.

This model for the plasma membrane is a first approximation of the real coupling of the cell with

the device surface. More precise models have been developed, taking into account for example the

actual space distribution of the potential in the cleft region or the different conductivity of the cleft

for the different ion species. In any case, the two-compartment model is sufficient to give at least a

qualitative idea of the effects of capacitive stimulation on the cell membrane potential.

Understanding if the fast spiking signals we measured in HEK-293 cells upon photostimulation of

the semiconducting polymer are just an artifact or are actually correlated with a variation in the

Page 88: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

83

membrane potential of the cell is not straightforward. Fromherz and coworkers have studied HEK-

293 cells transfected with genes for the expression of voltage-activated channels206

(both for K+ and

Na+) and have demonstrated that capacitive stimulation (with inorganic technologies) is able to

trigger the opening of these channels. However, in our work we only had at our disposal non-

transfected cells and observing the effects of capacitive charging of the interface in these systems is

not so direct. We thus performed some simulations to understand if, with reasonable values for the

system parameters in our case, capacitive stimulation of the cell membrane can be expected or not.

Starting from the electrical model of the photoactive interface proposed in Chapter 3.4, we used the

cell and cleft modeling presented in literature to obtain the equivalent circuit of Figure 4.23. In

particular, we made the following hypothesis:

The cell membrane is modeled as a two-component system like in the Fromherz model, with the

basal and lateral sections. The cell basal area has been taken from typical values for HEK-293

cells (Acleft = 300 μm2), while the total area has been estimated more or less three times the basal

area (Acell = 1000 μm2).

The cell membrane capacitance (Cm = 15 pF) and resistance (Rm = 500 MΩ) have been taken

from the electrophysiology data (Section 4.1) and divided between the basal (Cm,B e Rm,B) and

lateral (Cm,L e Rm,L) sections proportionally to their area.

The cleft is modeled as a compartment separated from the extracellular space by a cleft

resistance Rc. A typical value for this parameter taken from literature32,33

is 1 MΩ.

The cytoplasm is considered sufficiently conductive that the internal potential Vcell is the same

in every point of the cell. This hypothesis is reasonable for HEK-293 cells, given their compact

morphology without significant processes. However in other cells, like neurons that posses thin

dendrites protruding from the cell body, a more sophisticated model may be necessary to take

into account spatial variations of the internal potential.

The specific capacitance of the polymer/electrolyte interface has been considered the same for

both the cleft and the extracellular space regions.

Table 4.1 summarizes the base values used for the different parameters of the electric circuit. The

values for the description of the active interface have been selected in order to give a surface

potential similar in intensity and dynamics to that measured in P3HT:PCBM devices in Chapter 3.4

(I = 784 μW/mm2). The effect of photostimulation on the transmembrane potentials has been

calculated for both the basal (ΔVB = Vcell – Vcleft) and the lateral membrane (ΔVL = Vcell – Vsurf) upon

a light pulse of 20 ms; their dynamics at the onset of the light pulse are reported in Figure 4.24,

compared to the trace for the elicited surface potential (Vsurf) and the internal cell potential (Vcell, the

signal actually measured in a patch-clamp experiment). The simulations were performed with the

OrCAD® PSpice circuit simulator suite.

Page 89: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

84

Figure 4.23 | Electrical schematization of the two-compartment model applied to the case

of cell stimulation by the hybrid interface characterized in Chapter 3.4. The green box

represents the area of the device illuminated during stimulation.

The simulation shows that indeed there is an actual variation in the membrane potential. In

particular, the basal and lateral sections are subjected to different transients: the basal membrane is

hyperpolarized, while the lateral section is depolarized. This double behavior is common for

capacitive stimulation and is the main reason why with this kind of architecture it is difficult to

obtain suppression of activity in excitable cells.207

The two sections of the membrane are always

subject to signals with opposite signs; in the section that is depolarized the voltage-gated ion

channels tend to open, depolarizing the interior of the cell even more and overcoming the

hyperpolarization of the other section of the membrane.

Figure 4.24 | Simulation of the transmembrane potentials elicited by capacitive charging

of the interface in the later (green) and basal (orange) compartments of the cell, compared

to the surface potential (blue) and the potential inside the cell (magenta) measured with

respect to the reference electrode. Given the fast dynamics, the graph reports for clarity

only the first milliseconds after the light onset.

Page 90: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

85

Table 4.1 | Base values of the parameters used to simulate the electrical circuit of Figure

4.23.

In our case, it is interesting to observe that both the hyperpolarization and the depolarization

obtained are in absolute values higher than the surface potential elicited by the photoactive

interface. The drawback is that the elicited signals are very fast and, in the case of excitable cells

they could be too short to efficiently trigger the opening of voltage-gated channels. As already

discussed in Chapter 3.4, this issue is related to the value of the surface capacitances of the device

and could be addressed by proper engineering of the interface.

To have a better understanding of how the relevant parameters of the cell can influence the ability to

obtain an efficient stimulation, we performed different simulations by varying their values within

reasonable ranges. The relative variations in the peak signals and stimulus duration (estimated as the

pulse area divided by its peak value) are reported in Figure 4.25.

The cleft resistance Rcleft is the main parameters that determine the coupling of the surface

potential to actual variations in the cell membrane potentials. The membrane signals are linearly

proportional to this parameter, except for a slight saturation at very high values. It is interesting

to notice that the membrane potentials obtained can be much higher than the actual surface

potential elicited (in this case it is about ≈ 500 μV). The duration of the stimulus instead is

initially quite independent of Rcleft, increasing only at high resistances.

By changing the cleft area Acleft keeping the total cell area constant it is possible to simulate the

effect of the cell adhesion to the substrate. Increasing Acleft (i.e. the cell is well spread on the

interface) has a positive effect on both the lateral and basal signals, but while the increase is

more than linear on ΔVL, it saturates for ΔVB. As for the stimulus duration, it remains virtually

unchanged in all cases.

The cell area Acell was modified keeping constant the ratio with the basal membrane (Acleft =

Acell/3.33) in order to simulate cells with different dimensions. Also the membrane resistance

and capacitance have been modified proportionally. Both the lateral and basal signal increase

Description Value Description Value

Is Diode saturation current 17 nA Acleft Cleft area 300 μm2

N Diode ideality coefficient 2 Acell Cell total surface 1000 μm2

CJ Diode junct. capacitance 690 pF Alight Illuminated area 0.23 mm2

Iph Photogenerated current 13.8 μA cint Interf. specific capacitance 2 μF/cm2

Rbulk Bulk resistance 60 kΩ Clight Illuminat. area capacitance cint*(Alight - Acleft)

Rel Electrolyte resistance 100 Ω Ccleft Cleft interf. capacitance cint*Acleft

Cm Cell membrane capacitance 15 pF Rm Cell membrane resistance 500 MΩ

Rcleft Cleft resistance 1 MΩ

Page 91: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

86

with increasing cell dimension in a linear fashion, while the duration of the pulses are only

slightly lengthened.

Changing the membrane resistance Rm or the membrane capacitance Cm has no sizable

effects on the stimulation dynamics of the cell, both in terms of peak signals and pulse

durations.

This analysis shows that, to have efficient cell stimulation, the main factor to be considered is to

have a good adhesion of the cell to the substrate, with a highly resistive access to the cleft. Bigger

cells are excited better because of an increased cleft area, while the specific capacitance and

conductance of the membrane do not have significant effects.

Figure 4.25 | Dependence of the elicited variations in membrane potential in terms of

peak signal (orange: basal, green: lateral) and decay time (blue: basal, red: lateral) from

the main electrical parameters describing the cell and its adhesion to the substrate. For the

basal and lateral potential, the absolute value of the variation is reported for clarity; in any

case, it should be reminded that, for the electrical parameters used, ΔVB is always negative

(i.e. hyperpolarizing) and ΔVL is always positive (i.e. depolarizing). (a) cleft resistance;

(b) percentage of the cleft area with respect to the total area; (c) cell total surface, scaling

the cleft area accordingly; (d) membrane resistance; (e) membrane specific capacitance,

i.e. the ratio between the area and the total cell surface.

Page 92: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

87

Chapter 5 – Discussion and perspectives

In Chapter 3 and Chapter 4 we investigated the functioning of the hybrid polymer/electrolyte

interfaces in terms of their electrical and thermal properties, as well as the effects that these

phenomena have on the membrane potential of cells grown on their surface.

In this chapter we discuss how the results obtained compare with existing technologies for cellular

stimulation and in particular how they relate to our previous work of photoexcitation on neurons

and retinas. We then give some perspectives by identifying the main parameters of these systems

that need to be optimized and propose some possible future developments.

5.1 Discussion

From the investigation of hybrid polymer/electrolyte interfaces carried out in – Hybrid interfaces

characterization we have found that two main phenomena occur upon photoexcitation of the device

with short pulses of light:

In devices with an ITO electrode, it is possible to elicit a capacitive photocurrent upon

illumination. This signal is driven by the charge separation at the ITO/polymer junction and last

until the capacitance of the polymer/electrolyte interface has been charged. In devices based on

P3HT:PCBM active materials, in which the charge separation process is more efficient, peak

currents of several hundreds of μA/cm2 can be reached at the maximum high intensities used (jpc

≈ 900 μA/cm2 at Ilight ≈ 200 μW/mm

2), in a configuration with all the device active area (A ≈ 1

cm2) illuminated and the ITO contact short-circuited to the reference electrode. This current

however decays very rapidly to zero, with time constants shorter than 1 ms. We have also

shown that, even if in the device configuration used for cell stimulation the ITO electrode is not

contacted, the presence of a parasitic coupling with the electrolytic solution effectively close the

circuit and allow the charging of the polymer/electrolyte interface.

All devices, irrespectively of the presence of the ITO electrode, dissipate the great majority of

the energy of the photons absorbed via non-radiative decay and ultimately by heat transfer to

the surrounding environment. In particular, upon localized illumination of the active devices

(Alight = 0.23 mm2, intensities up to Ilight ≈ 57 mW/mm

2), temperature increases at the

Page 93: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

88

polymer/electrolyte interfaces on the order of few degrees Celsius have been measured (ΔT1 ≈ 3

°C and ΔT2 ≈ 7 °C with pulse duration of respectively t1 = 20 ms and t2 = 200 ms).

5.1.1 Capacitive stimulation

Capacitive stimulation of biological preparations has been widely investigated in the past with

architectures mainly based on inorganic semiconductors, in particular silicon chips or insulated

metal electrodes. In the great majority of reports, electrical stimulation, i.e. a voltage pulse, was

employed to elicit a capacitive current at the semiconductor/oxide/electrolyte interface. However,

some reports have also used a mixed electro-optical architecture in which the light was used to

trigger the charging of the device by increasing the semiconductor conductivity.

The capacitive current density elicited by a voltage pulse (Vs) applied to the device is basically

dependent on the temporal derivative of the voltage profile, following the relationship:207

where cint is the specific capacitance of the semiconductor/oxide/electrolyte interface. To improve

the efficiency of the device, i.e. the possibility of obtaining reliable stimulation at a lower applied

potential, a high value of the capacitance is thus desirable.

The standard silicon-based device used for capacitive stimulation is formed by a silicon chip

covered by a layer of silicon oxide that acts as the dielectric with thicknesses ranging from tens to

hundreds of nanometers.89,208

Typical values of specific interface capacitance are mainly determined

by the geometric capacitance of the silicon dioxide layer and are on the order of few hundreds of

nF/cm2 (the geometric capacitance of a 10 nm layer of SiO2 is CSiO2 = 0.34 μF/cm

2). In order to

improve this value, dielectrics with higher dielectric constant and with lower thicknesses have been

developed. For example, with hafnium oxide (HfO2),209

devices with a interface capacitance of

CHfO2 ≈ 1-3 μF/cm2 have been developed,

207,210 thanks to its very high dielectric constant (nHfO2 = 25,

compared to silicon oxide nSiO2 = 3.9).

Typical values of capacitive currents used for stimulation with inorganic devices are in the order of

hundreds of μA/cm2 to few mA/cm

2.210,211

Since the current depends on the derivative of the

potential pulse Vs, ramp of voltages are usually employed, in order to have constant current values.

Typical pulse durations in these stimulation protocols are in the order from few hundreds of

microseconds to few milliseconds, depending on the specific experiment. In order to obtain these

capacitive currents, the potential variations applied to the silicon capacitor are generally on the

order of few volts.

(5.1)

Page 94: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

89

In the case of our hybrid polymer based devices, the semiconductor is in direct contact with the

electrolyte, without a dielectric layer in between. In this case the actual identification of a dielectric

with a precise geometrical thickness and dielectric constant is not straightforward. In particular, the

capacitance derives from a combination of different factors, like the average extension of the

diffusion layer of ions in the electrolyte and the dimension of the region where charges are

accumulated in the semiconductor.170

These elements are actually present also in inorganic devices,

but usually give only secondary contributions with respect to the geometrical properties of the

dielectric layer. Another factor that can significantly influence the interface capacitance in polymer-

based devices is the conformation of the polymer at the surface and in particular the nature and

orientation of the side chains.212

Moreover, also the doping state of the material and the eventual

penetration of water molecules into it can effectively vary this parameter.213

From impedance

spectroscopy measurements, we have extracted for our P3HT-based devices capacitance values of

about cint ≈ 2 μF/cm2.197

However, examples reported in literature for devices based on thiophene

derivatives can reach specific interfacial capacitances on the order of cint = 10 μF/cm2.212

It has to be

noted that these values of capacitance have usually a significant dependence on frequency and bias

point of the device. However, in the range of voltages we are interested in, i.e. few hundreds of

millivolts (from photovoltages measurements), the device capacitance is basically constant.170,212

As

for the frequency, the fastest signals we measure have dynamics on the order of hundreds of

microseconds, which corresponds to frequencies of few kilohertz, comparable to the cut-off

frequency measured experimentally for the device capacitance.

The currents we are able to elicit with our hybrid interfaces upon photostimulation reach values

close to 1 mA/cm2 in the case of P3HT:PCBM based devices. While this value is comparable to

typical currents obtained for electrical stimulation, the elicited currents quickly drop to zero with

time constants shorter than 1 ms for the highest illumination intensities, i.e. the currents has a high

value only for few hundreds of microseconds. This spiking behavior can be related to the step

profile of the light pulse, which is reflected in a basically rectangular photovoltage profile at high

light intensities. Moreover, there is a limitation in the photopotentials achievable at the ITO

electrode. Their value is given by the polarization of the ITO/polymer Schottky junction due to the

current generated upon light absorption.192–195

Since this junction behaves like a diode, the increase

in the ITO potential is only logarithmic with the photogenerated current, and thus with the light

intensity.

In this work we focused the study on non-excitable cells in order to understand how the basic

properties of the cell membrane are influenced by the photoexcitation mediated by the hybrid

polymeric interface. As a further step, it would be now interesting to transfer the information

collected to excitable cells, where an active response can be elicited by a variation in the membrane

Page 95: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

90

potential due to the capacitive charging of the interface. It is indeed critical to understand to which

extent the capacitive charging of the interface determines the action potential generation by

triggering, for example, the opening of voltage gated channels, and to identify excitations thresholds

and critical conditions.

5.1.2 Thermal stimulation

The measurements of local heating at the polymer/electrolyte interface performed Chapter 3.5

demonstrate that basically all the energy of the absorbed light is released as vibrational energy

following non-radiative decay pathways of the photoexcited species in the active material. In

Chapter 4.3 we have then demonstrated that such increase in temperature is able to modulate the

membrane potential with different mechanisms depending on the considered timescale. In

particular, we have observed an initial transient depolarization related to a change in the membrane

capacitance with temperature. This depolarization reaches values of about 1.5 mV for the highest

power intensity used (57 mW/mm2), but it has been shown to be quite dependent on the membrane

electrical properties and the mechanism becomes less effective for more conductive cells. On longer

timescales, a sustained hyperpolarization is instead observed during illumination, with values again

of the order of 1-2 mV. This second process has been attributed to a variation in the membrane

equilibrium potential due to the temperature dependence of ion channels conductances.

A photothermal stimulation process in biological systems has been proposed to explain the

excitation of neurons observed upon illumination with pulses of infrared light (INS, infrared neural

stimulation).56,60

In 2012, Shapiro et al.58

demonstrated that a local increase in temperature is related

to a variation in the membrane capacitance and that this variation can indeed lead to a transient

membrane depolarization. The data we measured on HEK-293 cells with visible light illumination

of the photoactive devices are in accordance with the INS experiments. In particular, we found in

our system an increase in membrane capacitance of about 0.3 % for each degree Celsius of heating,

quite consistent with the measurements of Shapiro et al.58

on the same cell line, who reported a ΔC

≈ 1.7 % for a ΔT ≈ 10 °C. The main difference with respect INS is the phototransducing element: in

INS, the light is absorbed by water, while in our case the interaction is mediated by the active

polymer layer. INS is a very flexible tool for in-vivo experimentation,65,214–216

since it does not

require the use of an exogenous absorber. The main absorption peak of water usually used for INS

is the one at about 1.9 μm, where water has an absorption coefficient217

on the order of α ≈ 100 cm-

1. The light is usually delivered to the preparation via an optical fiber micromanipulated in close

proximity of the cells. Our polymer-based devices, however, have peculiar advantages that are

interesting for in-vitro experimentation. In particular, since the absorption is in the visible, the light

can be delivered to the preparation directly through the optical path of common microscopes, which

Page 96: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

91

usually do not transmit IR light above 1 μm. Desired pattern of illumination can be also easily

obtained with commercially available spatial light modulators or holographic systems. Moreover,

given the very high absorption length of semiconducting polymers like P3HT (α ≈ 105 cm

-1), thin

films of about 100 nm are already sufficient to absorb a large fraction of the impinging light.

The measurements presented in Chapter 4.3 were performed with pulses of light of 20 ms and 200

ms in order to investigate in better details the temporal evolution of the membrane potential.

However, to obtain an efficient thermally-mediated depolarization, it is better concentrate the

luminous energy in shorter pulses, for the following reasons:

Because of diffusion of thermal energy outside of the illuminated spot, increase of temperature

with time is not linear, but tends to saturate at an equilibrium value. The same amount of energy

concentrated in a shorter pulse is thus able to give a higher increase in temperature with respect

to a longer illumination.

The depolarization mechanism related to the change in the membrane capacitance is

counterbalanced by the flow of currents through the ion channels, which evolve on a timescale

dictated by the membrane time constant τmem. Using pulses of light with durations shorter than

τmem, this second mechanism has a lower impact on the actual depolarization reached for a

certain value of membrane capacitance variation.

For the hyperpolarizing effects observed on longer timescales, the measurements here performed

are again consistent with recent literature on photothermal stimulation. In particular, light-induced

heating, either mediated by water absorption or by exogenous transducers, has been shown to be

able to inhibit neural excitability and block transmission of action potentials along nerves.65

In

particular, using gold nanorods are used as near-IR absorbers, S. Yoo et al.87

reported that with

temperature variations up to 8 °C it is possible to reliably inhibit spiking activity in neurons. These

values are perfectly consistent with the heating measured in Chapter 3.5 upon prolonged

illumination in our devices and indeed preliminary experiments have shown that the polymer-based

hybrid interfaces are able to inhibit both stimulated and spontaneous activity in cultured neurons

and brain slices.

5.1.3 Comparison with previous works

In the previous works from our group, we have shown that the photoexcitation mediated by the

hybrid polymer/electrolyte interfaces was able to elicit neural activity both in-vitro with primary

cultures of neurons140,141

and ex-vivo with blind retinas from albino rats.141

In these works, an

electrical excitation due to charge generation in the active material was proposed. However, while

preliminary measurements were presented that already indicated the importance of the ITO

Page 97: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

92

electrode in the charge separation processes, no definitive coupling mechanism at the

polymer/electrolyte/cell interface was actually demonstrated. Despite the fact that the measurements

performed here on HEK-293 cells cannot be directly transposed to the case of neurons, given the

significant biophysical differences between the two systems, they can in any case provide important

information on the role that different stimulation mechanisms can play. However, given the

differences in the power intensities used, it is necessary to distinguish between the case of primary

neurons and explanted retinas.

In the measurements with primary neurons, excitation intensities of about 10-15 mW/mm2 were

used, while here we employed intensities up to 57 mW/mm2 for measurements on HEK-293 cells.

In those cases, however, a slightly different experimental configuration was used, with an upright

microscope and the illumination impinging directly on the sample surface. In the measurements

presented in Chapter 4, instead, an inverted microscope is used, with the light that has to travel

through the sample holder and the substrate before reaching the active layer. Taking into account

these losses and the fact that here we used a different wavelength (λ = 475 nm), which is less

efficiently absorbed by the P3HT layer than the one used in the previous works (λ = 530 nm), the

optical excitations in the two cases can be considered comparable. The observation in HEK-293

cells of a thermally-mediated transient depolarization upon illumination is qualitatively consistent

with the measured firing of action potentials in neurons and with literature on infrared neural

stimulation. However, the depolarization levels on the order of 1 mV found in Chapter 4.3 are too

small to explain by themselves a reliable triggering of action potential firing, for which

depolarizations on the order of 10 mV are usually needed. There are different possibilities that

could explain this difference:

The cell depolarization upon heating depends, other than the temperature gradient, also on the

biophysical properties of the cell membrane, in particular the actual coefficient of membrane

capacitance variation with temperature and the value of the Vσ potential related to membrane

asymmetries. The first parameter depends on the actual composition of the membrane in terms

of lipids and proteins and, for example, in the work of Shapiro et al.58

it is shown that a lipid

bilayer has a capacitance variation that is about two-fold with respect to that of an HEK-293

cell at the same light pulse energy. The second one is related to the difference in superficial

charge on the two leaflets of the membrane, which again can vary from cell to cell. A more

detailed analysis of these factors in neural cells is thus necessary to understand what could be

the actual effect of temperature variations on their membrane capacitance.

While Shapiro et al.58

attributed neural stimulation in INS to the transient depolarization given

by the effects on membrane capacitance, other groups proposed that this effect is mediated by

the opening of temperature-gated TRPV channels,57

which are expressed in different neural

Page 98: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

93

preparations (but are not endogenously present in HEK-293 cells). For example, TRPV4

channels have been demonstrated to open for temperatures higher than 25 °C,218

which are

consistent with the heating levels reached in our devices.

In the measurement with HEK-293 cells, we mainly characterized the effects of local heating.

However, also the capacitive charging of the interface, as proposed in our previous works,140,141

can take part into cell stimulation. The currents measured in our devices are quite fast compared

to standard protocols used with silicon-based devices; however, Fromherz and coworkers206

have actually shown that upon capacitive stimulation it is possible to elicit significant voltage-

gated sodium currents (mediated by NaV channels) even after few hundreds of microseconds.

For what concerns instead the measurements on explanted retinas,141

in that case the light intensities

used were between two and three orders of magnitude lower than those reported here for thermal

stimulation (significant responses were obtained with Ilight = 10-100 μW/mm2). Since the

temperature increase is linear with the absorbed intensity, the local heating in the experiments with

blind retinas should be on the order of few tens of mK. Such low values of temperature are not

expected to induce significant depolarizing effect or to gate the opening of TRPV channels. In this

case, we thus tentatively attributed the photostimulation effect to the capacitive charging of the

interface. Interestingly, also other groups have been able to elicit activity in blind retinas upon

optical stimulation using devices with architectures similar to ours. In particular, Narayan and

coworkers143

used organic devices based on bulk heterojunctions, while the group of Y. Hanein144

developed a systems based on inorganic nanorods absorbers dispersed on a mesh of carbon

nanotubes. In these works, illumination intensities on the order of tens to hundreds of μW/mm2

were employed and in both cases a capacitive coupling mechanism was proposed.

5.2 Perspectives

From the discussion of the experimental results presented in this work it certainly emerges that

indeed hybrid polymer/electrolyte interfaces harvest a huge interest for the modulation of

bioelectrical activity in biological tissues. These devices give the possibility to interact with cells

through stimuli of different nature, both electrical and thermal. The accepted models for electrical

capacitive stimulation, based on the concept of attached and free compartments of the membrane

described in Chapter 4.4, predict that both anodic and cathodic currents results in an excitation of

neural activity, while inhibition cannot be effectively obtained. In the case of our hybrid interfaces,

the possibility of having also thermal effects, that on long timescales have been demonstrated to

block the generation and propagation of action potentials, opens the way to the realization of

multifunctional platforms for controlling neural activity.

Page 99: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

94

However, there are still some issues that need to be addressed to reach a clear understanding of the

functioning of these devices.

The current densities elicited upon photoexcitation of the active material are comparable to the

value usually found in literature for electrical capacitive stimulation, but they rapidly fall to

zero on the millisecond timescale. An increase of the temporal duration of this current pulse is

thus desirable to obtain an efficient and reliable stimulation of neural cells. This goal can be

reached by increasing the interface capacitance and thus the total charge that can be

accumulated in two different ways. The first is to choose an active material with a higher

specific capacitance of the polymer/electrolyte interface; for example, thiophene-based

polymers with capacitance values in excess of 10 μF/cm2 (at f = 1 kHz) have been recently

demonstrated.212

The other way is to increase the active interfacial area between the polymer

and the electrolyte by using nano- or micro-structured electrodes. Also, porous materials like

hydrogels,219

which have been demonstrated to be ideal substrates for cell cultures, may be used

for this goal, upon functionalization with conjugated molecules.

The measurements performed on the HEK-293 cells revealed interesting information about the

effect of polymer-mediated photostimulation on general membrane properties like its

capacitance and overall conductivity; however, to understand the actual mechanisms that can

contribute to excitation in more complex systems like neurons, the effects on specific families

of membrane ion channels need to be taken into account.

o The eventual role of temperature-gated channels, like those of the TRPV family.57,218

These

channels, which are not endogenously expressed in HEK-293 cells, are found in different

kinds of neurons, like retinal and vestibular ganglion cells,220

dorsal root ganglion cells,63

CA1 neurons,62

and also in other cells like hyppocampal astrocytes.61

Interestingly, it is

possible to express TRPV channels in HEK-293 cells via transfection protocols and

photothermal stimulation of HEK-293 cells expressing different channels of this family198

has been already demonstrated with water-mediated IR light absorption.

o The effect of the light-evoked capacitive currents on voltage-gated ion channels. Voltage-

gated sodium (NaV) channels are fundamental in the generation of action potentials in

neurons. These channels can be expressed in HEK-293 cells and it has been shown that they

can be triggered via capacitive stimulation with inorganic silicon-based devices.206

Similar

measurements have also been performed on voltage-gated potassium (KV) channels.211

Moreover, it would be interesting to understand how the actual composition of the cell

membrane, in terms of the lipid and protein content but also of the ionic diffuse layer presents

on the two sides, affects the variation in capacitance and resistance upon local heating.58,146

Page 100: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

95

Given the complex nature of the plasma membrane, simplified systems like suspended artificial

lipid bilayers204

and vesicles89

may be used to decouple the effects of the various components.

In the meantime, other directions for a further development of these polymer-based hybrid

interfaces can be envisioned:

The devices presented in this work are based on P3HT as the active material, which absorb in

the blue/green region of the spectrum. However, a great variety of semiconducting polymers is

commercially available, with different absorption properties. A clear possible application of this

tunability is the development of a patterned interface with different organic materials absorbing

in different regions of the visible spectrum,221

in order to try to reproduce the color sensitivity of

the retina. Moreover, polymers absorbing in the near-IR222

could be used for applications

requiring in-vivo implants, since visible light has a low penetration depth in biological tissues.

Organic semiconductors have the great advantage that their molecular structure can be

chemically tuned to add different functionalities to the material without losing in conduction

properties. It is thus possible to produce conjugated polymers that are functionalized with

specific biochemical groups that can recognize particular cell types and allow their efficient

adhesion.118

In this way, it is possible to carefully control the cleft properties, which are

fundamental in determining an efficient electrical coupling between the cell and the device

interface.32,33

The hybrid interfaces discussed in this work have been mainly treated in terms of their

capability to modulate the membrane potential of cells for controlling neural activity. However,

electrical signals have been demonstrated to be involved in many other biological processes. A

fascinating example of an alternative use of our interface can be in the control of cellular

growth and differentiation.106,107,117

It has been widely reported in literature that extracellular

electrical stimulation of neural and neuronal-like cells can greatly affect their growth. Recently,

B. Zhu et al.118

have realized a functional substrate based on a conductive polymer and have

demonstrated that applying short pulses of voltage it is possible to enhance the neurite

outgrowth of PC12 cells grown on its surface. It would be thus interesting to understand if the

same effect could be obtained in our hybrid interfaces; if that was the case, new possibilities

could arise for controlling with precise spatio-temporal resolution the differentiation of

biological tissues. Moreover, it has been recently shown that the ability to modulate the cell

membrane potential can have important implications in the control of the life cycle of cells and

in cancer progression.148,149

The devices presented in this work are based on planar interfaces on which cells are grown.

While this configuration is perfectly suitable for in-vitro measurements, it has several

drawbacks for in-vivo experimentation. For these applications, it would thus be desirable to

Page 101: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

96

develop new architectures. One example could be the realization of polymer-coated optical

fiber,223

able to bring the active polymer deeper in tissues and at the same time deliver the

luminous stimuli. Another way could be that of realizing photoactive polymeric micro- and

nano-particle224,225

to be delivered in-vivo via injections; however, care should be taken in this

case on how the electrical and physico-chemical properties of the polymer/electrolyte interfaces

can vary with respect to planar devices.

Page 102: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

97

Appendix A

A.1 Optical Measurements

In this section are described the experimental setups used for the pump and probe measurements on

the different timescales presented in Chapter 3.3.

A.1.1 Femtosecond spectroscopy

Femtosecond dynamics of the primary photoexcitations in pristine P3HT and P3HT:PCBM blends

were measured with an ultrafast pump-and-probe setup with time resolution of ≈ 150 fs. A

schematic representation of the setup is depicted in Figure A.1.

Figure A.1 | Typical configuration of a femtosecond pump-and-probe experimental setup.

OPA: optical parametric amplifier.

Both the pump and the probe beam comes from an regeneratively amplified mode-locked Ti:Sapph

laser (Clark-MXR model CPA-1), with a fundamental emission wavelength at 780 nm at a

repetition rate of 1 kHz. The laser output is separated in two branches. The pump beam at 530 nm is

obtained from the fundamental via non-linear processes in an optical parametric amplifier (OPA).

The visible probe beam is achieved via supercontinuum generation focusing the laser through a 2

mm sapphire crystal. The pump beam is mechanically chopped at 500 Hz and it’s delayed with

respect to the probe by means of a motorized translation stage that increases its optical path. The

Page 103: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

98

two beams are then focused onto the sample on a spot with ≈ 50 μm diameter, taking care that they

have a good overlapping. The transmitted pump beam is blocked, while the probe is collected and

sent to an optical multichannel spectral analyzer. The acquisition of the probe spectra is

synchronized to the chopper modulation, in order to collect the probe transmission in presence (Ton)

and absence (Toff) of the pump. The final signal is then given by the normalized differential

transmission:

The energy of the pump pulses is kept low (≈ 20 nJ/pulse) in order to avoid bimolecular processes

upon photoexcitation, while the total energy of the probe beam is always at least an order of

magnitude lower.

A.1.2 Nanosecond spectroscopy

Transient absorption measurements on the nanoseconds/microseconds timescale were performed

with a commercial laser flash spectrometer (Edinburgh Instruments LP920). The pump pulses are

provided nanosecond tunable source (Opotek Opolett 355II) based on a Q-switched Nd:YAG laser

with a 10 Hz repetition rate. The probe light comes from a pulsed xenon arc lamp. The two beams

arrive at the sample position with a 90° angle, with the sample kept at a 45° angle respect to them.

The beams are properly focused onto the sample in order to ensure spatial overlap. The transmitted

probe light is spectrally filtered with a monochromator and then detected at a single wavelength by

a photomultiplier with a detection window in the visible.

For the traces reported in Chapter 3.3, an excitation wavelength of 530 nm was used, with an energy

density per pulse of about 20 μJ/cm2. The photobleaching signal was collected at 570 nm with a

spectral bandwidth of the monochromator of 3.5 nm. In order to get a good signal-to-noise ratio,

each of the traces reported is the results of averaging of 200 individual recordings.

(A.1)

Page 104: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

99

Figure A.2 | Experimental setup of the laser flash spectrometer used for ns-μs transient

absorption measurements.

A.1.3 CW Photoinduced Absorption

The measurements of CW absorption were performed in a home-made setup, whose schematic

structure is depicted in Figure A.3.

The probe beam is given by a 100 W tungsten halogen lamp that is focused on the sample by two

spherical mirrors; metal mirror are preferred to lenses to avoid chromatic aberrations and thus spots

of different size at different wavelength on the sample. The spot obtained on the sample plane has a

diameter of about 0.5 cm. The light transmitted through the sample is then recollected by two other

spherical mirrors that focus the beam on the entrance slit of a diffraction-grating based

monochromator (Spectral Products DK240). The pump beam is provided by a solid-state laser

source (Oxxius 561-100-COL-PP) with an optical power of about 100 mW, modualated by a

mechanical chopper at a frequency of f = 234 Hz. In order to obtain a good overlap between the

pump and the probe, the laser beam is enlarged by placing a defocusing lens in its path.

After being monochromated, the probe light is collected by a silicon photodetector (Thorlabs FDS

100) and the signal is analyzed with a lock-in amplifier (Stanford Research Systems SR830), locked

in frequency to the modulation of the chopper, with an integration time of 3 s. The measured signal

is thus the modulation in the transmitted light (ΔT) at a certain wavelength due to the presence of

the pump. The complete spectrum is reconstructed by scanning the monochromator through the

desired range of wavelengths, in this case 590 – 1100 nm. This signal is then normalized by the

sample transmission (T), measured in a second moment by modulating with the chopper the probe

beam in the absence of the pump.

Page 105: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

100

Figure A.3 | Scheme of the custom-built CW Photoinduced Absorption experimental

setup. PD: photodiode.

A.2 Electrical and thermal characterization

A.2.1 Photovoltage measurements

Photovoltage measurements have been performed in a three-electrodes configuration with a

potentiostat/galvanostat station (Metrohm Autolab PGSTAT). In particular, the station is used in

galvanostatic mode, i.e. controlling the value of the current flowing through the working electrode

(WE) and the counter-electrode (CE) and simultaneously monitoring the potential difference

between the WE and the reference electrode (RE). In the case of the measurements performed in

Chapter 3.4, the current value was kept to zero, i.e. the cell was in open circuit condition. The

samples measured are P3HT-based thin-films deposited on ITO-covered glass substrates (thickness

1 mm) with an active area of about 2 cm2. After the spin-coating of the films (as described in

Chapter 3.2), a strip of the absorbing layer is removed with acetone to allow the electrical contact to

the ITO. The device, which is the actual working electrode, is put in a custom electrochemical cell

in which the distance between the WE, the CE and the RE are fixed, in order to have repeatable

measurements. For the counter-electrode, a platinum wire is used, while the reference electrode is

an Ag/AgCl couple in a saturated KCl solution. The cell is filled with an electrolytic solution of

NaCl 0.2 M (if not otherwise specified) in ultrapure water.

Illumination to the sample is provided by a white LED (Thorlabs MCWHL5-C4), controlled in

intensity and pulse duration via a command signal provided by a function generator (Keithley

Page 106: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

101

3390). The light enters the cell through a planar quartz window and illuminates the entire sample

active area.

A.2.2 Photocurrent measurements

Photocurrent measurements were performed in a home-made setup with a two-electrodes

configuration. The samples for these measurements were deposited on ITO-covered glass substrates

(thickness 170 μm) with lateral dimensions of 9x18 mm2. On these substrates, the active material

was deposited via spin-coating methods, leaving a 9x5 mm2 strip of free ITO on the top for making

the electrical contact. The sample was then attached to the inner wall of a transparent cuvette with

square cross-section (1x1 cm2). The cuvette was filled with the electrolytic solution (NaCl 0.2 M)

until the active layer of the device was almost completely immersed, but leaving the free ITO strip

dry. As a counter-electrode, a platinum wire was immersed in the solution at the opposite side of the

cuvette. The current flowing from the ITO to the counter-electrode was measured by connecting the

two terminals to a transimpedance amplifier (Femto DHPCA-100). The amplified signal was then

collected with a digital oscilloscope (Tektronix MSO4054).

The illumination of the device was provided by a LED system (Lumencor Spectra X) with a

collimated output. In particular, a LED in the green was used (central wavelength λ = 530 nm) to

match the absorption peak of the P3HT. The LED intensity was controlled both via software and

with a set of neutral density filters. Light pulses (50 ms at 1 Hz repetition rate) were generated by

controlling the LED driver with a TTL signal provided by a function generator (Keithley 3390), that

also triggered the oscilloscope for acquisition.

A.2.3 Surface potential and temperature measurements

Measurements of surface potential a local temperature at the polymer/electrolyte interface were

performed with a similar experimental configuration using the patch-clamp setup (see Section

A.3.1).

The device to be analyzed was put in a plastic petri-dish and immersed in an electrolytic solution

(NaCl 0.2 M, if not specified otherwise) with Ag/AgCl counter-electrode. A glass micropipette

filled with the same electrolytic solution of the bath was micromanipulated in proximity of the

device surface. The approach was controlled visually by bringing the micropipette tip on the same

focal plane as the device surface. Both measurements were performed using the patch-clamp

amplifier in the voltage-clamp configuration and recording the signals upon illumination of the

active layer through the microscope objective, with a spot size dependent on the magnification used.

Page 107: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

102

The following scheme summarizes the electrical configuration in which the measurements were

performed.

Figure A.4 | Schematic representation of the electrical equivalent circuit for the

measurements of the surface potentials and temperatures.

In dark conditions, the current flowing through the pipette resistance (which is the actual parameter

measured in the experiment, Ipip) is determined by the potential different between Vsurf, the potential

at the device surface, and Vpip, which is kept by the amplifier at the offset value (Vset) set by the user:

where Rpip is the pipette resistance, determined by the solution concentration and the dimensions of

the pipette tip, and ΔVR the potential difference at its ends.

The main difference between the two experiments is in the offset applied to the recording pipette:

To measure the surface potentials, the offset was set in order not to have current flowing in the

pipette in dark conditions (ΔVR = 0). In this case, Vset and Vpip are thus equal to the value of Vsurf

in dark conditions, which is in turn equal to the reference potential (Vref = 0). Upon illumination,

Vsurf can change because of the capacitive charging of the interface, while Vpip is kept to the

reference value. The surface potential can thus be easily calculated as:

To measure the local temperature, the offset was set at a value that gave a constant current of

about Ioff = 4 nA. In this case, since the measurements were performed only on samples without

the ITO electrode, there is no capacitive charging of the interface upon illumination and Vsurf is

thus constant throughout the experiment, and thus ΔVR is also constant and it’s equal to the

product IoffR0, where R0 is the pipette resistance in dark conditions. Upon illumination, the value

(A.2)

(A.3)

Page 108: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

103

of the pipette resistance changes because of the local temperature variation, and can be

determined from the measured current from the following relationship:

The local temperature dynamics can then be recovered from the values of the pipette resistance

after proper calibration, as described in Chapter 3.5.

It should be noted that, also in the case of surface potential measurements the pipette resistance

maybe changing with time due to local heating. In principle, this variation should be taken into

account when calculating the surface potential value from the measured current. However, this

effect becomes important at longer times during the light pulse, while the surface potential signal is

falling to zero quite rapidly, especially in the blend, within the first few milliseconds. As a first

approximation we thus considered the pipette resistance constant in the surface potential

measurements.

A.3 Electrophysiology measurements

A.3.1 Electrophysiology setup

Measurements on HEK-293 cells, but also surface electrical and thermal characterizations of the

hybrid interface, were performed on a standard electrophysiology setup. This setup can be basically

divided in two parts:

An optical system for imaging the cells and guide the patch, with the possibility of shining light

pulses on the sample. It is composed of an inverted microscope (Nikon Eclipse Ti-S) coupled to

a CCD camera (Photometrics CoolSNAP MYO) for video acquisition. The light from the

microscope illuminator is filtered in order to use only wavelength longer than 750 nm for

imaging. In this way, the imaging light does not photoexcite the active material, which absorbs

in the visible. The excitation beam is provided by a collimated LED system (Lumencor Spectra

X) coupled to the microscope port for the fluorescence excitation source. The LED system is

equipped with six different LEDs, of which we use the blue (λ = 430 nm), the cyan (λ = 475

nm) and the green (λ = 530 nm). Power of the light impinging on the sample can be controlled

both via software and with neutral density filters in the optical path of the excitation light.

Measurements were performed with a 40x air objective (Nikon S Plan Fluor ELWD), giving an

(A.4)

Page 109: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

104

excitation spot on the sample with a diameter of 540 μm (that correspond to an active area of

about 0.23 mm2).

An electronic system for signal acquisition and amplification. The recording electrode is a

chlorinated silver wire enclosed in a glass pipette filled with an electrolytic solution, while the

counter-electrode is a pellet of Ag/AgCl. The acquisition is performed via a dedicated amplifier,

specifically designed for patch-clamp measurements (Molecular Devices Axopatch 200B). The

recording electrode is mounted on a headstage, which hold the pipette but also performs a first

amplification of the signal before sending it to the main unit of the amplifier. To allow precise

positioning of the pipette, the headstage/pipette system is mounted on a 3-axes

micromanipulator (Sutter Instruments MP-225). The amplifier is interfaced to a PC via a data

acquisition system (Molecular Devices Digidata 1550), which also commands the LED system

for triggering the light pulses.

For all the experiments, pipettes were freshly pulled just before each measurement with a

Flaming/Brown puller (Sutter Instruments P1000) from fire-polished glass capillaries (WPI

1B150F-4).

A.3.2 Electrolytic solutions and cell growth medium

HEK-293 cells were grown in Dulbecco’s modified Eagle’s medium (DMEM - high glucose, Sigma

D5796), supplemented with 10 % fetal bovine serum (Euroclone ECS 0180L), 100 U/ml penicillin

and 100 μg/ml streptomycin.

For patch-clamp experiments, the composition of the intracellular and extracellular solutions used

are reported in the tables below.

Intracellular solution

KCl K-Gluconate MgCl2 CaCl2 HEPES EGTA ATP-Na2

12 mM 125 mM 1 mM 0.1 mM 10 mM 10 mM 10 mM

Extracellular solution

NaCl KCl MgCl2 CaCl2 Glucose HEPES

135 mM 5.4 mM 1 mM 1.8 mM 10 mM 5 mM

In measurements not involving the presence of cells we normally employed, if not specified

otherwise, a solution of NaCl in ultrapure water at a concentration of 0.2 M. Sodium chloride was

Page 110: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

105

selected because Na+ and Cl

- are the two main ions in the extracellular solutions and the

concentration was chosen in order to roughly mimic the typical ionic strength of physiological

media.

Page 111: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

106

Appendix B

The analysis performed in this thesis has focused only on the behaviour of the polymer/electrolyte

hybrid interface, both on its own and coupled with cells, under illumination with short light pulses

up to few hundreds of milliseconds. The device, in its complete architecture with an ITO contact,

has been shown to sustain the generation of a capacitive photocurrent that charges the

polymer/electrolyte interface in few milliseconds. However, it has also been proven in other works

that under prolonged illumination electrochemical reactions can be promoted. We have thus also

performed a study on the effect of continuous photoexcitation of the active material on cells grown

on the device. In particular, we have performed experiments using primary neocortical astrocytes.

Astrocytes are cells that belong to the family of glial cells, which comprises the various types of

non-excitable cells present in the nervous system. Until the end of the last century, the function of

glial cells was thought to be only of support and protection for neurons. However, increasing

evidence in the recent years indicates that glial cells, and especially astrocytes, may have a more

active role in the functioning of the nervous system than previously believed. In particular,

astrocytes are found in the central nervous system, where they are the most numerous macroglial

cells. They have a fundamental role in regulating the concentrations of ions in the extracellular

space and in the recycling of neurotransmitters released by the synaptic terminals of neurons during

the propagation of the nervous signals. Moreover, astrocytes have a major role in the response of

the neural tissue to pathological situations like traumas, infections or neurodegenerative diseases

and also in the reaction to medical implants.

It is thus clear the importance of understanding the effects of photoexcitation on glial cells if the

polymeric interfaces here developed are to be applied for in vitro experimentations on models of the

nervous system or for in vivo implants.

Since the work that follows (which has been published in Ref. 142) is based on the use of prolonged

illumination protocols, the effects observed here cannot be directly compared to what observed in

the main part of this thesis. However, they are reported in this Appendix because they give in any

case interesting complementary information on the possible applications of the photoactive

polymeric interfaces.

Page 112: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

107

B.1 Astrocyte cultures and electrophysiological properties

In this work, we used mainly P3HT:PCBM films (with a thickness of about 50 nm) as the active

layer of the devices. We started by characterized the possibility of growing astrocytes on these

materials assessing the viability of cell cultures at several days in vitro. In order to promote

adhesion of the cells to the substrates, they were pre-coated with a layer of poly-d-lysine (PDL), a

commonly used molecule for cultures of brain cells.

We firstly performed a fluorescence assay based on the detection of fluorescein diacetate (FDA)

emission. This molecule, which is initially non-emissive, is hydrolyzed inside the cells by non-

specific esterases, transforming it into a fluorescent probe. Since this process can happen only in

live cells, the intensity of fluorescence can be related to the viability of cells. Figure B.1a,b shows

typical fluorescence images taken on control substrates (ITO+PDL) and on the active devices

(ITO/P3HT:PCBM+PDL) at 1 day in vitro (DIV) after re-plating. In the histograms of Figure B.1c

are reported the viability of the cells on the two different substrates at 1DIV and 4DIV, showing

that indeed astrocytes can be cultured on P3HT:PCBM thin films without significant effects on their

vitality.

Figure B.1 | (a,b) Fluorescence confocal images of astrocytes stained with fluorescein

diacetate grown on control ITO substrates (a) and on ITO/P3HT:PCBM devices (b); the

images are taken 1 day after plating. (c) Viability of astrocytes cultured on the two

different substrates at 1 and 4 DIV.

We then characterized the basic electrophysiological properties of the astrocytes cultured on the

active devices and compared them to those grown on control substrates. The cell membrane

properties were measured via patch-clamp techniques in whole-cells recordings with control

extracellular and intracellular solutions (see Section B.3 for details). A voltage ramp from -120 mV

to + 60 mV (duration 600 ms) was applied to the astrocytes and the relative membrane current was

measured in a voltage-clamp configuration. The complete protocol is depicted in the inset of Figure

B.2. The current traces recorded during the voltage ramp are reported in Figure B.2a,b for the case

of the cells grown on the control substrates and on the active material respectively. The membranes

Page 113: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

108

show a strong rectifying behavior for negative potentials, with negligible currents at values more

hyperpolarized than -40 mV.

Figure B.2 | Membrane currents recorded on ITO (a) and ITO/P3HT:PCBM (b) substrates

upon the application of the protocol reported in the inset. Only the current trace relative to

the ramp part of the stimulus is reported.

The electrophysiological parameters extracted during the measurements on both substrates are

reported in Table B.1. No significant difference can be observed between the values recorded on the

two cases (control: n = 9; blend: n = 10).

Table B.1 | Electrophysiological properties recorded on ITO (control) and

ITO/P3HT:PCBM (blend) substrates. Vm: membrane resting potential; Cp: membrane

capacitance; Rin: patch input resistance; spG: membrane specific conductance; I-120mV and

I60mV: current density (i.e. normalized by the cell capacitance) measured at -120 mV and

60 mV respectively.

The biophysical properties of the currents here measured are perfectly consistent with those already

characterized in immature glial cells grown on standard plastic Petri dishes for tissue culturing, that

have been mainly attributed to the presence of the delayed rectifier potassium channel (KDR).226,227

These data confirms that conjugated polymers are suitable substrates for culturing primary cells like

astrocytes and that they do not modify their basic electrophysiological property with respect to

control substrates normally employed in studies in vitro.

Vm

[mV]

Cp

[pF]

Rin

[MΩ]

spG

[pS/pF]

I-120mV

[pA/pF]

I60mV

[pA/pF]

Control −47±3 42.1 ± 5.9 443 ± 105 0.1 ± 0.02 −6.3 ± 1.2 52.3 ± 7.6

Blend −48±7 38.3 ± 9.8 599 ± 138 0.07 ± 0.03 −4.7 ± 2 44.3 ± 16.5

Page 114: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

109

B.2 Photostimulation of astrocytes membrane conductances

After having assessed the characteristics of membrane conduction in astrocytes in dark, we

investigated the effects on those properties of photostimulation of the active material. The

membrane potential was monitored during continuous illumination of the active layer with light

intensity of 13 mW/mm2 (λ = 561 nm) in a spot with a diameter of about 100 μm around the

patched astrocyte. Upon the onset of the light stimulus, the membrane potential of the cell starts to

depolarize, passing from -49 ± 4 mV to -19 ± 2 mV (n = 5) during 50 s photoexcitation, while no

effect can be observed for measurements on control substrates (Figure B.3). This behavior is

apparently in contrast with the measurements presented in Chapter 4 on HEK-293 cells, in which a

hyperpolarization was measured at longer times. However, it has to be reminded that the timescales

investigated here (several seconds) are completely different the pulses used in the HEK-293

recordings, which arrived up to 200 ms at maximum. Indeed, preliminary measurements (data not

shown) on astrocytes with short pulses confirm the presence of a behavior comparable to that of

HEK-293 cells on those timescales.

Figure B.3| Recording of the variations in membrane potential of an astrocyte upon

illumination with continuous light (13 mW/mm2) for P3HT:PCBM (red trace) and control

(grey trace) substrates.

To understand what biophysical phenomenon could be responsible of the depolarization observed in

Figure B.3, we repeated the membrane characterization with a voltage ramp protocol of Figure B.2

at different times during the illumination increasing the light intensity. In particular, the protocol

was applied every 10 s to measure how the membrane currents were affected by the

photostimulation. It can be seen in Figure B.4a that indeed the conductance properties of the

membrane vary during the illumination; in particular a significant increase of inward currents at

negative potentials could be observed, an effect that was augmented by using higher light intensities

(Figure B.4c). At the same time, a shift of the zero current potential (which correspond to the

equilibrium potential) towards more positive values was observed, consistently with the

Page 115: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

110

depolarization measured in Figure B.3. Again, no significant effect was recorded on control

substrates (Figure Figure B.4b,d); indeed, eventual small variations in the measured currents are

clearly independent of the presence of the light stimulus.

Figure B.4 | (a,b) Membrane currents recorded in dark and at different light intensities

during illumination with the voltage protocol of Figure B.2 for P3HT:PCBM (a) and

control (b) substrates. (c,d) Variation in time of the current measured at -120 mV (i.e. at

the start of the voltage ramp) during continuous illumination at different light intensities

(represented by the colored boxes) for P3HT:PCBM (c) and control (d) substrates.

The inward current observed upon photostimulation remained also after switching off the light and

we attributed the effect to the activation of a membrane conductance. Considering the simultaneous

shift of the zero-current potential toward more positive values, we could hypothesize the

involvement of a chloride conductance or of a nonspecific cation channel. We thus exposed the

cells after the light-mediated stimulation to an extracellular solution in which all the monovalent

cations were replaced equimolarly by the non-permeable ion NMDG+. The ramp currents elicited in

the membrane with control saline and with this modified solutions (NMDG-Cl) were however the

same (Figure B.5a), indicating that the photoactivated current was not mediated by cations and was

thus most likely due to Cl- ions. We then investigated the biophysical properties of this current by

analyzing its time dynamics. In order to isolate this current from the other conductances of the cell,

we replaced equimolarly the cations in the extracellular solutions with NMDG+ and the potassium

in the intracellular solution with Cs+. We measured the time traces of the evoked currents before

(Figure B.5b) and after (Figure B.5c) photostimulation of the astrocytes upon the application of a

Page 116: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

111

voltage step protocol from -120 mV to + 60 mV (see inset in Figure B.5b). While in resting

conditions virtually no conductances could be measured at every potential, after the light excitation

hyperpolarizing stimuli were able to elicit a time-dependent current. This current displayed a first,

fast component and a second slowly activated, non-deactivating one that grew larger for higher

hyperpolarizations. This behavior is characteristic of the chloride conductance mediated by the ClC-

2 channel, which has been previously characterized in cortical astrocytes both in vitro and in

situ.228,229

Interestingly, this channel is normally not activated in astrocytes cultured in vitro in

resting condition, unless a long-term pharmacological treatment is used.

Figure B.5 | (a) Ramp currents recorded on astrocytes after 50 s photostimulation (13

mW/mm2) with control extracellular saline (gray trace) and the NMDG

+-Cl solution

(black trace). (b,c) Current traces evoked in astrocytes upon a step voltage protocol (from

-120 mV to 60 mV in 20 mV steps) with intracellular CsCl and extracellular NMDG-Cl

before (b) and after (c) photostimulation for 50 s at 13 mW/mm2.

To corroborate the hypothesis of the activation of the ClC-2 conductance, we exposed the cells after

photostimulation to a 200 μM extracellular concentration of Cd2+

ions, which have been

demonstrated to inhibit these ion channels.229

Indeed, the currents evoked in the astrocytes upon the

application of the Cd2+

-containing solution were clearly reduced with respect to the ones measured

with the control NMDG-Cl saline (Figure B.6).

The evidence collected demonstrates that long-term photostimulation of astrocytes with visible light

mediated by the hybrid polymer-based interfaces results in the activation of specific chloride

currents mediated by the ClC-2 ion channel. This channel is expressed in different tissues, like the

brain, intestine, stomach, kidney, salivary glands, and heart.230

While a clear understanding of the

physiological functions of ClC-2 still has to be determined, there is a growing body of evidence that

this channel can play different roles, in the different tissues in which it is expressed. Results

obtained from mice in which its expression was inhibited revealed that ClC-2 disruption resulted in

leukoencephalopathy, blindness and male infertility.231,232

In this context, the possibility to stimulate

Page 117: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

112

ClC-2 currents by polymer photoexcitation may represent an interesting tool to help clarifying the

roles of this ion channel in central nervous system.

Figure B.6 | Current traces evoked with a voltage step protocol (inset) after

photostimulation (50 s light at 13 mW/mm2) with NMDG-Cl saline (a) and upon the

addition of sub-millimolar concentrations of Cd2+

ions (b).

Regarding the activation process of the channel upon photostimulation, the actual mechanism is still

unclear. It is known that ClC-2 conductances are modulated by different stimuli, like membrane

hyperpolarization, hypotonicity-induced cell swelling, moderate acidification of the extracellular

medium and disruption of F-actin filaments.226,228,233,234

During the measurements, we did not

observe any visible morphological change of the cell upon photostimulation. Membrane

hyperpolarization was indeed measured in HEK-293 cells upon prolonged illumination. However,

also electrochemical reactions at the polymer/electrolyte interface have been demonstrated to occur

on these long timescales and these reactions may be accompanied by a variation in the local pH felt

by the cell. More in-depth characterizations of these processes are thus necessary to clearly

understand the biophysical mechanisms leading the specific activation of ClC-2 channels.

B.3 Experimental methods

Organic conducting polymers preparation

Rr-P3HT has a regio-regularity of 99.5 % and molecular weight of 17500 g/mol. An accurate

cleaning of the substrate was required: the substrate was rinsed in an ultrasonic bath with,

sequentially, a specific tension-active agent in water solution (HELLMANEX® II, 3 %), deionised

water, pure acetone and isopropyl alcohol. 1,2-Chlorobenzene solutions of P3HT and PCBM were

prepared separately. P3HT was diluted to a final concentration of 15g/l. PCBM was prepared at a

concentration of 15g/l and then mixed (1:1 volume ratio) with P3HT using a magnetic stirrer.

Solutions were then heated at 50 °C, stirred and finally deposited on ITO-covered and ITO-less

glass substrates, previously heated, by spin-coating. Spinning parameters (first step: 800 rpm,

Page 118: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

113

angular acceleration 1500 rad/s, rotation duration 2 s; second step: 1500 rpm, angular acceleration

4000 rad/s, rotation duration 30 s) were carefully selected in order to obtain suitable optical quality

and film thickness. After deposition, organic layers were annealed and properly sterilized by heating

at 120 °C for 2 hrs. Control substrates were sterilized in the same way.

Cell culture preparation and maintenance

Briefly, cerebral cortices devoid of meninges were triturated and placed in cell culture flasks

containing Dulbecco’s modified Eagle’s medium (DMEM)–glutamax medium with 15% fetal

bovine serum (FBS) and penicillin–streptomycin (100 U/mL and 100 lg/mL respectively) (Gibco-

Invitrogen, Milan, Italy). Culture flasks were maintained in a humidified incubator with 5% CO2

for 2–5 weeks. Immunostaining for glial fibrillary acidic protein (GFAP) and the flat, polygonal

morphological phenotype of the cultured cells indicated that more than 95% were type 1 cortical

astrocytes. At confluence, astroglial cells were enzymatically dispersed using trypsin–EDTA in

P3HT-PCBM/ITO or ITO substrates treated for 30 min with poly-D-lysine (0,01 mg/ml in PBS) at a

concentration of 1x104 per substrate and maintained in culture medium containing 10% FBS.

Cell Viability Assay and Counting

Astrocytes plated on the different substrates were mounted in a custom-made perfusion chamber

and incubated for 5 min with FDA (Sigma Aldrich). After rinsing with physiological saline

solution, a sequence of confocal images (10 to15 different fields of 0.6 mm × 0.6 mm for each

sample) was taken using a Nikon TE 2000 inverted confocal microscope (20× objective). Living

cells were counted and number of cells per mm 2 was calculated and compared at each time point

analyzed.

Electrophysiology and photostimulation

Current recordings were obtained with the whole-cell configuration of the patch-clamp technique.

Light excitation was provided by a CW laser diode (OXXIUS) peaking at 561 nm, with an intensity

was properly varied by a neutral density filter with variable optical density, in the range 0.7–13

mW/mm2. Laser light was coupled to the microscope, impinging on the sample from the ITO side.

Patch pipettes were prepared from thin-walled borosilicate glass capillaries to have a tip resistance

of 2-4 MW when filled with the standard internal solution. Membrane currents were amplified (List

EPC-7) and stored on a computer for off-line analysis (pClamp 6, Axon Instrument and Origin 6.0,

MicroCal). Because of the large current amplitude, the access resistance (below 10 MW) was

corrected 70-90%. Experiments were carried out at room temperature (20-24°C). The reference

electrode was an agar bridge filled with either 150 mM NaCl or 1 M KCl saline for experiments in

which substitution of extracellular Cl– was required. Experiments were carried out at room

Page 119: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

114

temperature (20–24°C). Current densities were calculated by dividing the current values measured

at each membrane potential by the cell capacitance derived from the correction of the capacitive

transients of the recorded cells by means of the analogue circuit of the patch-clamp amplifier.

Solutions and chemicals

For electrophysiological experiments the standard bath saline was (mM): 140 NaCl, 4 KCl, 2

MgCl2, 2 CaCl2, 10 HEPES, 5 glucose, pH 7.4 with NaOH and osmolarity adjusted to ~315 mOsm

with mannitol. The intracellular (pipette) solution was composed of (mM): 144 KCl, 2 MgCl2, 5

EGTA, 10 HEPES, pH 7.2 with KOH and osmolarity ~300 mOsm. When using external solutions

with different ionic compositions, salts were replaced equimolarly. The different salines containing

pharmacological agents were applied with a gravity-driven, local perfusion system at a flow rate of

~200 ml/min positioned within ~100 mm of the recorded cell. In order to isolate Cl– current, the

external bath perfusion, termed control saline, was (in mM): 140 NMDG-Cl, 4 NaCl, 2 MgCl2, 2

CaCl2, 10 TES, 5 glucose. The intracellular (pipette) solution was composed of (in mM): 126 CsCl,

2 MgCl2, 1 EGTA, 10 TES, pH 7.2 with CsOH, and osmolarity adjusted to approximately 290

mOsm with mannitol.

Page 120: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

115

Bibliography

1. Reilly, P. J. Applied Bioelectricity: From Electrical Stimulations to Electropathology.

(Springer Science & Business Media, 1998).

2. Gennis, R. B. Biomembranes: molecular structure and function. (Springer-Verlag, 1989).

3. Kandel, E., Schwartz, J. H. & Jessell, T. M. Principles of Neural Science. (McGraw Hill

Professional, 2013).

4. Barnett, M. W. & Larkman, P. M. The action potential. Pract. Neurol. 7, 192–197 (2007).

5. Bers, D. M. Cardiac excitation–contraction coupling. Nature 415, 198–205 (2002).

6. Ebashi, S. & Endo, M. Calcium and muscle contraction. Prog. Biophys. Mol. Biol. 18, 123–

183 (1968).

7. MacDonald, P. E. & Rorsman, P. Oscillations, Intercellular Coupling, and Insulin Secretion in

Pancreatic β Cells. PLoS Biol. 4, (2006).

8. Jacob, R. Calcium oscillations in electrically non-excitable cells. Biochim. Biophys. Acta BBA

- Mol. Cell Res. 1052, 427–438 (1990).

9. Chen, L. B. Mitochondrial Membrane Potential in Living Cells. Annu. Rev. Cell Biol. 4, 155–

181 (1988).

10. Masui, Y. & Clarke, H. J. in International Review of Cytology (ed. G.H. Bourne, J. F. D. and

K. W. J.) 57, 185–282 (Academic Press, 1979).

11. D, M., Colin, Rajnicek, A. M., Song, B. & Zhao, M. Controlling cell behavior electrically:

current views and future potential. Physiol. Rev. 85, 943–978 (2005).

12. Famm, K., Litt, B., Tracey, K. J., Boyden, E. S. & Slaoui, M. Drug discovery: a jump-start for

electroceuticals. Nature 496, 159–161 (2013).

13. Birmingham, K. et al. Bioelectronic medicines: a research roadmap. Nat. Rev. Drug Discov.

13, 399–400 (2014).

14. Verkhratsky, A., Krishtal, O. A. & Petersen, O. H. From Galvani to patch clamp: the

development of electrophysiology. Pflüg. Arch. 453, 233–247 (2006).

15. Galvani, L. De Viribus Electricitatis In Motu Musculari Comentarius. Bononiensi Sci. Artium

Inst. Atque Acad. Comment. VII, (1792).

16. Bois-Reymond, E. H. D. Untersuchungen über thierische elektricität. (G. Reimer, 1848).

17. Hodgkin, A. L. & Huxley, A. F. Currents carried by sodium and potassium ions through the

membrane of the giant axon of Loligo. J. Physiol. 116, 449–472 (1952).

Page 121: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

116

18. Hodgkin, A. L. & Huxley, A. F. A quantitative description of membrane current and its

application to conduction and excitation in nerve. J. Physiol. 117, 500–544 (1952).

19. Neher, E. & Sakmann, B. Single-channel currents recorded from membrane of denervated

frog muscle fibres. Nature 260, 799–802 (1976).

20. Neher, E., Sakmann, B. & Steinbach, J. H. The extracellular patch clamp: A method for

resolving currents through individual open channels in biological membranes. Pflüg. Arch.

375, 219–228 (1978).

21. Sakmann, B. & Neher, E. Single-channel Recording. (Springer Science, 2009).

22. Neher, E. Ion channels for communication between and within cells. Biosci. Rep. 12, 1–14

(1992).

23. Sakmann, B. Elementary steps in synaptic transmission revealed by currents through single

ion channels. Neuron 8, 613–629 (1992).

24. Markram, H. The Human Brain Project. Sci. Am. 306, 50–55 (2012).

25. Alivisatos, A. P. et al. The Brain Activity Map Project and the Challenge of Functional

Connectomics. Neuron 74, 970–974 (2012).

26. Alivisatos, A. P. et al. Nanotools for Neuroscience and Brain Activity Mapping. ACS Nano 7,

1850–1866 (2013).

27. Spira, M. E. & Hai, A. Multi-electrode array technologies for neuroscience and cardiology.

Nat. Nanotechnol. 8, 83–94 (2013).

28. HajjHassan, M., Chodavarapu, V. & Musallam, S. NeuroMEMS: Neural Probe

Microtechnologies. Sensors 8, 6704–6726 (2008).

29. Delgado Ruz, I. & Schultz, S. R. Localising and classifying neurons from high density MEA

recordings. J. Neurosci. Methods 233, 115–128 (2014).

30. Swindale, N. V. & Spacek, M. A. Spike sorting for polytrodes: a divide and conquer

approach. Front. Syst. Neurosci. 8, 6 (2014).

31. Poghossian, A., Ingebrandt, S., Offenhäusser, A. & Schöning, M. J. Field-effect devices for

detecting cellular signals. Semin. Cell Dev. Biol. 20, 41–48 (2009).

32. Gleixner, R. & Fromherz, P. The Extracellular Electrical Resistivity in Cell Adhesion.

Biophys. J. 90, 2600–2611 (2006).

33. Hess, L. H. et al. Electrical Coupling Between Cells and Graphene Transistors. Small (2014).

doi:10.1002/smll.201402225

34. Buzsáki, G., Anastassiou, C. A. & Koch, C. The origin of extracellular fields and currents —

EEG, ECoG, LFP and spikes. Nat. Rev. Neurosci. 13, 407–420 (2012).

35. Duan, X. et al. Intracellular recordings of action potentials by an extracellular nanoscale field-

effect transistor. Nat. Nanotechnol. 7, 174–179 (2012).

Page 122: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

117

36. Patolsky, F. et al. Detection, Stimulation, and Inhibition of Neuronal Signals with High-

Density Nanowire Transistor Arrays. Science 313, 1100–1104 (2006).

37. Scanziani, M. & Häusser, M. Electrophysiology in the age of light. Nature 461, 930–939

(2009).

38. Martino, N., Ghezzi, D., Benfenati, F., Lanzani, G. & Antognazza, M. R. Organic

semiconductors for artificial vision. J. Mater. Chem. B 1, 3768 (2013).

39. Antognazza, M. R. et al. Shedding Light on Living Cells. Adv. Mater. (2014).

doi:10.1002/adma.201403513

40. Loew, L. M. in Membrane Potential Imaging in the Nervous System (eds. Canepari, M. &

Zecevic, D.) 13–23 (Springer New York, 2010).

41. Peterka, D. S., Takahashi, H. & Yuste, R. Imaging Voltage in Neurons. Neuron 69, 9–21

(2011).

42. Knöpfel, T. Genetically encoded optical indicators for the analysis of neuronal circuits. Nat.

Rev. Neurosci. 13, 687–700 (2012).

43. Mutoh, H., Perron, A., Akemann, W., Iwamoto, Y. & Knöpfel, T. Optogenetic monitoring of

membrane potentials. Exp. Physiol. 96, 13–18 (2011).

44. Stosiek, C., Garaschuk, O., Holthoff, K. & Konnerth, A. In vivo two-photon calcium imaging

of neuronal networks. Proc. Natl. Acad. Sci. 100, 7319–7324 (2003).

45. Stepnoski, R. A. et al. Noninvasive detection of changes in membrane potential in cultured

neurons by light scattering. Proc. Natl. Acad. Sci. 88, 9382–9386 (1991).

46. Ross, W. N. et al. Changes in absorption, fluorescence, dichroism, and birefringence in

stained giant axons: Optical measurement of membrane potential. J. Membr. Biol. 33, 141–

183 (1977).

47. D’ Arsonval, J.-A. La fibre musculaire est directement excitable par la lumiere. CR Soc Biol

43, 318–320 (1891).

48. Chalazonitis, N. Light Energy Conversion in Neuronal Membranes. Photochem. Photobiol. 3,

539–559 (1964).

49. Kerkut, G. Can fibre optic systems drive lower motoneurones? Prog. Neurobiol. 14, 1–23

(1980).

50. Fork, R. Laser stimulation of nerve cells in Aplysia. Science 171, 907–908 (1971).

51. Reece, P. J., Dholakia, K., Thomas, R. C. & Cottrell, G. A. Green laser light (532 nm)

activates a chloride current in the C1 neuron of Helix aspersa. Neurosci. Lett. 433, 265–269

(2008).

52. Hirase, H., Nikolenko, V., Goldberg, J. H. & Yuste, R. Multiphoton stimulation of neurons. J.

Neurobiol. 51, 237–247 (2002).

53. Iwanaga, S., Smith, N., Fujita, K., Kawata, S. & Nakamura, O. Single-pulse cell stimulation

with a near-infrared picosecond laser. Appl. Phys. Lett. 87, 243901 (2005).

Page 123: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

118

54. Smith, N. I. et al. Photostimulation of two types of Ca2+ waves in rat pheochromocytoma

PC12 cells by ultrashort pulsed near-infrared laser irradiation. Laser Phys. Lett. 3, 154–161

(2006).

55. Wells, J. et al. Optical stimulation of neural tissue in vivo. Opt. Lett. 30, 504–506 (2005).

56. Wells, J. et al. Biophysical Mechanisms of Transient Optical Stimulation of Peripheral Nerve.

Biophys. J. 93, 2567–2580 (2007).

57. Albert, E. S. et al. TRPV4 channels mediate the infrared laser-evoked response in sensory

neurons. J. Neurophysiol. 107, 3227–3234 (2012).

58. Shapiro, M. G., Homma, K., Villarreal, S., Richter, C.-P. & Bezanilla, F. Infrared light excites

cells by changing their electrical capacitance. Nat. Commun. 3, 736 (2012).

59. Migliori, B., Ventra, M. D. & Jr, W. K. Photoactivation of neurons by laser-generated local

heating. AIP Adv. 2, 032154 (2012).

60. Liljemalm, R., Nyberg, T. & von Holst, H. Heating during infrared neural stimulation. Lasers

Surg. Med. 45, 469–481 (2013).

61. Bai, J.-Z. & Lipski, J. Differential expression of TRPM2 and TRPV4 channels and their

potential role in oxidative stress-induced cell death in organotypic hippocampal culture.

NeuroToxicology 31, 204–214 (2010).

62. Lipski, J. et al. Involvement of TRP-like channels in the acute ischemic response of

hippocampal CA1 neurons in brain slices. Brain Res. 1077, 187–199 (2006).

63. Cao, D.-S., Yu, S.-Q. & Premkumar, L. S. Modulation of transient receptor potential vanilloid

4-mediated membrane currents and synaptic transmission by protein kinase C. Mol. Pain 5, 5

(2009).

64. Ryskamp, D. A. et al. The Polymodal Ion Channel Transient Receptor Potential Vanilloid 4

Modulates Calcium Flux, Spiking Rate, and Apoptosis of Mouse Retinal Ganglion Cells. J.

Neurosci. 31, 7089–7101 (2011).

65. Duke, A. R. et al. Transient and selective suppression of neural activity with infrared light.

Sci. Rep. 3, 2600 (2013).

66. Mou, Z., Triantis, I. F., Woods, V. M., Toumazou, C. & Nikolic, K. A Simulation Study of the

Combined Thermoelectric Extracellular Stimulation of the Sciatic Nerve of the Xenopus

Laevis: The Localized Transient Heat Block. IEEE Trans. Biomed. Eng. 59, 1758–1769

(2012).

67. Banghart, M. R. et al. Photochromic Blockers of Voltage-Gated Potassium Channels. Angew.

Chem. Int. Ed. 48, 9097–9101 (2009).

68. Lester, H. A. & Nerbonne, J. M. Physiological and Pharmacological Manipulations with Light

Flashes. Annu. Rev. Biophys. Bioeng. 11, 151–175 (1982).

69. Tochitsky, I. et al. Restoring Visual Function to Blind Mice with a Photoswitch that Exploits

Electrophysiological Remodeling of Retinal Ganglion Cells. Neuron 81, 800–813 (2014).

Page 124: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

119

70. Polosukhina, A. et al. Photochemical Restoration of Visual Responses in Blind Mice. Neuron

75, 271–282 (2012).

71. Callaway, E. M. & Katz, L. C. Photostimulation using caged glutamate reveals functional

circuitry in living brain slices. Proc. Natl. Acad. Sci. 90, 7661–7665 (1993).

72. Kandler, K., Katz, L. C. & Kauer, J. A. Focal photolysis of caged glutamate produces long-

term depression of hippocampal glutamate receptors. Nat. Neurosci. 1, 119–123 (1998).

73. Noh, J., Seal, R. P., Garver, J. A., Edwards, R. H. & Kandler, K. Glutamate co-release at

GABA/glycinergic synapses is crucial for the refinement of an inhibitory map. Nat. Neurosci.

13, 232–238 (2010).

74. Boyden, E. S., Zhang, F., Bamberg, E., Nagel, G. & Deisseroth, K. Millisecond-timescale,

genetically targeted optical control of neural activity. Nat. Neurosci. 8, 1263–1268 (2005).

75. Zhang, F., Wang, L.-P., Boyden, E. S. & Deisseroth, K. Channelrhodopsin-2 and optical

control of excitable cells. Nat. Methods 3, 785–792 (2006).

76. Berndt, A., Lee, S. Y., Ramakrishnan, C. & Deisseroth, K. Structure-Guided Transformation

of Channelrhodopsin into a Light-Activated Chloride Channel. Science 344, 420–424 (2014).

77. Wietek, J. et al. Conversion of Channelrhodopsin into a Light-Gated Chloride Channel.

Science 344, 409–412 (2014).

78. Airan, R. D., Thompson, K. R., Fenno, L. E., Bernstein, H. & Deisseroth, K. Temporally

precise in vivo control of intracellular signalling. Nature 458, 1025–1029 (2009).

79. Konermann, S. et al. Optical control of mammalian endogenous transcription and epigenetic

states. Nature 500, 472–476 (2013).

80. Kim, B. Y. S., Rutka, J. T. & Chan, W. C. W. Nanomedicine. N. Engl. J. Med. 363, 2434–

2443 (2010).

81. Doane, T. L. & Burda, C. The unique role of nanoparticles in nanomedicine: imaging, drug

delivery and therapy. Chem. Soc. Rev. 41, 2885–2911 (2012).

82. Winter, J. O., Liu, T. Y., Korgel, B. A. & Schmidt, C. E. Recognition Molecule Directed

Interfacing Between Semiconductor Quantum Dots and Nerve Cells. Adv. Mater. 13, 1673–

1677 (2001).

83. Pappas, T. C. et al. Nanoscale Engineering of a Cellular Interface with Semiconductor

Nanoparticle Films for Photoelectric Stimulation of Neurons. Nano Lett. 7, 513–519 (2007).

84. Zhao, Y., Larimer, P., Pressler, R. T., Strowbridge, B. W. & Burda, C. Wireless Activation of

Neurons in Brain Slices Using Nanostructured Semiconductor Photoelectrodes. Angew. Chem.

Int. Ed. 48, 2407–2410 (2009).

85. Lugo, K., Miao, X., Rieke, F. & Lin, L. Y. Remote switching of cellular activity and cell

signaling using light in conjunction with quantum dots. Biomed. Opt. Express 3, 447–454

(2012).

Page 125: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

120

86. Eom, K. et al. Enhanced Infrared Neural Stimulation using Localized Surface Plasmon

Resonance of Gold Nanorods. Small 10, 3853–3857 (2014).

87. Yoo, S., Hong, S., Choi, Y., Park, J.-H. & Nam, Y. Photothermal Inhibition of Neural Activity

with Near-Infrared-Sensitive Nanotransducers. ACS Nano 8, 8040–8049 (2014).

88. Farah, N. et al. Holographically patterned activation using photo-absorber induced neural–

thermal stimulation. J. Neural Eng. 10, 056004 (2013).

89. Fromherz, P., Kiessling, V., Kottig, K. & Zeck, G. Membrane transistor with giant lipid

vesicle touching a silicon chip. Appl. Phys. A 69, 571–576 (1999).

90. Fromherz, P. Electrical Interfacing of Nerve Cells and Semiconductor Chips. ChemPhysChem

3, 276–284 (2002).

91. Fromherz, P., Eick, S. & Hofmann, B. in Nanoelectronics and Information Technology:

Advanced Electronic Materials and Novel Devices (ed. Waser, R.) Chapter 32, 781–810

(Wiley-VCH, 2012).

92. Colicos, M. A., Collins, B. E., Sailor, M. J. & Goda, Y. Remodeling of Synaptic Actin

Induced by Photoconductive Stimulation. Cell 107, 605–616 (2001).

93. Hafeman, D. G., Parce, J. W. & McConnell, H. M. Light-addressable potentiometric sensor

for biochemical systems. Science 240, 1182–1185 (1988).

94. Goda, Y. & Colicos, M. A. Photoconductive stimulation of neurons cultured on silicon

wafers. Nat. Protoc. 1, 461–467 (2006).

95. Hung, J., Chansard, M., Ousman, S. S., Nguyen, M. D. & Colicos, M. A. Activation of

microglia by neuronal activity: Results from a new in vitro paradigm based on neuronal-

silicon interfacing technology. Brain. Behav. Immun. 24, 31–40 (2010).

96. Suzurikawa, J., Nakao, M., Jimbo, Y., Kanzaki, R. & Takahashi, H. Light-Addressed

Stimulation Under Imaging of Cultured Neurons. IEEE Trans. Biomed. Eng. 56, 2660–2665

(2009).

97. Suzurikawa, J., Nakao, M., Jimbo, Y., Kanzaki, R. & Takahashi, H. A light addressable

electrode with a TiO2 nanocrystalline film for localized electrical stimulation of cultured

neurons. Sens. Actuators B Chem. 192, 393–398 (2014).

98. Mathieson, K. et al. Photovoltaic retinal prosthesis with high pixel density. Nat. Photonics 6,

391–397 (2012).

99. Forrest, S. R. The path to ubiquitous and low-cost organic electronic appliances on plastic.

Nature 428, 911–918 (2004).

100. Aregueta-Robles, U. A., Woolley, A. J., Poole-Warren, L. A., Lovell, N. H. & Green, R. A.

Organic electrode coatings for next-generation neural interfaces. Front. Neuroengineering 7,

15 (2014).

101. Liao, C. et al. Flexible Organic Electronics in Biology: Materials and Devices. Adv. Mater.

(2014). doi:10.1002/adma.201402625

Page 126: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

121

102. Rivnay, J., Owens, R. M. & Malliaras, G. G. The Rise of Organic Bioelectronics. Chem.

Mater. 26, 679–685 (2014).

103. Berggren, M. & Richter-Dahlfors, A. Organic Bioelectronics. Adv. Mater. 19, 3201–3213

(2007).

104. Angione, M. D. et al. Interfacial electronic effects in functional biolayers integrated into

organic field-effect transistors. Proc. Natl. Acad. Sci. 109, 6429–6434 (2012).

105. Mulla, M. Y. et al. Capacitance-modulated transistor detects odorant binding protein chiral

interactions. Nat. Commun. 6, 6010 (2015).

106. Saltó, C. et al. Control of Neural Stem Cell Adhesion and Density by an Electronic Polymer

Surface Switch. Langmuir 24, 14133–14138 (2008).

107. Wan, A. M. D., Brooks, D. J., Gumus, A., Fischbach, C. & Malliaras, G. G. Electrical control

of cell density gradients on a conducting polymer surface. Chem. Commun. 35, 5278–5280

(2009).

108. Benfenati, V. et al. A transparent organic transistor structure for bidirectional stimulation and

recording of primary neurons. Nat. Mater. 12, 672–680 (2013).

109. Campana, A., Cramer, T., Simon, D. T., Berggren, M. & Biscarini, F. Electrocardiographic

Recording with Conformable Organic Electrochemical Transistor Fabricated on Resorbable

Bioscaffold. Adv. Mater. 26, 3874–3878 (2014).

110. Khodagholy, D. et al. High transconductance organic electrochemical transistors. Nat.

Commun. 4, 2133 (2013).

111. Khodagholy, D. et al. In vivo recordings of brain activity using organic transistors. Nat.

Commun. 4, 1575 (2013).

112. Khodagholy, D. et al. NeuroGrid: recording action potentials from the surface of the brain.

Nat. Neurosci. 18, 310–315 (2015).

113. Simon, D. T. et al. Organic electronics for precise delivery of neurotransmitters to modulate

mammalian sensory function. Nat. Mater. 8, 742–746 (2009).

114. Isaksson, J. et al. Electronic control of Ca2+ signalling in neuronal cells using an organic

electronic ion pump. Nat. Mater. 6, 673–679 (2007).

115. Tybrandt, K. et al. Translating Electronic Currents to Precise Acetylcholine–Induced

Neuronal Signaling Using an Organic Electrophoretic Delivery Device. Adv. Mater. 21, 4442–

4446 (2009).

116. Simon, D. T., Larsson, K. C., Berggren, M. & Richter-Dahlfors, A. Precise Neurotransmitter-

Mediated Communication with Neurons In Vitro and In Vivo Using Organic Electronics. J.

Biomech. Sci. Eng. 5, 208–217 (2010).

117. Bolin, M. H. et al. Active Control of Epithelial Cell-Density Gradients Grown Along the

Channel of an Organic Electrochemical Transistor. Adv. Mater. 21, 4379–4382 (2009).

Page 127: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

122

118. Zhu, B. et al. Large enhancement in neurite outgrowth on a cell membrane-mimicking

conducting polymer. Nat. Commun. 5, 4523 (2014).

119. Günes, S., Neugebauer, H. & Sariciftci, N. S. Conjugated Polymer-Based Organic Solar Cells.

Chem. Rev. 107, 1324–1338 (2007).

120. Yu, G., Gao, J., Hummelen, J. C., Wudl, F. & Heeger, A. J. Polymer Photovoltaic Cells:

Enhanced Efficiencies via a Network of Internal Donor-Acceptor Heterojunctions. Science

270, 1789–1791 (1995).

121. Li, G. et al. High-efficiency solution processable polymer photovoltaic cells by self-

organization of polymer blends. Nat. Mater. 4, 864–868 (2005).

122. Jørgensen, M., Norrman, K. & Krebs, F. C. Stability/degradation of polymer solar cells. Sol.

Energy Mater. Sol. Cells 92, 686–714 (2008).

123. Hintz, H. et al. Photodegradation of P3HT−A Systematic Study of Environmental Factors.

Chem. Mater. 23, 145–154 (2011).

124. Voroshazi, E. et al. Influence of cathode oxidation via the hole extraction layer in

polymer:fullerene solar cells. Org. Electron. 12, 736–744 (2011).

125. Antognazza, M. R., Ghezzi, D., Musitelli, D., Garbugli, M. & Lanzani, G. A hybrid solid-

liquid polymer photodiode for the bioenvironment. Appl. Phys. Lett. 94, 243501 (2009).

126. Lanzarini, E. et al. Polymer-Based Photocatalytic Hydrogen Generation. J. Phys. Chem. C

116, 10944–10949 (2012).

127. Guerrero, A. et al. Organic photoelectrochemical cells with quantitative photocarrier

conversion. Energy Environ. Sci. 7, 3666–3673 (2014).

128. Sirringhaus, H. et al. Two-dimensional charge transport in self-organized, high-mobility

conjugated polymers. Nature 401, 685–688 (1999).

129. Kline, R. J. et al. Dependence of Regioregular Poly(3-hexylthiophene) Film Morphology and

Field-Effect Mobility on Molecular Weight. Macromolecules 38, 3312–3319 (2005).

130. Hugger, S., Thomann, R., Heinzel, T. & Thurn-Albrecht, T. Semicrystalline morphology in

thin films of poly(3-hexylthiophene). Colloid Polym. Sci. 282, 932–938 (2004).

131. Deibel, C. et al. Energetics of excited states in the conjugated polymer poly(3-

hexylthiophene). Phys. Rev. B 81, 085202 (2010).

132. Choulis, S. A. et al. High ambipolar and balanced carrier mobility in regioregular poly(3-

hexylthiophene). Appl. Phys. Lett. 85, 3890–3892 (2004).

133. Seemann, A. et al. Reversible and irreversible degradation of organic solar cell performance

by oxygen. Sol. Energy 85, 1238–1249 (2011).

134. Nicolai, H. T. et al. Unification of trap-limited electron transport in semiconducting polymers.

Nat. Mater. 11, 882–887 (2012).

135. Abdou, M. S. A. & Holdcroft, S. Mechanisms of photodegradation of poly(3-alkylthiophenes)

in solution. Macromolecules 26, 2954–2962 (1993).

Page 128: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

123

136. Abdou, M. S. A., Orfino, F. P., Son, Y. & Holdcroft, S. Interaction of Oxygen with

Conjugated Polymers:  Charge Transfer Complex Formation with Poly(3-alkylthiophenes). J.

Am. Chem. Soc. 119, 4518–4524 (1997).

137. Hoshino, S. et al. Influence of moisture on device characteristics of polythiophene-based

field-effect transistors. J. Appl. Phys. 95, 5088–5093 (2004).

138. Bellani, S. et al. Reversible P3HT/Oxygen Charge Transfer Complex Identification in Thin

Films Exposed to Direct Contact with Water. J. Phys. Chem. C 118, 6291–6299 (2014).

139. Scarpa, G., Idzko, A.-L., Götz, S. & Thalhammer, S. Biocompatibility Studies of

Functionalized Regioregular Poly(3-hexylthiophene) Layers for Sensing Applications.

Macromol. Biosci. 10, 378–383 (2010).

140. Ghezzi, D. et al. A hybrid bioorganic interface for neuronal photoactivation. Nat. Commun. 2,

166 (2011).

141. Ghezzi, D. et al. A polymer optoelectronic interface restores light sensitivity in blind rat

retinas. Nat. Photonics 7, 400–406 (2013).

142. Benfenati, V. et al. Photostimulation of Whole-Cell Conductance in Primary Rat Neocortical

Astrocytes Mediated by Organic Semiconducting Thin Films. Adv. Healthc. Mater. 3, 392–

399 (2014).

143. Gautam, V., Rand, D., Hanein, Y. & Narayan, K. S. A Polymer Optoelectronic Interface

Provides Visual Cues to a Blind Retina. Adv. Mater. 26, 1751–1756 (2014).

144. Bareket, L. et al. Semiconductor Nanorod–Carbon Nanotube Biomimetic Films for Wire-Free

Photostimulation of Blind Retinas. Nano Lett. 14, 6685–6692 (2014).

145. O’Connor, C. M. & Adams, J. U. Essentials of Cell Biology. (2010).

146. Heimburg, T. Thermal Biophysics of Membranes. (John Wiley & Sons, 2008).

147. Hille, B. Ion Channels of Excitable Membranes. (Sinauer Associates, 2001).

148. Sundelacruz, S., Levin, M. & Kaplan, D. L. Role of Membrane Potential in the Regulation of

Cell Proliferation and Differentiation. Stem Cell Rev. Rep. 5, 231–246 (2009).

149. Yang, M. & Brackenbury, W. J. Membrane potential and cancer progression. Front. Physiol.

4, 185 (2013).

150. Singer, S. J. & Nicolson, G. L. The fluid mosaic model of the structure of cell membranes.

Science 175, 720–731 (1972).

151. Mouritsen, O. G. & Bloom, M. Mattress model of lipid-protein interactions in membranes.

Biophys. J. 46, 141–153 (1984).

152. Bevers, E. M. & Williamson, P. L. Phospholipid scramblase: An update. FEBS Lett. 584,

2724–2730 (2010).

153. Op den Kamp, J. A. F. Lipid Asymmetry in Membranes. Annu. Rev. Biochem. 48, 47–71

(1979).

Page 129: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

124

154. Kay, J. G., Koivusalo, M., Ma, X., Wohland, T. & Grinstein, S. Phosphatidylserine dynamics

in cellular membranes. Mol. Biol. Cell 23, 2198–2212 (2012).

155. Bigay, J. & Antonny, B. Curvature, Lipid Packing, and Electrostatics of Membrane

Organelles: Defining Cellular Territories in Determining Specificity. Dev. Cell 23, 886–895

(2012).

156. Conner, S. D. & Schmid, S. L. Regulated portals of entry into the cell. Nature 422, 37–44

(2003).

157. Bezanilla, F. Ion Channels: From Conductance to Structure. Neuron 60, 456–468 (2008).

158. Doyle, D. A. et al. The Structure of the Potassium Channel: Molecular Basis of K+

Conduction and Selectivity. Science 280, 69–77 (1998).

159. Gouaux, E. & MacKinnon, R. Principles of Selective Ion Transport in Channels and Pumps.

Science 310, 1461–1465 (2005).

160. Roux, B. et al. Ion selectivity in channels and transporters. J. Gen. Physiol. 137, 415–426

(2011).

161. The 2003 Nobel Prize in Chemistry - Advanced Information. Nobelprize.org at

<http://www.nobelprize.org/nobel_prizes/chemistry/laureates/2003/advanced.html#>

162. Beckstein, O. Principles of Gating Mechanisms of Ion Channels. (University of Oxford,

2004).

163. Thompson, S. M., Masukawa, L. M. & Prince, D. A. Temperature dependence of intrinsic

membrane properties and synaptic potentials in hippocampal CA1 neurons in vitro. J.

Neurosci. Off. J. Soc. Neurosci. 5, 817–824 (1985).

164. DeCoursey, T. E. & Cherny, V. V. Temperature Dependence of Voltage-gated H+ Currents in

Human Neutrophils, Rat Alveolar Epithelial Cells, and Mammalian Phagocytes. J. Gen.

Physiol. 112, 503–522 (1998).

165. Purves, D. et al. Neuroscience. (Sinauer Associates, 2001).

166. Goldman, D. E. Potential, Impedance, and Rectification in Membranes. J. Gen. Physiol. 27,

37–60 (1943).

167. Skou, J. C. in Membrane Transport (ed. Tosteson, D. C.) 155–185 (Springer New York,

1989).

168. Gentet, L. J., Stuart, G. J. & Clements, J. D. Direct measurement of specific membrane

capacitance in neurons. Biophys. J. 79, 314–320 (2000).

169. Service, R. F. Outlook Brightens for Plastic Solar Cells. Science 332, 293–293 (2011).

170. Gautam, V., Bag, M. & Narayan, K. S. Dynamics of Bulk Polymer

Heterostructure/Electrolyte Devices. J. Phys. Chem. Lett. 1, 3277–3282 (2010).

171. Gautam, V., Bag, M. & Narayan, K. S. Single-Pixel, Single-Layer Polymer Device as a

Tricolor Sensor with Signals Mimicking Natural Photoreceptors. J. Am. Chem. Soc. 133,

17942–17949 (2011).

Page 130: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

125

172. Mayer, A. C., Scully, S. R., Hardin, B. E., Rowell, M. W. & McGehee, M. D. Polymer-based

solar cells. Mater. Today 10, 28–33 (2007).

173. Thompson, B. C. & Fréchet, J. M. J. Polymer–Fullerene Composite Solar Cells. Angew.

Chem. Int. Ed. 47, 58–77 (2008).

174. Shaw, P. E., Ruseckas, A. & Samuel, I. D. W. Exciton Diffusion Measurements in Poly(3-

hexylthiophene). Adv. Mater. 20, 3516–3520 (2008).

175. Halls, J. J. M., Pichler, K., Friend, R. H., Moratti, S. C. & Holmes, A. B. Exciton diffusion

and dissociation in a poly(p‐phenylenevinylene)/C60 heterojunction photovoltaic cell. Appl.

Phys. Lett. 68, 3120–3122 (1996).

176. Grancini, G. et al. Hot exciton dissociation in polymer solar cells. Nat. Mater. 12, 29–33

(2013).

177. Tang, C. W. Two‐layer organic photovoltaic cell. Appl. Phys. Lett. 48, 183–185 (1986).

178. Ameri, T. et al. Fabrication, Optical Modeling, and Color Characterization of Semitransparent

Bulk-Heterojunction Organic Solar Cells in an Inverted Structure. Adv. Funct. Mater. 20,

1592–1598 (2010).

179. McNeill, C. R. et al. Photophysics and Photocurrent Generation in

Polythiophene/Polyfluorene Copolymer Blends. Adv. Funct. Mater. 19, 3103–3111 (2009).

180. Pandit, B., Gautam, B. R., Basel, T. P. & Vardeny, Z. V. Correlation between ultrafast

transient photomodulation spectroscopy and organic photovoltaic solar cell efficiency based

on RR-P3HT/PCBM blends. Org. Electron. 15, 1149–1154 (2014).

181. Dupuis, A., Tournebize, A., Bussière, P.-O., Rivaton, A. & Gardette, J.-L. Morphology and

photochemical stability of P3HT:PCBM active layers of organic solar cells. Eur. Phys. J.

Appl. Phys. 56, 34104 (2011).

182. Cabanillas-Gonzalez, J., Grancini, G. & Lanzani, G. Pump-Probe Spectroscopy in Organic

Semiconductors: Monitoring Fundamental Processes of Relevance in Optoelectronics. Adv.

Mater. 23, 5468–5485 (2011).

183. Gadermaier, C. & Lanzani, G. Photophysics of conjugated polymers: the contribution of

ultrafast spectroscopy. J. Phys. Condens. Matter 14, 9785 (2002).

184. Cerullo, G., Manzoni, C., Lüer, L. & Polli, D. Broadband pump–probe spectroscopy system

with sub-20 fs temporal resolution for the study of energy transfer processes in

photosynthesis. Photochem. Photobiol. Sci. 6, 135–144 (2007).

185. Hwang, I.-W., Moses, D. & Heeger, A. J. Photoinduced Carrier Generation in P3HT/PCBM

Bulk Heterojunction Materials. J. Phys. Chem. C 112, 4350–4354 (2008).

186. Korovyanko, O. J., Österbacka, R., Jiang, X. M., Vardeny, Z. V. & Janssen, R. A. J.

Photoexcitation dynamics in regioregular and regiorandom polythiophene films. Phys. Rev. B

64, 235122 (2001).

187. Young, H. D., Freedman, R. A., Ford, A. L., Zemansky, M. W. & Sears, F. W. Sears and

Zemansky’s University Physics, 12th Edition. (Pearson Education, Limited, 2008).

Page 131: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

126

188. Albert-Seifried, S. & Friend, R. H. Measurement of thermal modulation of optical absorption

in pump-probe spectroscopy of semiconducting polymers. Appl. Phys. Lett. 98, 223304

(2011).

189. Kobayashi, T., Kinoshita, K., Nagase, T. & Naito, H. Continuous-wave photoinduced

absorption studies in polythiophene and fullerene blended thin films. Phys. Rev. B 83, 035305

(2011).

190. Jahng, W. S., Francis, A. H., Moon, H., Nanos, J. I. & Curtis, M. D. Is indium tin oxide a

suitable electrode in organic solar cells? Photovoltaic properties of interfaces in organic p/n

junction photodiodes. Appl. Phys. Lett. 88, 093504–093504–3 (2006).

191. Song, Q. L. et al. Exciton dissociation at the indium tin oxide-N,N′-Bis(naphthalen-1-yl)-

N,N′-bis(phenyl) benzidine interface: A transient photovoltage study. Appl. Phys. Lett. 88,

232101 (2006).

192. Tomozawa, H., Braun, D., Phillips, S., Heeger, A. J. & Kroemer, H. Metal-polymer schottky

barriers on cast films of soluble poly(3-alkylthiophenes). Synth. Met. 22, 63–69 (1987).

193. Tomozawa, H. et al. Metal-polymer Schottky barriers on processible polymers. Synth. Met.

28, 687–690 (1989).

194. Gustafsson, G., Inganäs, O., Sundberg, M. & Svensson, C. Rectifying metal/poly(3-

hexylthiophene) contacts. Synth. Met. 41, 499–502 (1991).

195. Yih, F. & Show-An, C. Studies on aluminum/poly(3-octylthiophene)/indium-tin oxide

Schottky barrier electronic device: rectification property and its temperature dependence.

Mater. Chem. Phys. 32, 380–385 (1992).

196. MacKenzie, R. C. I., Shuttle, C. G., Chabinyc, M. L. & Nelson, J. Extracting Microscopic

Device Parameters from Transient Photocurrent Measurements of P3HT:PCBM Solar Cells.

Adv. Energy Mater. 2, 662–669 (2012).

197. Porrazzo, R. et al. Field-effect and capacitive properties of water-gated transistors based on

polythiophene derivatives. APL Mater. 3, 014905 (2015).

198. Yao, J., Liu, B. & Qin, F. Rapid Temperature Jump by Infrared Diode Laser Irradiation for

Patch-Clamp Studies. Biophys. J. 96, 3611–3619 (2009).

199. Patrício, P. S. O. et al. Correlation between thermal, optical and morphological properties of

heterogeneous blends of poly(3-hexylthiophene) and thermoplastic polyurethane. J. Phys.

Condens. Matter 18, 7529 (2006).

200. Bavel, S. S. van, Sourty, E., With, G. de & Loos, J. Three-Dimensional Nanoscale

Organization of Bulk Heterojunction Polymer Solar Cells. Nano Lett. 9, 507–513 (2009).

201. Duda, J. C., Hopkins, P. E., Shen, Y. & Gupta, M. C. Thermal transport in organic

semiconducting polymers. Appl. Phys. Lett. 102, 251912 (2013).

202. Thomas, P. & Smart, T. G. HEK293 cell line: A vehicle for the expression of recombinant

proteins. J. Pharmacol. Toxicol. Methods 51, 187–200 (2005).

Page 132: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

127

203. Varghese, A., TenBroek, E. M., Coles Jr., J. & Sigg, D. C. Endogenous channels in HEK cells

and potential roles in HCN ionic current measurements. Prog. Biophys. Mol. Biol. 90, 26–37

(2006).

204. Rajapaksha, S. P., Wang, X. & Lu, H. P. Suspended Lipid Bilayer for Optical and Electrical

Measurements of Single Ion Channel Proteins. Anal. Chem. 85, 8951–8955 (2013).

205. Twentyman, P. R. & Luscombe, M. A study of some variables in a tetrazolium dye (MTT)

based assay for cell growth and chemosensitivity. Br. J. Cancer 56, 279–285 (1987).

206. Schoen, I. & Fromherz, P. Activation of Na+ channels in cell membrane by capacitive

stimulation with silicon chip. Appl. Phys. Lett. 87, (2005).

207. Schoen, I. & Fromherz, P. The Mechanism of Extracellular Stimulation of Nerve Cells on an

Electrolyte-Oxide-Semiconductor Capacitor. Biophys. J. 92, 1096–1111 (2007).

208. Zeck, G. & Fromherz, P. Noninvasive neuroelectronic interfacing with synaptically connected

snail neurons immobilized on a semiconductor chip. Proc. Natl. Acad. Sci. U. S. A. 98,

10457–10462 (2001).

209. Huang, A. P., Chu, P. K. & Yang, Z. C. Hafnium-based high-k gate dielectrics. (INTECH

Open Access Publisher, 2010).

210. Eickenscheidt, M., Jenkner, M., Thewes, R., Fromherz, P. & Zeck, G. Electrical Stimulation

of Retinal Neurons in Epiretinal and Subretinal Configuration using a Multi-Capacitor-Array.

J. Neurophysiol. 107, 2742–2755 (2012).

211. Ulbrich, M. H. & Fromherz, P. Opening of K+ channels by capacitive stimulation from silicon

chip. Appl. Phys. A 81, 887–891 (2005).

212. Toss, H. et al. On the mode of operation in electrolyte-gated thin film transistors based on

different substituted polythiophenes. Org. Electron. 15, 2420–2427 (2014).

213. Stavrinidou, E. et al. Direct Measurement of Ion Mobility in a Conducting Polymer. Adv.

Mater. 25, 4488–4493 (2013).

214. Izzo, A. D. et al. Laser Stimulation of Auditory Neurons: Effect of Shorter Pulse Duration and

Penetration Depth. Biophys. J. 94, 3159–3166 (2008).

215. Duke, A. R. et al. Combined optical and electrical stimulation of neural tissue in vivo. J.

Biomed. Opt. 14, 060501 (2009).

216. Jenkins, M. W. et al. Optical pacing of the embryonic heart. Nat. Photonics 4, 623–626

(2010).

217. Curcio, J. A. & Petty, C. C. The Near Infrared Absorption Spectrum of Liquid Water. J. Opt.

Soc. Am. 41, 302–302 (1951).

218. Voets, T. et al. The principle of temperature-dependent gating in cold- and heat-sensitive TRP

channels. Nature 430, 748–754 (2004).

219. Choi, M. et al. Light-guiding hydrogels for cell-based sensing and optogenetic synthesis in

vivo. Nat. Photonics 7, 987–994 (2013).

Page 133: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

128

220. Bec, J.-M. et al. Characteristics of laser stimulation by near infrared pulses of retinal and

vestibular primary neurons. Lasers Surg. Med. 44, 736–745 (2012).

221. Antognazza, M. R., Scherf, U., Monti, P. & Lanzani, G. Organic-based tristimuli colorimeter.

Appl. Phys. Lett. 90, 163509 (2007).

222. Soci, C. et al. Photoconductivity of a Low-Bandgap Conjugated Polymer. Adv. Funct. Mater.

17, 632–636 (2007).

223. Binda, M., Natali, D., Iacchetti, A. & Sampietro, M. Integration of an Organic Photodetector

onto a Plastic Optical Fiber by Means of Spray Coating Technique. Adv. Mater. 25, 4335–

4339 (2013).

224. Hu, Z. & Gesquiere, A. J. PCBM concentration dependent morphology of P3HT in composite

P3HT/PCBM nanoparticles. Chem. Phys. Lett. 476, 51–55 (2009).

225. Hu, Z., Tenery, D., Bonner, M. S. & Gesquiere, A. J. Correlation between spectroscopic and

morphological properties of composite P3HT/PCBM nanoparticles studied by single particle

spectroscopy. J. Lumin. 130, 771–780 (2010).

226. Ferroni, S., Marchini, C., Schubert, P. & Rapisarda, C. Two distinct inwardly rectifying

conductances are expressed in long term dibutyryl-cyclic-AMP treated rat cultured cortical

astrocytes. FEBS Lett. 367, 319–325 (1995).

227. Benfenati, V., Caprini, M., Nobile, M., Rapisarda, C. & Ferroni, S. Guanosine promotes the

up-regulation of inward rectifier potassium current mediated by Kir4.1 in cultured rat cortical

astrocytes. J. Neurochem. 98, 430–445 (2006).

228. Ferroni, S., Nobile, M., Caprini, M. & Rapisarda, C. pH modulation of an inward rectifier

chloride current in cultured rat cortical astrocytes. Neuroscience 100, 431–438 (2000).

229. Makara, J. K. et al. Astrocytes from mouse brain slices express ClC-2-mediated Cl− currents

regulated during development and after injury. Mol. Cell. Neurosci. 23, 521–530 (2003).

230. Jentsch, T. J. CLC Chloride Channels and Transporters: From Genes to Protein Structure,

Pathology and Physiology. Crit. Rev. Biochem. Mol. Biol. 43, 3–36 (2008).

231. Bösl, M. R. et al. Male germ cells and photoreceptors, both dependent on close cell–cell

interactions, degenerate upon ClC‐2 Cl− channel disruption. EMBO J. 20, 1289–1299 (2001).

232. Jeworutzki, E. et al. GlialCAM, a Protein Defective in a Leukodystrophy, Serves as a ClC-2

Cl− Channel Auxiliary Subunit. Neuron 73, 951–961 (2012).

233. Ferroni, S., Marchini, C., Nobile, M. & Rapisarda, C. Characterization of an inwardly

rectifying chloride conductance expressed by cultured rat cortical astrocytes. Glia 21, 217–

227 (1997).

234. Fava, M., Ferroni, S. & Nobile, M. Osmosensitivity of an inwardly rectifying chloride current

revealed by whole-cell and perforated-patch recordings in cultured rat cortical astrocytes.

FEBS Lett. 492, 78–83 (2001).

Page 134: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

129

Page 135: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

130

Acknowledgements

“No man is an island entire of itself; every man

is a piece of the continent, a part of the main”

John Donne (1624)

It was the spring of 2008 when I first entered a lab for my bachelor degree. It was in the Physics

Department and it was called “Cappa-Chimico”, from the fumehood in a corner behind the big

optical table. I witnessed its moving to the infamous garage of the Administrative Building, where it

became the first ARCO-lab, not knowing that I would later spend an entire year in its cold and dark

rooms working on the first version of the cw-confocal for my master thesis. Nor did I knew that

ARCO was meant to be just the original nucleus of what is now the Center for NanoScience and

Technology (CNST for friends), which I witnessed being transformed from a series of empty (and

wet) rooms into a thriving center at the forefront of international research, and which pretty much

became my second home during the four years of my PhD. It has been seven long years since this

adventure started, and it has now arrived the time to thank all the people that have accompanied me

during this period, also because I’m still in debt of a proper acknowledgment to many of them since

my master thesis.

First of all, I have to thank my supervisor, Dr. Maria Rosa Antognazza for supporting me during

these five years, even if I sometimes made it difficult with my chronic lateness, and for always

taking into high consideration my personal and professional interests and needs. Thanks also to

prof. Guglielmo Lanzani, for all his ideas and for allowing me to work in such a scientifically

exciting and socially pleasant environment as the CNST. Thanks to both of them for having given

me the opportunity to work on extremely fascinating projects and to meet very interesting people.

I would like to thank the colleagues from my research group that were essential in providing both

experimental help and theoretical discussions fundamental for the outcomes of this work:

Sebastiano, for the electrochemical measurements; Matteo, for all the numerical calculations; Katia,

for teaching me how to grow cells. Moreover, I have to thank Erica, Elena, Susana, Gabriele,

Andrea, Giovanni, Daniele and Lucia.

I also have to thank our collaborators that guided me from the reductionist world of physics into the

holistic realm of biology and neuroscience. In particular, I would like to thank Dr. Valentina

Benfenati and Paul Feyen for introducing me to this strange discipline called electrophysiology and

Page 136: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

131

for teaching me the art of the patch-clamp, but also for all the invaluable discussions we had. I also

have to thank Prof. Fabio Benfenati, Dr. Diego Ghezzi, Dr. Fabrizia Cesca, Dr. Elisabetta Colombo

from the NBT Department in Genova, as well as Prof. Michele Muccini, Prof. Stefano Ferroni and

Dr. Stefano Toffanin at the CNR and Univeristy in Bologna.

I need to thank Dr. Giulia Grancini for being my first guide into ultrafast and confocal

spectroscopy, Dr. Mario Caironi and Dr. Calogero Sciascia for the fruitful collaboration on Charge

Modulation Microscopy and Dr. Alessandro Luzio for making me state-of-the-art transistors. I also

wish to thank all the people I have collaborated with in the disparate projects I was involved in

during these five years: Dr. Annamaria Petrozza, Dr. Daniele Fazzi, Dr. Ajay Ram Srimath

Kandada, Dr. Margherita Zavelani-Rossi, Dr. Valeria Russo, Dr. Massimiliano Bianchi, Dr. Assunta

Pistone, Dr. Grazia Pertile, Dr. Maurizio Mete and Dr. Silvia Bisti.

I wish also to thank all my friends and colleagues from CNST that have made (or still do) working

in this institute not only an invaluable opportunity for my scientific formation, but also a unique

experience at the personal level: Ross, for always seeing the best in people; Vale (Vinz), for being

around since the old times of the garage-ARCO; Vitto, for always knowing the right thing to say;

Ale L, not only for making transistors; Marco (and wife) and Mary for the Friday sushis and the

good wines; Giorgio, Andrea and Joy, for making CNST not so lonely on weekends (and for the

darts matches); Sadir and Pupi, the steady couple since the university years; Michele DB, for always

having the latest news; Vale N, for always having a good word for everyone; Piva, for not letting

me be the only Nicola in the center; Luca P, for his half-veals and cassoeulas; Marcelo, for his

perfect English; Ale Mez, for always being the spirit of the parties. Thanks to all the others, past and

present: Sara, James, Alex, Isis, Berri, GEB, Andrea DV, Erika and Simone, Nava, Giacomo,

Giusy, FM, Michele G, Antonio, Nicolas and all those I have for sure forgotten. Thanks to Rob for

the nice time we had during his stay in Italy. A big thanks also to the administrative and technical

offices, Silvia, Tessa, Elena, Alessandra, Stefano, Luca and Enrico.

A long-awaited and sincere thanks to Marghe, who has always been next to me through the nine

years of highs and lows from the first lessons of calculus to the hard life of the researcher, and with

whom I look forward to have an equally exciting time as a new chapter of our life starts in the U.S.

An equally long-awaited thanks to Lalla (and husband) for always being present and supportive,

even if she betrayed science for the dark side of a real job. Not less important, a special thanks is

due to the parents of both of them, who were always concerned with feeding me in a proper way. I

wish to also to thank everyone else from the years of University, especially Ale, Paul, Mirco, Fra,

Nico and Vale, and all my other friends here in Milan, Sica, Macchia, Gigio, Dani, Tobia, Ieva,

Luz, Cora, Giulia and Tommy.

Page 137: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

132

A special thanks to Vicky for our 14-years long friendship that has been a constant and essential

element in my life from Alassio to Milano and that grows stronger every year that passes. A second

special thanks to Fede, for being the best friend (and roommate) anyone could hope for. Thanks to

Ivan for always giving me a place in the sun and to Marco for not letting me be the only nerd

engineer at the beach. Thanks also to Silvia and Jessica for being there since I can remember.

Last, but not least, thanks to my mother, for the endless support and love she has always given me

throughout all my life and for never letting me lack anything, to my father, my grandmother, my

cousin, my aunt, my brothers and sister.

Page 138: Mechanisms of cellular photostimulation in hybrid interfaces … · 2016-02-20 · i Abstract Hybrid interfaces between organic semiconductors and living tissues represent a new tool

133

List of publications

1. N. Martino, P. Feyen, M. Porro, C. Bossio, E. Zucchetti, D. Ghezzi, F. Benfenati, G. Lanzani,

M.R. Antognazza, Phototermal cellular stimulation in functional bio-polymer interfaces. Sci.

Rep. 5, 8911 (2015).

2. M.R. Antognazza, N. Martino, D. Ghezzi, P. Feyen, E. Colombo, D. Endeman, F. Benfenati, G.

Lanzani, Shedding Light on Living Cells. Adv. Mater. (2014) DOI: 10.1002/adma.201403513

3. N. Martino, D. Fazzi, C. Sciascia, A. Luzio, M.R. Antognazza, M. Caironi, Mapping

Orientational Order of Charge-Probed Domains in a Semiconducting Polymer. ACS Nano 8,

5968 (2014)

4. V. Benfenati, N. Martino, M.R. Antognazza, A. Pistone, S. Toffanin, S. Ferroni, G. Lanzani, M.

Muccini, Photostimulation of Whole‐Cell Conductance in Primary Rat Neocortical Astrocytes

Mediated by Organic Semiconducting Thin Films. Adv. Health. Mater. 3, 392 (2014)

5. D. Endeman, P. Feyen, D. Ghezzi, M.R. Antognazza, N. Martino, E. Colombo, G. Lanzani, F.

Benfenati, The Use of Light-Sensitive Organic Semiconductors to Manipulate Neuronal Activity

in “Novel Approaches for Single Molecule Activation and Detection”. Springer Berlin

Heidelberg (2014).

6. G. Grancini, M. De Bastiani, N. Martino, D. Fazzi, H.-J. Egelhaaf, T. Sauermann, M.R.

Antognazza, G. Lanzani, M. Caironi, L. Franco, A. Petrozza, The critical role of interfacial

dynamics in the stability of organic photovoltaic devices. Phys. Chem. Chem. Phys. 16, 8294

(2014).

7. D. Ghezzi, M.R. Antognazza, R. Maccarone, S. Bellani, E. Lanzarini, N. Martino, M. Mete, G.

Pertile, S. Bisti, G. Lanzani, F. Benfenati, A polymer optoelectronic interface restores light

sensitivity in blind rat retinas. Nature Photon. 7, 400 (2013).

8. N. Martino, D. Ghezzi, F. Benfenati, G. Lanzani, M.R. Antognazza, Organic Semiconductors

for Artificial Vision. J. Mater. Chem. B 1, 3768 (2013).

9. G. Grancini, N. Martino, M. Bianchi, L.G. Rizzi, V. Russo, A. Li Bassi, C.S. Casari, A.

Petrozza, R. Sordan, G. Lanzani, Ultrafast spectroscopic imaging of exfoliated graphene. Phys.

Status Solidi B 249, 2497 (2012).

10. G. Grancini, N. Martino, M.R. Antognazza, M. Celebrano, H.-J. Egelhaaf, G. Lanzani,

Influence of blend composition on ultrafast charge generation and recombination dynamics in

low band gap polymer-based organic photovoltaics. J. Phys. Chem. C 116, 9838 (2012).

11. C. Sciascia, N. Martino, T. Schuettfort, B. Watts, G. Grancini, M.R. Antognazza, M.

Zavelani‐Rossi, C.R. McNeill, M. Caironi, Sub‐Micrometer Charge Modulation Microscopy of

a High Mobility Polymeric n‐Channel Field‐Effect Transistor. Adv. Mater. 23, 5086 (2011).