the use of microreactors/optofluidics to study

55
The use of microreactors and optofluidic waveguides to study photochemical reactions MSc Chemistry - Literature study By Amber Jaspars

Upload: others

Post on 03-Dec-2021

9 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The use of microreactors/optofluidics to study

The use of microreactors and

optofluidic waveguides to

study photochemical reactions

MSc Chemistry - Literature study

By

Amber Jaspars

Page 2: The use of microreactors/optofluidics to study

1

The use of microreactors and optofluidic

waveguides to study photochemical reactions

Literature project - MSc Chemistry – Analytical track

By

Amber Jaspars

UvA/VU

Supervised by:

Daily supervisor: I. Groeneveld

First examiner: F. Ariese

Second examiner: G. Somsen

April 2020

Page 3: The use of microreactors/optofluidics to study

2

Abstract

In this literature review, the use of different microreactors and optical waveguides for

photochemical reactions is discussed. Photochemical reactions can be carried out in a batch

reactor, but these batch reactors are not homogeneous, are hard to irradiate and take a

long time before it is fully reacted. With a microreactor, these problems can be solved.

Because of the small volume, mixing is faster and more efficient and the sample is more

easily irradiated, creating a faster reaction. The microreactor can be irradiated sideways or

along the length. For sideways irradiation, mostly chips with microchannels are covered with

a glass plated. A light source above the chip irradiates the sample, the light penetrates

through most of the sample and the mixture is easily irradiated.

The microreactor can also be irradiated along the length of the reactor, using a liquid core

waveguide (LCW). The LCW is like a tube made from a reflective cladding material, the light

is reflected within and guided through the LCW. The light can be guided either by total

internal reflection. The light is guided if the cladding material has a smaller refractive index

than the liquid core, meaning the cladding material should be carefully chosen. Another

principle of light guiding is based on interference based waves, multiple layers of material

are used as cladding, creating multiple reflections. This principle can be based on photonic

bandgaps or on the Fabry–Pérot reflector.

Page 4: The use of microreactors/optofluidics to study

3

Table of contents

1 Introduction ............................................................................................................................. 4

2 Photochemical reactions and photodegradation .............................................................. 5

3 Microreactors and waveguides ............................................................................................. 7

3.1 Micro photoreactors ....................................................................................................... 8

3.2 Optofluidic waveguides ................................................................................................. 9

4 Total internal reflection based waveguides....................................................................... 12

4.1 Teflon AF ........................................................................................................................ 14

4.1.1 Fabrication ............................................................................................................. 15

4.1.2 Detection and Reaction use ................................................................................. 18

4.2 Aerogel microchannels ................................................................................................ 21

4.2.1 Fabrication ............................................................................................................. 22

4.2.2 Detection and Reaction use ................................................................................. 24

5 Interference based waveguides .......................................................................................... 28

5.1 Bragg fibres ................................................................................................................... 29

5.1.1 Fabrication ............................................................................................................. 29

5.1.2 Detection and Reaction use ................................................................................. 30

5.2 Photonic crystal fibres (PCF) ........................................................................................ 31

5.2.1 Fabrication ............................................................................................................. 31

5.2.2 Detection use ......................................................................................................... 34

5.2.3 reaction use ............................................................................................................ 35

5.3 Anti-resonant reflecting optical waveguides (ARROW) ............................................ 38

5.3.1 Fabrication ............................................................................................................. 39

5.3.2 Detection and reaction use .................................................................................. 41

6 Conclusions and outlook ..................................................................................................... 42

7 The List of References .......................................................................................................... 44

Page 5: The use of microreactors/optofluidics to study

4

1 Introduction

There is a growing interest in microreactors since the beginning of the 21st century. They

consist of a solid support with channels of 10–1000 µm in width and depth [1][2]. The

microreactors have several advantages over bulk or batch reactors, due to the small

volumes and the narrow channels. For photochemical reaction, a light source for bulk

reactors can be put on the outside of the reactor, but it would only irradiate the compounds

near the edges, irradiating a small surface in a large volume. When using microreactors for

photochemical reactions the main advantage is the surface-to-volume ratio [1][2].

For photochemical reactions in a microreactor, the sample can be irradiated in two ways.

The first way is irradiating sideways by using a micro photoreactor. The micro photoreactor

consists of channels in a solid plate covered with a transparent material. The light comes

from above and due to the narrow channels the light penetrates through most of the reactor

and the reaction mixture [1][3][4]. The other way of irradiating the microreactor is along the

length by using a liquid core waveguide. A liquid core waveguide is a tube made of a

reflective cladding material that reflects the light from within [1][5]. The maximized surface-

to-volume ratio in the micro photoreactor enables highly efficient photochemistry, with an

improved irradiation and better light penetration through the reaction volume [6]. Because

of the better surface-to-volume, there is more light in a smaller volume, the molecules

absorb the light easier and the reaction goes faster [7].

This study is for the TooCOLD project, the TooCOLD project has the goal to improve

photodegradation research. They want to create an online two-dimensional liquid

chromatography system, which is combined by a photodegradation microreactor. The goal

of this study is to find what kind of micro photoreactors have been developed and which

optofluidic devices or waveguides could or have been used for photochemical reactions.

The questions would be if the found microreactors and waveguides could be coupled to a

HPLC or other liquid separation methods and what the important parameters would be. This

study will not be focused on photodegradation, but will take a broader view and look at

photochemical reactions.

In the first chapter, the introduction with the goal of the research is given. The second

chapter will give more information about the photochemical reactions, it will discuss the

parameters that need to be considered. In chapter three more information will be given

about the microreactors and waveguides. The general overview of the differences will be

discussed together with the important parameters. Different waveguides with their principle,

fabrication and the detection and reaction used will be discussed in chapter four and five.

This study will be concluded and a future outlook will be given in chapter six.

Page 6: The use of microreactors/optofluidics to study

5

2 Photochemical reactions and photodegradation

There is a growing interest in photochemical reaction in many areas, including medicine

and chemical synthesis [8]. The photochemical reactions differ from the usually used

thermal reactions. The photochemical reaction is activated mainly by the absorption of light,

while the thermal reaction is mainly activated by the use of heat [9]. With the absorption of

a photon, the molecule reaches an electronically excited state. In the excited state, the

molecule possesses quite a high internal energy, the electronic structure and the nuclear

configuration of the molecule will be different from the ground state [4]. These excited

molecules can fall apart, change to new structures, combine with each other or other

molecules, or transfer electrons, hydrogen atoms or protons. These photochemical

reactions may result in products that cannot be formed by thermally driven reactions in the

ground state. This allows photoreactions at very low temperatures, near-zero Kelvin [9].

There are many different photochemical reactions, mostly they are categorised based on

the required conditions (homogeneous or heterogeneous) and not on their reaction type

[1]. In homogeneous reactions all the reagents are in the same phase or solution [1], [3], [10],

[11]. For homogeneous catalytic reactions, the catalyst is dissolved in the reaction solvent.

Heterogeneous gas-liquid reactions are between two phases and acquire a gas supply,

commonly Photooxygenations or photochlorinations [3][5][6]. However, gasses are difficult

to handle in batch conditions. Filling a batch reactor with gas in unpressurized systems

results in serious mass-transfer limitations. Although, in a pressurized system it has serious

safety concerns when using toxic or reactive gases. A microreactor avoids these problems

by using high amounts of gas in each part of the reactor, but it can also be easily pressurized

in a safe manner [5]. For heterogeneous solid-liquid reactions, same as for homogeneous

reactions where solid intermediates or products are formed, the microreactor is a challenge.

The solid particles can interfere with the light, but can also clog in the small channels of the

microreactor [5]. When using a heterogeneous catalyst, the catalyst can be immobilised

within the reaction channel. Due to the large surface-to-volume ratio, there is a maximised

interaction between the sample, the catalysts and the photons [1], [11]–[13]

The TooCOLD project is mainly focused on the photodegradation of products. The goal is

to create an online two-dimensional liquid chromatography system, which is combined by

a photodegradation microreactor. Photodegradation is beneficial for water and wastewater

treatment, however, it is unwanted in many areas, for example, food products, packaging

materials or cultural heritage like paintings [14]. Photodegradation is a photochemical

reaction, where the exposure to light may cause significant degradation of many materials

[15]. It may happen direct or indirect, by direct photodegradation the chemical absorbs the

light itself. With indirect photodegradation, the light is absorbed by photosensitizers and

the excited photosensitizers generate photoreactants, which can react with the compounds

[16]. Most commonly the photosensitizer (like H2O2 or O3) generates singlet oxygen, the

singlet oxygen reacts with for example food components, photooxidative degradation. This

can influence flavour quality, nutritional quality and toxicity of food products [17]–[19]. The

degradation rate depends on different parameters, like pH, temperature, concentration and

purity of the chemicals and the concentration of the photosensitizer [20][21].

Page 7: The use of microreactors/optofluidics to study

6

The flow is another important parameter in the photochemical reactions, especially when

minimizing them. The reactions can be performed without a flow. The reactor is filled with

the reaction mixture before it is irradiated, after irradiation the product can be collected. A

detector could measure the progress of the reaction. When performed under a continuous-

flow the irradiation time is directly proportional to the flow rate of the reactor. This makes

the optimization faster, because the irradiation time can be easily changed [1][5][11]. A

detector could be installed at the outlet of the reactor, it would measure the speed of finding

the equilibrium of the reaction rate. When the equilibrium is found, the detector would see

if the equilibrium is unstable. A UV-VIS or a fluorescence detector could be used inline and

measure the absorbance of the used light. Online many other detectors could be used for

the same purpose, McQuitty et al. [22] used an online mass spectrometer to identify the

reaction products.

Another question to ask when using a flow, is what kind of flow should be used. A

hydrodynamic flow (shown in Figure 2.1 A) occurs because of a pressure difference

between the inlet and outlet of the reactor channel. The HPLC uses a hydrodynamic flow, so

the coupling of the two would be easy. The advantages of the hydrodynamic flow are that it

may be used with any liquid and with devices constructed of any material. However, the

hydrodynamic flow has a parabolic flow profile, the compounds in the middle of the flow

will go faster than the particles close to the wall, these particles stay longer in the reactor

and create a decrease in yield and selectivity. Another disadvantage from the hydrodynamic

flow is the pressure involved, when the channels are too small the pressure will be too high

[2]. An electrokinetic flow (shown in Figure 2.1 B) could also be used, it occurs due to a

direct movement of ions in solution toward the electrode of opposite charge, as in capillary

electrophoresis. The velocity profile of the electrokinetic flow is nearly flat across the

channel, leading to a more homogeneous mixture and a higher yield compared to a

hydrodynamic flow. Unfortunately, the electrokinetic flow needs ions, which can interfere

with the reaction. Because of the ions, the electrokinetic flow is also limited with polar

solvents and device materials that can handle the charges in the system [2]. The HPLC uses

a hydrodynamic flow, so the coupling of the HPLC with a microreactor would be easy

compared to the electrokinetic flow.

Figure 2.1 (A) hydrodynamic flow and (B) electrokinetic flow [2]

Page 8: The use of microreactors/optofluidics to study

7

3 Microreactors and waveguides

Organic synthesis have mostly been proceeded in the round-bottom flask, with these

traditional synthetic methods almost any organic compound can be fabricated. However,

microreactors have received a lot of attention since the beginning of the 21st century.

Microreactors, also known as microstructured reactors or micro channelled reactors, consist

of a solid support with channels of 10–1000 µm in width and depth [1][2]. The microreactors

have several advantages over these batch reactors, due to the narrow channels and the

small volumes. The mixing in the microreactors is very homogenous and can happen really

fast, using a static mixer in the microreactor makes the mixing happen in milliseconds. The

stirring in classical reactors, for example, a round-bottom flask, is limited by its

inhomogeneities due to its stirring mechanism. Due to the small size, heat is effectively

dissipated. This prevents the hotspots typically found in batch reactors, contributing in a

better selectivity and better control over the heat transfer. The small volumes decrease

waste and increase safety. Because of the small volumes used in the microreactors, it allows

the safe use of highly toxic or explosive reactants [2][13]. However, it is important that the

device is minimized in adsorption to the wall, crystallization and clogging or it can ruin the

experiments due to loss, pressure and scattering [3].

When using microreactors for photochemical reactions the large surface-to-volume ratio is

a big advantage. In bulk reactors, the light can shine on the outside of the reactor, but you

would only irradiate the compounds on the outside of the sample reactor, irradiating a small

surface in a high volume. Another way is to place a lamp in the sample solution, but still only

a part of the sample is irradiated, so again a small surface is irradiated in a high volume.

When using a microreactor for the photochemical reactions, the sample can be irradiated

in two ways. Irradiating sideways by using a micro photoreactor (see Figure 3.1 a) or by

irradiating along the length using a waveguide for instance (see Figure 3.1 b) [1][5]. The

maximized surface-to-volume ratio for the micro photoreactor enables highly efficient

photochemistry, with an improved irradiation and better light penetration through the

reaction volume [6]. Because there is a larger surface-to-volume ratio there is more light in

a smaller volume, the molecules absorb the light easier and the reaction goes faster [7].

Figure 3.1 microreactors set up (a) irradiated sideways (micro photoreactor) [12] (b) irradiated along the length (optical

waveguide) [6]

a b

Page 9: The use of microreactors/optofluidics to study

8

3.1 Micro photoreactors

Most micro photoreactors (shown in Figure 3.1 a) consists of channels in a chip covered with

a transparent material. The sample is irradiated from a light source outside the channel,

when the channel thickness is small it allows a high surface-to-volume ration. The light

penetrates through most of the reactor and the reaction mixture can be easily irradiated

[1][3][4]. For photochemical microreactors, many examples already exist in the literature.

Many describe different reactions or they published a review comparing the reactions with

the advantages of the microstructured photoreactors. Coyle and Oelgemöller [1] wrote a

review about the different reactions with all the advantages on microreactors and

Sambiagio and Noël [5] wrote one February 2020 about different micro photoreactors with

a continuous-flow.

Micro photoreactors allow off-chip detection of the reaction, for example with a

spectrometer at the outlet of the chip [1][3]. The detection could also be done on-chip,

having an immediate detection before the outlet of the chip. This inline detector could also

be on different spots in the channel, when using a continuous flow the reaction process

could be measured [3]. However, these channels are really small, with really small

pathlengths, making it hard to have a suitable inline detector. Instead of a detector, the

micro photoreactor could also be coupled to a HPLC, but HPLC analysis can be relatively

slow compared to most photochemical reactions [3]. A stop flow, loops or a really fast LC

separation are needed to keep up with the flow of the micro photoreactor.

For homogenous reactions, where all reagents are dissolved in the liquid, many examples

exist. With different junctions multiple liquids can be mixed before or in the microreactor

itself. Commercially the serpentine channel type reactors are available but many researcher

custom build their homogenous photoreactors. This gains them the flexibility for the length

and depth, as well as to choose the materials for the solid substrate and the cover glass

[1][3][10][11]. Heterogeneous gas-liquid reactions require a gas supply, where some kind

of membrane could be used. Commercially is the falling film type reactor available, but they

can also be custom build reactors [1][10][11].

The concentrations in the reaction mixture are also important, because in a reaction a high

yield and a high amount of product are wanted. In the microreactors only small amount of

starting materials are used, so only a small amount of reaction products can be produced.

When increasing the concentration also the light absorption increases in the microreactor,

due to the law of Lambert-Beer, see equation (1). This can affect the yield, because there

are too many compounds and not all are excited for the photochemical reaction.

Page 10: The use of microreactors/optofluidics to study

9

3.2 Optofluidic waveguides

In optofluidic microreactors or liquid core waveguides (LCW), the microreactor is a tube

where the light irradiates along the length (see Figure 3.1 b). A reflective cladding material

is used to reflect the light within the LCW and with the small volumes used, it creates a large

surface-to-volume ratio. The light guiding can be achieved by two different principles [23].

Either by total internal reflection (TIR), where the light is guided if the cladding material has

a smaller refractive index than the liquid core, meaning the cladding material should be

carefully chosen. This will be discussed further in chapter 4 [24]. The other principle is based

on interference based waves, where multiple layers of material are used as cladding,

creating multiple reflections. This is further discussed in chapter 0 [25].

For homogenous reactions, where all reagents are dissolved in the liquid, both TIR and

interference based waveguides work, because only one phase (the liquid phase) enters the

tube. Multiple examples are already made and discussed in the next chapters. Most of them

are commercially available, but the fabrication procedure to custom build the different

waveguide is also given. For heterogeneous gas-liquid reactions a gas supply is needed,

the cladding should not only be the reflecting material but also be gas permeable. This is

easier with TIR based waveguide and might be difficult for the interference based

waveguides because of the multiple cladding layers. Homogeneous catalytic reactions can

proceed in both waveguide types, the catalyst is solved in the reaction mixture so it would

not interfere with the light. For heterogeneous catalytic reactions, the interference based

waveguides might have the advantage, because of the multiple layers starting with a

catalytic layer might not worsen the loss that much. However, using a catalytic layer would

change the refractive index of the cladding material and make total internal reflection way

less possible.

The Lambert-Beers law calculates the absorbance (A), where ɛ is the molar extinction

coefficient, L the optical pathlength, c the concentration and I0 and It the intensity of the light

in the absence and the presence of the analyte.

𝐴 = 𝑙𝑜𝑔 (𝐼0

𝐼𝑡) = 𝜀𝐿𝑐 (1)

The concentration is an important parameter to think about, because with a high

concentration and a high pathlength gives a high absorption. In the waveguide only small

amount of starting materials are used, so only a small amount of reaction products can be

produced. To increase the amount of reaction product the starting concentration needs to

be increased, but this increases the absorbance. When using a cuvette (L = 1 cm), the

wanted absorption is normally found between 0,2 and 1. When using the same

concentration and molar extinction coefficient but you increase the length to a meter (L =

100 cm), the absorption increases by a 100 times and cannot be normally measured. With

a high concentration and a long pathlength, the absorbance is too high. In this way, the light

is already absorbed before it even reached the end of the waveguide and the yield is

decreased. Therefore, small concentrations should be used [26].

Page 11: The use of microreactors/optofluidics to study

10

To make sure the light penetrates through the entire waveguide the optical loss (dB/m) must

be minimized. Typical the losses are created by Rayleigh scattering, OH absorption (~1310

and ~1550 nm) and the imperfection of the waveguide [27]. The optical losses are mostly

given in attenuation (in dB) per meter. When using the waveguide as a microreactor, the

waveguide would not be longer than a few meters and a loss around 2 dB/m is acceptable.

Less would always be better but is not necessary. Having a loss of 1 dB/cm would be a big

loss if you have a waveguide of one meter long. However, when using the waveguide in

different lengths or for different purposes, the acceptable losses could be different.

The waveguides have already been used for increasing the optical pathlength of a

spectrophotometer. According to Lambert-Beer (equation (1)) the sensitivity of

spectrophotometry can be enhanced by increasing optical pathlengths. So at really low

concentrations, increasing the length would still give a measurable absorbance [28][29].

The absorbance can also be measured during the photoreactions, as an inline detection.

The absorbed light could be measured at the exit of the waveguide. With the absorption

the stability of the reaction flow could be measured and the flow could be slowed down to

increase the yield. When performing the reactions without a flow the reaction rate could be

measured, because the sample is not moving the absorbance tells something about the

entire reactor at that time [11]. The waveguides could also have online detection, where a

detector is coupled at the end of the waveguide, this doesn’t have to be a spectrometer but

can also be an MS [22]. Coiling the waveguide should for absorbance spectroscopy not be

a problem, only if the light is guided through the waveguide. However, when the waveguide

coils too much the angle of incidence can be too small and the light is not reflected in the

waveguided and more losses can be found.

When increasing the pathlength, the chance to absorb light increases. When increasing the

pathlength in fluorescence, you increase the absorbed excitation light, increasing the

emitted fluorescence, which increases the sensitivity [30]. Fujiwara and Ito [31][32] were the

first to execute fluorescence spectrometry in an LCW. They illuminated the waveguide

axially, see Figure 3.2 a. The excitation light that does not excite a molecule, survives to the

end of the waveguide and increases the fluorescence intensity signal. In the detector, the

fluorescence signal needs to be separated from the excitation light to obtain the correct

values. However, it will be hard to separate them when there is an overlap between the

excitation and the emission wavelengths [30][33][34]. Another way to irradiated the LCW is

by transverse illumination through the waveguide side-wall, shown in Figure 3.2 b. Less light

goes into the waveguide, but the light that is not absorbed by the components passes

through the tube, so almost no excitation light is found in the detector [30][33][34].

Dasgupta et al. [34] used a transverse illumination in an LCW for longer pathlength

fluorescence spectrometry. They also changed the geometry by coiling the waveguide, but

this causes a slight loss from the light source. It also increases the light that enters transverse,

but propagate through the axial mode. However, the transverse illumination can only be

used for TIR based waveguides. Due to the critical angle and a small angle of incidence the

light is not reflected and goes in and out of the waveguide, but with interference based

waveguides this is not always the case.

Page 12: The use of microreactors/optofluidics to study

11

Figure 3.2 (a) axial and (b) transverse illuminated fluorescence in a liquid core waveguide [31][34]

The LCW can also be used to improve the sensitivity of Raman spectroscopy, normally the

sensitivity is really low because of the rare inelastic scattering. After the laser excites the

molecule a majority is Rayleigh scattering, which has no exchange of energy. A small part is

inelastically scattered and has a change in energy, the difference in energy between the

excitation and the scattered photons corresponds to a vibrational mode of the molecule,

called Raman scatter [24][35]. When increasing the pathlength there will be more scattering,

more Rayleigh scattering but also more inelastic scattering, this is also called fibre enhanced

Raman spectroscopy [36][37]. This significantly improves the Raman signal [35], [38]–[41].

In Raman spectroscopy the illumination can also be done axially, Altkorn et al. [42] used 6

different axial geometries for an LCW. Consisting of forward and/or backward scattering

with or without the used of a long pass filter or a mirror. The Raman intensity was measured

over the cell length, giving the losses for the different geometries. Using a long bandpass

filter or a mirror increases the intensity, the highest intensity was found using combined

forward and backward scattering with a long pass filter. The illumination in Raman

spectroscopy can also be done transverse, Holtz et al. [43] have successfully used an LCW

for Raman spectroscopy with transverse illumination. Next to stokes and anti-stokes Raman,

the LCW can also be used for resonance Raman spectroscopy [35][38].

a b

Page 13: The use of microreactors/optofluidics to study

12

4 Total internal reflection based waveguides

The first waveguides to be discussed are the total internal reflection (TIR) based waveguides.

Snell’s law or the law of refraction, shown in Figure 4.1, provides the fundamental principle

for a total internal reflection based optical waveguide [44]. The formula describes the

relationship between the angle (𝜃) and the refractive index (n).

sin(θ2)

sin(θ1)=

n 1

n2 (2)

Figure 4.1 Snell’s law

When light reaches another medium two things can happen, either the light penetrates into

the medium, it is (partly) reflected. When the second medium has a bigger refractive index

than the first one (n1 < n2) or when the angle (θ1) is smaller than the critical angle, the light

penetrates into the second medium. The light undergoes total internal reflection (TIR) when

the second medium has a smaller refractive index (n1 > n2) and has an angle (θ1) equal to

or bigger than the critical angle [24]. The critical angle can be calculated by (sin(θ2) = 1),:

θcritical = sin−1 (n2 (cladding)

n1 (core)) (3)

For TIR-based waveguides, the cladding material needs to have a smaller refractive index

than the refractive index of the liquid core [45]. The bigger the difference between n1 and

n2, the smaller the critical angle will be. When the critical angle is small, the region for θ1 is

bigger and more light will be reflected. Figure 4.2 shows a schematic side view of a liquid

core waveguide, where the cladding material has a lower refractive index. The light follows

the red arrows and is reflected at the wall, with an angle less than the critical angle so the

light does not enter in the cladding material [45].

Figure 4.2 Schematic drawing of a liquid core waveguide [46]

Page 14: The use of microreactors/optofluidics to study

13

Many solvents have low refractive indexes, most aqueous solutions have a refractive index

around 1,33, so the cladding material needs to have a really low index of refraction to realize

TIR. For common solvents like methanol and acetonitrile their refractive index are also

relatively low, 1,33 and 1,34. Solvents with higher refractive indexes are for example

chloroform (n = 1,45) or toluene (n = 1,49), compared to glass (n = 1,46) is this still low [47].

In this chapter two materials are discussed which have a refractive index lower than water,

Teflon AF (RI of 1,29-1,33) and aerogels (RI = 1, as air ).

Page 15: The use of microreactors/optofluidics to study

14

4.1 Teflon AF

In 1989 the fluoroplastic material was announced by DuPont, a material with the

requirements for a TIR based LCW [48]. Teflon AF [chemical name: fluorinated (ethylenic-

cyclo oxyaliphatic substituted ethylenic) copolymer] is a family of amorphous copolymer of

2,2-bistrifluoromethyl-4,5-difluoro-1,3-dioxol (PDD) and tetrafluoroethylene (TFE), the

structure is shown in Figure 4.3 structure of Teflon AFFigure 4.3 [48][49]. The difference

between Teflon AF varieties (for example AF1600 and AF2400) is in the relative amount of

the dioxol monomer (TFE: PDD) in the basic polymer chain. The labelling of Teflon AF refers

to the glass transition temperature. As the relative concentration of PDD increases the glass

transition temperature increases, so the transition temperature of Teflon AF1600 is 160⁰C

and for AF2400 the transition temperature is 240⁰C [48][49].

Figure 4.3 structure of Teflon AF [49]

Teflon AF has many advantages, next to outstanding chemical, thermal, and surface

properties it also has unique electrical and solubility characteristics. But the major

advantage of Teflon AF is their low refractive index, making them very suitable as low-index

claddings for waveguide applications [49].

Figure 4.4 shows the index of refraction for three different grades of Teflon AF (AF1300,

AF1601 and AF2400) by different wavelengths. It shows that Teflon AF2400 has a lower

refractive index, and AF1300 the highest, so the refractive index increases when the PDD

concentration decreases. After 200 nm the index values are fairly stable (1,29-1,33), it would

be recommended to use wavelengths longer than 200 nm [49].

Figure 4.4 Refraction index determined for three Teflon AF grades at different wavelengths [49]

Page 16: The use of microreactors/optofluidics to study

15

4.1.1 Fabrication

Teflon AF waveguides can be fabricated in multiple ways: a capillary made of Teflon or a

capillary coated internally or externally with Teflon AF. The first one to be discussed is a

liquid core waveguide composed solely of Teflon AF (often called type I, shown in Figure

4.5) [24] [50]. The light travelling through the liquid core can interact with the Teflon AF walls.

Most of the light will be reflected back through the liquid core, but some can be transmitted

in the Teflon AF. At the Teflon AF/air interface the transmitted light can be transmitted to

the air and lost or reflected back in the Teflon AF and transmitted in the liquid core

waveguide [50]. The waveguides are formed by melting Teflon pellets into a preform and

drawing capillaries in a similar way glass capillaries are made [25].

Figure 4.5 schematic drawing of a type I Teflon AF waveguide

Altkorn et al. [51] used Teflon AF 2400 capillary tubing. The waveguides were drawn from

preform, these preforms were really expensive at that time ($10 000/kg). The LCW had an

outer diameter of 525 µm and an inner diameter of 250 µm with a variation of 8%. The

attenuation at 632,8 nm was tested over different waveguide lengths, see Figure 4.6 (a). The

loss can be found in the slope of the line, the solid line shows the least-squares fit and

indicates a loss of 1,9 dB/m from 0-0,45 meters and 1,4 dB/m from 0,55-0,75 meter.

Between 0,45 and 0,55 meter the loss was high (5,9 dB/m), but this was probably caused by

scattering and imperfection. They calculated with two different waveguide lengths (0,05 m

and 0,95 m), the losses over a wavelength range of 400 and 800nm, see Figure 4.6 (b) (solid

is water and dashed is methanol). The loss is lowest with wavelengths around 600 nm for

water (3dB/m) and methanol (2dB/m).

Figure 4.6 (a) Attenuation versus length in a Type I Teflon AF 2400 capillary filled with methanol, solid lines show the least-

square fit to the data (b) loss spectra of water (solid) and methanol (dashed) in a Teflon AF 2400 capillary [51]

a b

Page 17: The use of microreactors/optofluidics to study

16

One of the advantages of Teflon AF is its flexibility. It can be easily coiled, which makes it

easier to use longer lengths without using a lot of space [24]. Another advantage is the

porous structure of Teflon AF, which makes it very gas permeable. This makes it possible to

use the Teflon AF waveguide as a gas sensor [24]. An example is given by Dasgupta et al.

[28], who used two different Teflon AF devices for gas sensing. Their first device was made

from commercially available Teflon AF 2400 tubes. The waveguides were surrounded by

glass to keep the gas inside the device. The lengths of Teflon ranged between 15 and 30

cm. For the second device they bought a 24 cm long silica capillary externally coated with

Teflon AF. Again the waveguide was surrounded by glass, but they removed the silica by

pumping an aqueous solution of 20% HF through the capillary. The two devices were used

as sensors for CO2 in water samples, with an increase of CO2 there was a decrease in pH

(using NaHCO3). With the use of an indicator (phenol red), the colour in the liquid core

changed and a difference in absorption could be measured. Because the colour change of

the indicator is extremely fast it did not interfere with the response. Wang et al. [52] also

used a Teflon AF 2400 waveguide (inner diameter of 550 µm, an outer diameter of 625 µm

and 12-21 cm long) to measure CO2, but they used gas samples. The amount of CO2 caused

a decrease or increase of the pH (using Na2CO3) of the filling medium that is reflected in

turn by the optical behaviour of the indicator (phenol red), creating more or less absorption

in the waveguide. Milani and Dasgupta [53] used the same setup as Dasgupta et al. [28],

but used the LCW to measure nitrous acid and nitrogen dioxide.

Teflon AF can also be coated on the inside of glass, silica or PDMS. Creating the same effect

as Teflon AF alone, but surrounded by an extra protection layer of tubing. The light

travelling through the liquid core can interact with the Teflon AF walls, the light will be

reflected back through the liquid core, but some of the light can be transmitted in the Teflon

AF. By interaction from Teflon AF to waveguide there is no reflection, because the refractive

index of the extra cladding is bigger than the refractive index of Teflon AF. The light will be

reflected in the cladding/air interface and travel back to the liquid core. A way to coat the

inside of the waveguide with Teflon AF is by using a wafer with channels, as used in the

micro photoreactors. Manor et al. [30] used soda-lime glass and Pyrex, Datta et al. [54] used

silicon and Wu and Gong. [55] and Cho et al. [56] used polydimethylsiloxane (PDMS). But

the adhesion Teflon AF is difficult. Manor et al. [30] used a monolayer fluorocarbon-based

solvent which was spun coated on the wafer. It provides low surface energy to oxide surfaces,

therefore, allowing better adhesion of Teflon AF. Datta et al. [54] used thin fluorocarbon

films formed by plasma-enhanced-chemical vapour deposition using inductively-coupled

high density plasma methods.

Dress et al. [57] tested the influence of the layer thickness on the optical properties of the

LCW. For different layer thicknesses they calculated the reflectivity of Teflon AF 2400 as a

function of the angle. The reflectivity is defined as the ratio of the reflected to the incident

intensity, or also known as the transmission. Figure 4.7 summarizes all data in a three-

dimensional plot. The critical angle is measured at 75,41⁰ with respect to the normal. A

detailed calculation demands a minimum value of the Teflon AF2400 coating of about 5 mm

to obtain total internal reflection. To get a smooth coating Datta et al. [54] recommends to

Page 18: The use of microreactors/optofluidics to study

17

bake the Teflon AF above its transition temperature, this softens the polymer and produces

a smooth surface.

Figure 4.7 Three-dimensional plot of the reflectivity spectrum of the layer system for different thicknesses of the Teflon AF

2400 inner coating in the LCW [57]

Another type of Teflon AF waveguide is a silica or quartz capillary coated on the outside

with Teflon AF to create the LCW effect, this is often called a type II LCW, Figure 4.8 [24].

The waveguide made of glass or quartz is surrounded by Teflon AF. Because the refractive

index of the cladding material is bigger than the refractive index of the liquid core the total

internal reflection does not happen at the liquid/glass interface. For this reason, the light

travels both in the liquid core and in the glass itself. The light is effectively trapped in the

waveguide wall, because its refractive index is much larger than the ones from the liquid

and the Teflon AF. The light that does exit the waveguide wall will propagate within the

liquid core and again enter the tube wall. The light is trapped in the waveguide walls but it

also involves a large number of traverses of the water/quartz interface [50].

Figure 4.8 schematic drawing of an externally coated Teflon AF waveguide

Altkorn et al. [58] used a fused-silica tubing (530 µm inner diameter and 630 µm outer

diameter) coated externally with a thin (~15 µm) layer of Teflon AF, but they do not describe

which type of Teflon AF. The light is coupled into the waveguide by an optical fibre. The

attenuation at 632,8nm was tested over different waveguide lengths, see Figure 4.9 (a).

Page 19: The use of microreactors/optofluidics to study

18

Inherent waveguide loss, given by the slopes of these lines, is approximately 1,6 dB/m for

water. They also tested the losses between 400 and 800 nm, see Figure 4.9 (b), at two

different waveguide lengths (0,05 m and 0,95 m). The loss is lowest with wavelengths

shorter than 700 nm and is lowest around 600 nm being 1 dB/m.

Figure 4.9 (a) Attenation versus length in a Type II silica capillary coated with Teflon AF 2400 at 632,8 nm filled with water

(b) loss versus wavelength for water filled waveguide [58]

When comparing the type I and the type II, there are not many differences, the losses for

both are almost the same, as are the used diameter and length of the LCW. Probably the

main reason why the type I Teflon AF waveguide is more used it the travel path of the light.

In the type II Teflon AF waveguides the light propagates mostly through the glass walls, but

in type I the light mostly propagates through the core. When the light only needs to travel

to the other side of the waveguide this does not matter, but when the light needs to interact

with the liquid in the core. Another advantage of the type I Teflon AF waveguides is the gas

permeability, which makes it possible to use the LCW for heterogeneous gas-liquid

reactions. These require a gas supply, the type I Teflon AF waveguide can homogeneously

add gas to the reaction without creating bubbles. The optical losses in both waveguides are

acceptable when using them for photochemical reactions, because it is not to be expected

to use waveguides longer than maximal a few meters.

4.1.2 Detection and Reaction use

As said in chapter 0 the LCW can be used as a longer pathlength in absorption spectroscopy.

The Teflon AF waveguide is used a lot for trace analysis, for example in water and marine

chemistry [24][59][60]. Many other general applications have also been described, for

example, Cheng et al. [29] used the LCW to create a longer pathlength for the determination

of organophosphorus pesticides in vegetables and fruits. The LCW can also improve

fluorescence detection, it improves the light–fluorophore interaction and fluorescence

collection efficiency and improves the detection limit [61]. The LCW significantly improves

a b

Page 20: The use of microreactors/optofluidics to study

19

the Raman signal by the propagation of Raman scattered light along the length of the

waveguide [37]–[41].

The Teflon AF liquid core waveguide can also be used as a microreactor. Its optical

properties are good for photoreactions and the gas permeability of type I Teflon LCW

makes it perfect for heterogeneous gas-liquid reactions. However, the reactions executed

in the Teflon AF LCW are not photochemical reactions. The reactions make use of the gas-

permeable Teflon AF type I and some use the reflective properties as detection of the

reaction. The gas permeability of type I Teflon LCW, is a good advantage because of the

possibilities of photooxidation and other gas-liquid photochemical reactions.

O’Brien, Ley and others of their group [62]–[65] did multiple studies on the use of Teflon AF

for flow reactions for processing of gas-liquid reactions. They used the set-up from Figure

4.10 (a) to prove their concept that Teflon AF works as microporous gas-permeable material

that can afford a practically simple method of achieving efficient gas−liquid reactions [62].

They had a tubing of Teflon AF-2400 passing through a chamber of ozone. The permeation

of O3 across the Teflon was demonstrated by bleaching a 0,1 M flow stream of 1,1-

diphenylethene in methanol. After this, they built the set-up shown in Figure 4.10 (b) where

they did many experiments with different gas-liquid reactions.

Figure 4.10 (a) proof of concept reactor [65] (b) schematic of the tube-in-tube reactor configuration [64]

Ponce et al. [66] used a type I LCW made of Teflon AF 2400 tube as a microreactor. Figure

4.11 (a) shows their schematic setup. They used a custom-made T-fluidic cell to connect the

liquid and optic pathway. The Teflon AF was used as a waveguide but also as a membrane

for a gas-liquid reaction. The liquid filled Teflon AF waveguide allowed the gas (N2, O2) to

transfer through the Teflon AF in the liquid core filled with an aqueous solution of methylene

blue, sodium hydroxide and glucose at pH 12,75.

a b

Page 21: The use of microreactors/optofluidics to study

20

When no oxygen was present, the solution turns colourless, meaning that methylene blue

has been reduced to leuco-methylene blue by glucose. If oxygen is transported to the

solution, it will cause the oxidation of methylene blue and create a blue colour. All these

steps are schematically illustrated in Figure 4.11 (b). Ponce et al. [67] also demonstrated that

their LCW membrane reactor is suitable for the studies of kinetics of gas/liquid reactions.

They studied the redox kinetics of vanadium substituted heteropoly acids.

Figure 4.11 (a) Schematic liquid core waveguide membrane setup of in situ sensing and gas permeation using a Teflon AF

tube [66] [68] (b) Schematic illustration of the methylene blue redox cycle.

However, these reactions are not photochemical reactions, they showed that it would be

possible to do the photoreactions in the Teflon AF waveguides. The waveguide could be

filled with a photodegradable solution and the light could be guided through the solution.

As mentioned in chapters 2 and 3.2 the flow matters, when doing the reaction without a flow

the transmission could be measured at the end of the waveguide and the reaction could be

followed. The disadvantage of not using a flow are the small reaction volumes that can be

used. When using a flow for the reaction, the speed determines the yield of the reaction.

a b

Page 22: The use of microreactors/optofluidics to study

21

4.2 Aerogel microchannels

Another possible material for TIR is aerogel, where the aerogel is used as a cladding

material. Aerogel is a nanostructured material, consisting of a crosslinked polymers with a

solid structure and a large number of air-filled pores. These pores create a high porosity,

which is sometimes called “solid air”, creating refractive indexes close to one (air). This

makes them perfect for total internal reflection based waveguides [23], [69]–[72].

Aerogels consist of three categories: inorganic, organic, and inorganic-organic hybrids. The

inorganic aerogels were the first to be studied and are most widely used. They are formed

from metal alkoxides, like alumina (Al2O3), titania (TiO2), or zirconia (ZrO2), but mostly used

is silica (SiO2) [23], [69], [73]–[76]. Organic aerogels consist of connected colloidal particles

or polymeric chains, like Resorcinol–formaldehyde and melamine–formaldehyde aerogels.

Carbon aerogels are obtained by pyrolysis of organic aerogels [23], [73]–[75], [77], [78].

Hybrid aerogels are synthesized by building organic polymers within an inorganic aerogel

matrix [23][73][74][79][80].

Aerogels have many advantages due to their high porosity and the structural nature of their

microscopic and macroscopic features. Aerogels have an ultralow thermal conductivity,

ultralow dielectric constant, ultralow sound speed, high specific surface area, ultrawide

adjustable ranges of the density and most important they have an ultralow refractive index

[75]. The refractive index of the aerogel is almost one, being so low it will almost always be

lower than the refractive index of any liquid. Figure 4.12 (a) shows a liquid core waveguide

in aerogel cladding. The aerogel consists of connected secondary particles, with pores filled

with air. The secondary particles are formed from primary particles (Figure 4.12 (b)), the air

in the aerogel pockets is responsible for guiding light through the core. To create the LCW

the aerogel-based optofluidic waveguides exists of microchannels inside the aerogel. The

channels are filled with a suitable liquid as liquid core, as long as the liquid doesn’t penetrate

the aerogel network and stays in its channel, any type of liquid can be used [23][71].

Figure 4.12 (a) liquid core optofluidic waveguide created in aerogel (b) nano-scaled particles and pores in aerogel [23]

Page 23: The use of microreactors/optofluidics to study

22

4.2.1 Fabrication

Aerogels are synthesized using a sol-gel process, which is a method for producing solid

materials from small molecules. The typical steps of the synthesis are shown in Figure 4.13

and can be divided into 3 general steps. The first step is the gel preparation: where the sol

is formed from reactants with the addition of a catalyst (b). The condensation of the sol leads

to the formation of alcogel, where the sol particles are linked (c). The second step is the

aging of the gel (d), where the aging process strengthens the gel. The last step is the drying

of the gel, where the gel is freed of the pore liquid (e). To prevent the collapse of the gel

structure, drying has to take place under special conditions and is mostly done with

supercritical CO2 [23], [69], [70], [74]–[76], [81].

Figure 4.13 Preparation of monolithic aerogels by the sol-gel method [23]

Some types of aerogels are hydrophilic, because of the hydroxyl groups on the surface of

the secondary particles [76]. When filling the channels with water, the water enters the

porous microchannel system of the aerogels. The hydroxyl groups can take part in hydrogen

bonding with water and together with high capillary stress this leads to erosion which leads

to collapse of the structure and the waveguide. Therefore the channel walls should be made

hydrophobic to prevent the penetration of water. With the presence of a non-polar

hydrophobic group they can meet the stability requirements for long-term use when using

an aqueous core. These surface modifications can either be performed after the channel

forming, during the aging step or after the drying step [23]. Eris et al. [70] used

hexamethyldisilazane (HMDS) solved in supercritical CO2 to turn the hydrophilic silica

aerogels into hydrophobic ones. Özbakır et al. [46] treated silica aerogels with (HMDS)

vapour, the hydrophobic nature was tested during experiments over several weeks and

water did not penetrate into the pores.

There are several possibilities to fabricate channels in aerogel blocks. One of the ways is by

forming the aerogel around a capillary and after formation withdrawing the capillary from

the block. Xiao et al. [71] created a silica aerogel block around a glass capillary and after the

drying step it was withdrawn from the block. They filled the channel with water and used a

Page 24: The use of microreactors/optofluidics to study

23

fibre to guide a 635 nm laser into the channel. The light was well guided and the waveguide

loss was measured for a 16 mm long channel filled with water, the total loss was 1,5 dB/cm.

This is much more than the Teflon AF, which were in dB/m. The withdrawal generally leads

to breakage due to the attachment of the capillary to the aerogel network. There are limited

shapes and sizes for the channels with this technique [23].

Furthermore, preformed shapes can also be used and the aerogels are formed around it.

When the preformed shapes are soluble in the supercritical CO2 (from the drying step) and

not in the solvents used in the aging steps, the preform can be easily removed. However,

there are not that many polymers that are soluble in supercritical CO2 and insoluble in the

other solvents used [23]. Eris et al. [70] created a fibre from trifluoropropyl POSS in a U-

shape. Before gelation, they placed U-shaped trifluoropropyl POSS fibre inside the sol

solution. Because the trifluoropropyl POSS solves in supercritical CO2, it was removed

during the drying step. Figure 4.14 (a) shows that the U form is clearly visible at the top of

the aerogel block. In Figure 4.14 (b) the channel is still filled with air and due to the absence

of water, the light is not guided. Figure 4.14 (c) shows that when the channel is filled with

water, the light is nicely guided through the channel and creates an LCW.

Figure 4.14 (a) aerogel block with an empty U-shape channel (b) laser light focused from the top of the channel before it was

filled with water (c) laser light focused from the top on the right end of the water filled channel [70]

Another possibility is to fabricate the aerogels and to create channels with a laser. Yalızay et

al. [69][72] used femtosecond laser ablation to create channels in a hydrophobic silica

aerogel. The ultrafast laser has a better focus, creating a better precision and a smoother

channel. After the formation of the channel, they filled it with ethylene glycol instead of water,

because water was evaporating relatively fast due to open-ended channels and ethylene

glycol is less volatile than water. A laser at 632,8 nm was coupled into the channel through

an optical fibre. The propagation loss of the optical waveguide was then calculated as 9,9

dB/cm. This high loss is mainly due to increased roughness of the channel surfaces

produced by the laser, leading to higher scattering losses in the waveguides.

A manual drilling technique is also a possibility [23]. Özbakır et al. [46] synthesised

hydrophobic silica aerogels and created microchannels by manual drilling using a drill bit.

A hand drill bit was put to the aerogel surface and moved slowly into the sample (rotating

continuously) to prevent stress build up, as shown in Figure 4.15 (a). A straight channel with

a diameter of ~2,1 mm was opened. An L-shaped channel was formed by creating a

horizontal channel and another channel starting from the top, see Figure 4.15 (b), creating

an L-shaped channel shown in Figure 4.15 (c).

Page 25: The use of microreactors/optofluidics to study

24

Figure 4.15 (a) channel formation in aerogel by manual drilling (b) L-shaped channel formation in aerogel by manual

drilling (c) Side-view of inclined L-shaped channel in monolithic aerogel [46]

The laser was coupled into the horizontal part of the channel by an optical fibre, when the

light was coupled into an empty channel, no guiding of light to the opposite end of the

channel was observed. As can be seen in Figure 4.16 (a), the light is scattered strongly.

When filling the channel with water, see Figure 4.16 (b), the light was guided and delivered

to the opposite end of the channel. A propagation loss of 1,45 dB/cm was found.

Figure 4.16 (a) light coupled into an empty channel (b) light propagation in the water filled channel [46]

The losses in the aerogel waveguides are large, in dB/cm. This is probably because the

channels are not smooth, causing the light not to be in the right angle for TIR and leaving

the waveguide. Likewise, the not so smooth channels create more light scattering, causing

more losses.

4.2.2 Detection and Reaction use

Aerogels guide light by the principle of total internal reflection, which works to create a

longer pathlength. Özbakır et al. [46], Eris et al. [70] and Xiao et al. [71] showed that the light

can be guided through the waveguide channel. Therefore, the aerogel-based optofluidic

waveguides could be used for applications including light-driven detection or identification

and quantification of particular chemical compounds [46]. However, none has done

experiments other than proving it to be possible to use the aerogels for light-driven

detection.

Page 26: The use of microreactors/optofluidics to study

25

The aerogels are gas permittable due to the open porous network. The pores are open, so

the aerogels allow gasses to diffuse from the surroundings in the liquid core and the other

way around [23][76][82]. Like Teflon AF, they could be used for gas sensing or

heterogeneous gas-liquid reactions.

The aerogel based waveguides could also be used as a micro photoreactor, Özbakır et al.

[6] demonstrated an aerogel-based micro photoreactor for the degradation of methylene

blue (MB). They used the same fabrication as before [6], a hydrophobic silica based aerogel

with an L-shaped channel (1,1 cm long horizontal channel and 3,7 cm long inclined channel).

Figure 4.17 shows a schematic overview of the experimental set-up that was used for the

photodegradation of MB dye in an aqueous solution, an L-shaped waveguide was used.

Using a T-connector a fibre was connected to the channel entrance to connect the laser to

the reaction, the other port of the T-connector was connected to a syringe to fill the channel

with the aqueous MB solution. Due to the hydrophobic walls, the MB solution did not

penetrate into the pores of the aerogel. The MB might have been adsorbed on the surface

of the aerogel, influencing the concentration and the results. The aerogel channel was filled

with MB solution and kept in the dark for 2 hours, samples were collected at different times

and the concentration was measured by a nanodrop spectrophotometer. The

concentrations were constant, indicating that MB was not adsorbed at the channel surface.

Figure 4.17 Schematics of experimental set-up used for photochemical reactions, the waveguide is pictured linear but the L-

shaped waveguide was used [6]

Page 27: The use of microreactors/optofluidics to study

26

A laser with wavelength 388 nm was coupled into the aerogel waveguide filled with MB

solution. Because of the guiding properties, the MB could be degraded along the full length

of the channel. The amount of degradation is then depending on the exposure time and the

laser power. Figure 4.18 shows the decrease in MB concentration by a longer exposure time

and a constant power. The photodegradation rate was highest at the entrance, due to

absorption and scattering. These experiments were executed using no liquid flow, but it

should be possible to operate with a continuous flow, using a suitable pump. The difference

in photodegradation rate would then be solved and bigger amounts of MB could be

degraded.

Figure 4.18 The concentration of aqueous MB solution in an aerogel channel under irradiation at 388 nm at different times

with a varying power of the incident light [6]

Özbakır et al. [83] also used the photo microreactor for the photocatalytic degradation of

phenol over titania. The silica-titania aerogels were fabricated by adding anatase titania

powder in the sol during the sol preparation in different weight percentage ranging from

1,7 wt% to 50 wt%. Afterwards, they were aged, dried and transformed from hydrophilic

into hydrophobic using HMDS. For their experiments they used the same set up as Figure

4.17, just the waveguide contains titania particles now, see Figure 4.19.

Figure 4.19 aerogel waveguide with catalytic titania [83]

Because the phenol may be adsorbed on the channel wall, the channels were filled and kept

in the dark for 60 minutes, Figure 4.20 (a) shows that the concentration of phenol remained

almost constant. The photodegradation of phenol was also investigated without titania

particles at 366 nm in a silica aerogel waveguide. Figure 4.20 (b) shows that the

concentration remained constant which indicated that there is no photodegradation, just

the photocatalytic degradation of phenol. Figure 4.20 (c) shows a decrease in phenol

Page 28: The use of microreactors/optofluidics to study

27

concentration over time within the first 10 minutes, after 10 minutes the reactor system

henceforth operated at steady state at constant flow.

Figure 4.20 (a) concentration of phenol in the dark over time (b) concentration of phenol in silica aerogels without titania at

a laser wavelength of 366 nm over time (c) concentration of phenol in silica-titania aerogels at a laser wavelength of 366 nm

over time [83]

Compared to the Teflon AF waveguides the aerogel has a much better refractive index,

creating a smaller critical angle. Another advantage from the aerogel over Teflon AF is the

possibility to use heterogeneous catalysts. However, the losses in the aerogel are too high

(~dB/cm) and the channels can only be a few centimetres long. These dimensions and

losses need to be improved before the aerogel could be used as microreactor of

photochemical reactions.

Page 29: The use of microreactors/optofluidics to study

28

5 Interference based waveguides

Another approach for light guiding is the use of interferences-based waveguides. This

means the interference of two waves forming a wave with a higher, lower or the same

amplitude. The waves can be correlated or coherent with each other, either because they

come from the same source or because they have the same or nearly the same frequency,

resulting in constructive (Figure 5.1 left) or destructive (Figure 5.1 right) interference [45].

Figure 5.1 interference of waves, (left) constructive and (right) destructive [Wikipedia]

In interference-based waveguides, multiple layers of dielectric materials are used as the

waveguide cladding. These layers create multiple reflections that can interfere

constructively or destructively. Figure 5.2 shows that the light is partially reflected at each

interface of the structured cladding material. The figure shows only a small amount of the

reflected waves, but for the reflected power all the reflected waves need to be added up.

The idea is that an almost perfect reflection in the liquid core can be achieved even if the

core has a lower refractive index than all of the cladding layer materials [23][25][45].

Figure 5.2 Schematic view of multiple interfering reflections [45]

The interference waveguides can be based on the photonic bandgap effect. Photonic

bandgap materials or photonic crystals are dielectric materials with a periodic structure. The

photonic bandgap represents the forbidden range where certain wavelengths cannot be

transmitted through the material [84]–[86]. The photonic crystal can be layered in different

dimensions, see Figure 5.3. The first dimension (Bragg fibres) and second dimension

(photonic crystal fibres) are discussed below. The other interference based waveguides that

will be discussed are Anti-resonant reflecting optical waveguides (ARROW).

Figure 5.3 One, two and three dimensional photonic crystals [87]

Page 30: The use of microreactors/optofluidics to study

29

5.1 Bragg fibres

The first interference based waveguide to be discussed is the Bragg fibre, also known as the

one dimensional photonic crystal. The Bragg fibres contain multiple dielectric layers of low

(n1) and high (n2) refractive indexes around the hollow core, shown in Figure 5.4. The light

is guided due to the photonic bandgap [88]–[91].

Figure 5.4 A schematic of a Bragg fibre with a liquid core. ncore < n0 < n1 [92]

5.1.1 Fabrication

Bragg fibres rely on the layered dielectric cladding and are mostly used for air filled

mediums [25]. The Bragg fibres are mostly made by placing multiple layers of film

alternately on each other, the films are rolled into a hollow multilayer tube that acts as a

preform. The preform is then placed in a draw tower to draw a long fibre with thin films

[25][93][94]. Low losses (~ dB/km) were found with the Bragg waveguides [95]. A few years

ago they filled the fibres with a liquid and tried to guide the light in a liquid medium, like

water. Qu and Skorobogatiy [88] demonstrate a liquid-core Bragg fibre. The Bragg fibre

was 40 cm long and had a large core (~0,8 mm) surrounded by an alternating polymethyl

methacrylate and polystyrene. The core was filled with NaCl solutions with different weight

concentration. The refractive indexes of the salt solutions were between 1,33 and 1,38. The

transmission spectra of the NaCl solutions that were analysed by a spectrometer are shown

in Figure 5.5. Observed is a blueshift in transmission spectra when the RI of the analyte

increases.

Figure 5.5 Experimental transmission spectra of Bragg fibre filled with NaCl solutions [88]

Page 31: The use of microreactors/optofluidics to study

30

5.1.2 Detection and Reaction use

Filling the Bragg fibres with a liquid is a relatively new phenomenon and first happened less

than 10 years ago. Only a few Bragg fibres filled with liquids have been published, there

were all about fabrication and how to fill the fibres with a liquid. Little is known about using

the Bragg fibres as a longer pathlength or as a reaction vessel, but that does not make it

impossible. However, little is known about its use, guidance and loss when filled with liquid,

so I would not recommend it.

Page 32: The use of microreactors/optofluidics to study

31

5.2 Photonic crystal fibres (PCF)

Photonic crystal fibres are optical fibres based on the properties of photonic crystals, also

known as the two dimensional photonic crystals. The cladding consists of microscopic

hollow capillaries along the length of the entire fibre [8]. The holes of the first PCF were too

small to expect a photonic bandgap and they had a solid core. The guiding principle was

based on modified total internal reflection, where the holes create a low refractive index.

When using bigger holes or a hollow core the light is guided by the photonic bandgap

principle due to the photonic crystal [84][96]. Here the photonic crystal cladding forbid the

light with a certain range of frequencies to propagate, the light within these bandgaps is

confined and guided in the core [97].

5.2.1 Fabrication

PCFs are commonly made of silica, but glass and polymers are also used [25]. Their

fabrication is based on the ‘stack-and-draw’ method, see Figure 5.6 [8][98][96]. First, the

capillaries and rods need to be drawn with specific diameters and cut into short lengths,

they have a very strong impact on the transmission characteristics of the fibre. The capillaries

and rods are stacked in a larger diameter glass tube, also called a jacket tube, forming a

preform. The preform is pulled down into a cane under pressure to prevent holes between

the capillaries. Finally, the canes are drawn into a fibre with a specific diameter using a

drawing tower, a typical fibre diameter is 40 to 100 µm [98][99]. This is already smaller than

the core of the TIR based waveguides, but this is just the fibre diameter, not even the holes.

It is important to remember that even smaller volumes go into the PCFs.

Figure 5.6 schematic of the 'stack-and-draw' method and some examples: Solid-core photonic crystal fibres (a);

suspended-core photonic crystal fibres (b); hollow-core photonic bandgap fibre (c) and Kagome photonic crystal fibre (d)

[8][96]

Page 33: The use of microreactors/optofluidics to study

32

As mentioned before the core of the PCF can be solid and hollow, the first one to be

discussed is the solid-core PCF (SC-PCF). The SC-PCF has a solid core surrounded by

periodic cladding channels, see Figure 5.6 (a). The hollow channels reduce the refractive

index of the cladding below the refractive index of the core, creating a guidance mechanism

based on total internal reflection [8]. Kurokawa et al [27] created a SC-PCF to improve the

fibre bandwidth for internet traffic. They created a 25 km long silica SC- PCFs with 60 holes.

The hole diameter was 7,6 µm and the outer diameter of the fibre was 125 µm. Figure 5.7

shows a loss spectrum of the fabricated PCF with the lowest loss, a loss of 0,18 dB/km was

found in the 1550 nm wavelength region. The big loss at 1380 nm is caused by the OH

absorption. This is really low, which is pretty impressive compared to the waveguides

discussed above. On the other hand, it is not necessary, to have such low losses or long

waveguides when using them as a microreactor. When using the PCF for sensing or

detecting applications the holes can be filled with liquid. The light is guided in the solid core,

but interacts with the samples by an evanescent field that penetrates into the holes. The

amount of evanescent field penetration into the holes can be controlled by parameters such

as core size, hole diameter and the distance between the hollow channels [8][100]. This

principle might work, but the question would still be of how much light would go into the

channels by evanescent field penetration and would it also go into the second row of

channels? This seems a little bit unlikely, but it would also be really hard to fill the extremely

small channels, use a flow and keep a practical pressure.

Figure 5.7 Loss spectrum of solid-core PCF with the lowest losses [27]

Because of the large flow resistance in the micro cladding holes in SC-PCF, the suspended-

core PCF was fabricated. It consists of a solid core held in the air by three nanowebs, shown

in Figure 5.6 (b) [8]. Webb [101] fabricated multiple suspended-core PCFs. Instead of

stacking, they drilled holes into a silica preform, but the drawing inside a jacked process

was the same. The outer diameter of the fibres was 125 µm, the core variated from 0,8 to

1,8 µm and the holes were around 8 µm in diameter. In each case, they drew several tens of

meters fibre to prove there was no change in structure, but they used a meter of fibre for

the experiments. They found a loss of 0,29 dB/m at 1550 nm, the relatively high loss was

mainly due to scattering because of the rough surface followed by the drilling. This is an

acceptable loss when using the suspended-core PCF as a microreactor for photochemical

reactions. For sensing applications the holes are filled with the sample, the amount of

evanescent field that extends into the holes can be changed by changing the core diameter

[8][102]. The suspended-core PCF works ideal for broadband transmission [103]. The

channels are already a little bit bigger than the SC-PCF and would be better for the flow

resistance. However, the evidenced field penetration would still be questionable, but there

Page 34: The use of microreactors/optofluidics to study

33

is only one layer of channels around the core and the core is smaller. The chance of the light

to be in the channels is bigger but also is the loss bigger. The loss is still acceptable when

using waveguides of a few meters. The reaction may take a little bit longer because the light

is mostly in the solid core and sometimes in the liquid channels. However, it would be

acceptable/neglectable because of the small volumes being used.

The other type of PCF is the hollow-core photonic crystal fibre (HC-PCF), it has a core with a

low refractive index. It can be fabricated in different structures, mostly by the “stack and

draw” method, where the hollow core is created by leaving out some capillaries at the core,

usually seven [99]. The first type of HC-PCF, known as hollow-core photonic bandgap fibre

(HC-PBF) is shown in Figure 5.6 (c). It consists of a central hollow core surrounded by a

honeycomb lattice of hollow channels. The light is guided by a two-dimensional photonic

bandgap, because of the photonic crystal [8][104]. Due to these photonic bandgaps, in the

HC-PBF the wavelength of the light has to be matched to a specific spectral bandgap,

otherwise, high losses are to be found. By filling the hollow cores completely with a liquid

media, the bandgap can shift to lower wavelengths [36]. Like the SC-PCF, the HC-PBF can

provide extremely low transmission losses, as low as 1,2 dB/km [105]. Which is again pretty

impressive, but not completely necessary when not using more than a few meters for the

microreactor. Because the HC-PBF only reflects the light at certain wavelength ranges, it is

not very useful when using multiple wavelengths.

Another type of HC-PCF is the kagome PCF, which is named after its particular lattice

arrangement in the cladding structure, see Figure 5.6 (d) [8]. An important property of

kagome HC-PCF is the wide transmission bandwidth, covering from the UV to the near-

infrared, this fibre is, therefore, the preferred choice for broadband spectral measurements

[106][107]. But the kagome PCF has the disadvantage of a higher transmission loss, 1 dB/m

[108]. This is much higher than the HC-PBF, but the advantage is that the light reflects in a

bigger range of wavelengths. This loss is still acceptable when using it as a microreactor

because maximal a few meters of waveguide would be used. Williams et al. [109] used a 30

cm long kagome PCF with a core diameter of 19 µm and a loss of ~1 dB/m. They used the

kagome PCF for the photoisomerization of azobenzene. In 10 seconds the reaction process

was completed, whereas using a 1 cm cuvette, almost three orders of magnitude more

power was needed.

When filling the HC-PCF with a liquid sample, the sample must be pumped into the core of

the fibre. The fibre can be filled in two different ways, selectively or completely filled. When

only filling the core, the cladding holes are air-filled. The refractive index of the core exceeds

the average refractive index of the fibre cladding, creating a total internal reflection and a

broadband guiding waveguide. Sometimes the holes are thermally collapsed by a fusion

splicer to prevent them from being filled [110][111][112]. When completely filling the HC-

PCF, the core is filled with the sample. The holes are homogeneously filled with a liquid

whose index of refraction is lower than the refractive index glass [97][113]. However, only

filling the core is proved to be more suitable than filling the airholes and the core, because

light can be more easily confined in the core region [114].

Page 35: The use of microreactors/optofluidics to study

34

5.2.2 Detection use

The PCF is a waveguide, this means that it could be used to create a longer pathlength as

described in chapter 0. The pathlength could be improved for absorbance, creating a

longer pathlength so low concentration compounds could be detected [114][115]. Williams

et al [116] compared a suspended-core PCF with a Kagome photonic crystal fibre for long

pathlength fluorescence measurements, each type of PCF was ~30 cm long and the

excitation source was a picosecond pulsed 470 nm diode laser. Detection limits were

established for both fibres using fluorescein. Figure 5.8 (left) shows the fluorescence

spectrum recorded in HC-PCF (concentration of 2x10-11M) and SC-PCF (concentration of

3x10-10M) approaching the detection limit. The solid line is after the subtraction of the water

background, it allowed the removal of the Raman band. To study fluorescence lifetime

detection from within the fibres, rhodamine B fluorophore was used. The fluorescence

decay measured for 10-9M is shown for each type in Figure 5.8 (right). In the SC-PCF a

lifetime of 3,64 ns was measured, in the HC-PCF a shorter decay time of 2,12 ns was

observed. However, the decay is deviating from a mono-exponential behaviour. This is a

large uncertainty in the latter lifetime value, but it is probably because of surface-adsorbed

rhodamine B.

Figure 5.8 (left) Fluorescence spectra of fluorescein, with (solid) and without (dached) water background removal.

(right) fluorescence decays measurements for an aqueous solution of rhodamine B [116]

The PCF waveguides can also be used to improve the Raman signals [117][118][119]. Yan

et al. [36] used the HC-PBF for the detection of low-concentrated analyte molecules in

aqueous solutions. They used an HC-PBF with a core diameter of 10 µm and a length of 10

cm and a green excitation laser (λexc = 532 nm). They measured the drug levofloxacin, using

the waveguide and compared the measurements to using a cuvette. As can be seen in

Figure 5.9 a, the waveguide increased the signal to noise ratio more than one order of

magnitude. Figure 5.9 b shows that the limit of detection was strongly improved and the

Page 36: The use of microreactors/optofluidics to study

35

Raman intensities in the waveguide show an excellent linearity to the levofloxacin

concentration.

Figure 5.9 Comparison of Raman measurements of levofloxacin in a cuvette and using the HC-PBF. (a) The Raman spectrum

of a solution of levofloxacin measured using the cuvette (black) and the HC-PBF (red). (b) The Raman peak intensities vs the

concentration of the solution [36]

5.2.3 reaction use

The photonic crystal fibres are not only used for detection use. Many scientists wrote reports

on photoreactions using PCF reactors and multiple reviews are written about the PCFs

and/or their use in photochemical reactions. Cubillas et al. [8] have published a good review

about the use of PCF for photochemical reactions. Below are some examples as proof of

principle and some newer publications.

As a proof of principle of using PCF for photoreactions, Chen et al. [7] used a liquid filled

kagome PCF to study the photoaquation of vitamin B12 (cyanocobalamin, CNCbl).

Depending on the pH of the aqueous solution, the CN- in CNCbl exchanges for H2O.

Forming [H2OCbl]+ creates a shift in the absorption to a shorter wavelength, see Figure 5.10

(a). The solution is most stable at pH 7 and converts rapidly at both the higher and lower

extremes of pH. At pH 2,5 the conversion in the fibre occurred within 1 minute, roughly 1000

times faster than in a cuvette. Using pH 7,5 the reaction was completed in the PCF in 10

minutes, but using a cuvette no change was observed after 17 hours exposure time, see

Figure 5.10 (b). Unterkofler et al. [120] used the same reaction and also a kagome PCF, but

they showed that the reaction products can be monitored online by a high resolution mass

spectrometer.

Page 37: The use of microreactors/optofluidics to study

36

Figure 5.10 (a) Change in the absorption spectrum as a result of the photochemical conversion of CNCbl to [H2OCbl]+

(b) Temporal evolution of the photoaquation reaction at 500 and 550 nm in a cuvette and a kagome PCF at pH 2,5 and 7,5.

Note the different time scales [7]

McQuitty et al. [22] used a kagome PCF as a micro flow-reactor for the efficient activation

and analysis of photoactivatable drugs on a much shorter time scale than normally used

methods. They used a di-nuclear ruthenium complex, under irradiation it can lose an indane

ligand and bind to DNA. The complex can also undergo aquation in aqueous solution in the

dark, losing a chloride ligand. Aquation can be prevented by the presence of a high

concentration of Cl- ions. They studied the process in the presence of multiple biomolecules

as a guide to its possible intracellular behaviour. They coupled a high resolution mass

spectrometer online with the fibre, to gain more information about the process. This

combined optofluidic device proved very useful for rapid analysis of photoactivatable drugs.

The samples were irradiated in a cuvette for 14 hours and transferred through the PCF in 12

seconds, with the same results.

The PCF can also be used as a microreactor for photocatalytic reactions. As proof of

principle Cubillas et al. [26] used a kagome PCF to study complex multicomponent

homogeneous catalytic reaction. The well-known, photo-Fenton chemistry was chosen as

the photocatalytic system, where hydrogen peroxide is catalytically decomposed by ferrous

ions to give hydroxyl radicals. The photo-Fenton reaction samples consisted of methylene

blue (MB) as model dye pollutant, FeCl3 as the catalyst, oxalate as the ligand, hydrogen

peroxide as the source for OH radicals, and water as the solvent. The activity in the PCF was

two times higher than that in the cuvette and much less concentration and volume was

needed.

The advantage of the PCF over the Teflon AF waveguides is the possibility of using a

heterogeneous catalyst. The solid catalyst particles are placed in the fibre core without

disturbing the light guiding properties. Cubillas et al. [121] demonstrated the prove of

principle using a kagome PCF for heterogeneous photocatalytic reactions. The main

challenge for combining heterogeneous catalysis with the HC-PCF was placing the solid

catalyst particles in the fibre core without disturbing the light guiding properties. The fibre

Page 38: The use of microreactors/optofluidics to study

37

core acts as an optofluidic microreactor for heterogeneous catalysis, and the catalyst

nanoparticles (Rh) are located on the inner wall of the core, away from the intensity

maximum. This minimizes the overlap between the light and the particles, preventing

unwanted heating of the catalyst. The performance of the catalytic PCF was studied by the

heterogeneous catalytic hydrogenation of azobenzene (Disperse Red 1 or DR1). The

hydrogenation of DR1 gives an absorbance decrease around 480 nm and can, therefore,

be conveniently monitored by absorption spectroscopy at a single wavelength. The

measured catalytic activity of the fibres was proportional to the amount of catalyst surface

coverage and in good agreement with standard plug-flow reactor models.

Compared to the TIR based waveguides the PCF has smaller losses and smaller volumes.

However, the PCF has one big disadvantage, it is not gas-permeable. There are no pores in

the material and gasses cannot go through. When filling the pores with gasses they still

would not be in contact with the liquid reaction mixture and no heterogeneous gas-liquid

reaction would happen. Gas bubbles could be made in the reaction mixture, but the

presence of bubbles destroys light guidance [66].

Page 39: The use of microreactors/optofluidics to study

38

5.3 Anti-resonant reflecting optical waveguides (ARROW)

The last type of interference-based waveguides to be discussed is the anti-resonant

reflecting optical waveguide (ARROW), using the principle of thin-film interference to guide

light with low loss. In Thin-film interference, light waves are reflected by the upper and lower

boundaries of a thin film, either enhancing or reducing the reflected light, see Figure 5.11.

The thickness of the film is very important, when the thickness of the film is an odd quarter-

wavelength of the light (1/4 λ, 3/4 λ, 5/4 λ), the reflected waves from the surfaces interfere

and cancel each other out. When the thickness is half a wavelength of the light (1/2 λ, λ, 3/2

λ), the reflected waves reinforce each other, increasing the reflection [122].

Figure 5.11 thin film interference [122]

For simplicity, consider a one dimensional structure of only two cladding layers with

refractive index n1 and n2 (n1 > n2), shown in Figure 5.12. In order to increase the light

reflected back into the core, each cladding layer must be designed to fulfil a specific

condition. By considering the similarity of a cladding layer with a Fabry–Pérot reflector, such

conditions correspond to the anti-resonance state, which performs the maximum reflectivity

[123][124]. When the Fabry-Perot is in resonance, the light in the layer constructively

interferes with itself, a high transmission is found. When the Fabry-Perot etalon is in an anti-

resonance, the light in the layer destructively interferes with itself and no transmission is

found providing a large reflection. This does not mean that the reflection only works at a

narrow wavelength. The resonances of a Fabry-Perot occur over a narrow band of

wavelengths and the anti-resonances occur over a very broad band of wavelengths [125].

Figure 5.12 one dimensional ARROW waveguide principle

Page 40: The use of microreactors/optofluidics to study

39

These specific conditions of antiresonance make the thickness of the thin layer important.

The required layer thickness (t1 and t2) can be calculated by [125]:

𝑡1,2 = 𝜆

4𝑛1 (2𝑁 − 1) (1 −

𝑛 𝑐2

𝑛1,2 2 +

𝜆2

4 𝑛1,22 𝑡𝑐

2)−0,5

𝑁 = 0, 1, 2, … (4)

Where N is the order of interference, n1,2 is the refractive index of the cladding layers, nc the

refractive index of the core, tc the thickness of the core and λ the used wavelength

[45][126][127]. The guiding efficiency can be improved by adding more layers that fulfil the

antiresonance conditions.

5.3.1 Fabrication

The ARROW fabrication can happen in one of the two main approaches: by bulk or surface

micromachining process [124]. By bulk micromachining, the microchannels are formed by

selectively removing parts from a bulk silicon substrate, see Figure 5.13. The substrate is

covered with a protective layer on places where the substrate may not be removed [128].

After forming the channels the protective layers are removed and the thin cladding layers

are formed. Using deposition techniques, like low pressure chemical vapour deposition

(LPCVD), plasma enhanced chemical vapour deposition (PECVD) or atomic layer deposition

(ALD), thin layers are formed on the substrate. The channels are covered by a second

substrate with the same thin layer formation, but without any channels [124]. Campopiano

et al. [129] used the bulk micromachining process to form a hollow core ARROW waveguide.

Two layers of silicon dioxide and silicon nitride were deposited by PECVD onto silicon

substrates, one with and one without a core. The two substrates were joined together to

form the waveguide.

Figure 5.13 Typical steps of a bulk micromachining process (a) substrate preparation; (b) deposition of a protective layer

(SiO2); (c) removing parts of the protective layer; (d) substrate removing; (e) deposition of protective layer for a selective

area; (f) substrate removing for creating deeper channels; (g) creation of channels in a substrate [128]

Page 41: The use of microreactors/optofluidics to study

40

By surface micromachining process the channel core is obtained by depositing and

patterning a sacrificial material on a single substrate in between the thin layers, see Figure

5.14. The thin layers can be formed, like in bulk micromachining process, by techniques like

low pressure chemical vapour deposition (LPCVD), plasma enhanced chemical vapour

deposition (PECVD) or atomic layer deposition (ALD). The sacrificial material is removed at

the end of the process [124][128].

Figure 5.14 Typical steps in a surface micromachining process: (a) substrate preparation; (b) deposition of a sacrificial

layer; (c) creation of a hole by patterning and removing the sacrificial layer; (d) deposition of a structural layer; (e) shape

definition by patterning and removal of the structural layer; (f) release of the suspended structure [128]

Yin et al. [130] used the surface micromachining process to form a hollow core ARROW

waveguide. Alternating silicon dioxide and silicon nitride were deposited on a silicon

substrate using PECVD and a photosensitive polyamide was used as sacrificial material. The

channels had a width varying from 6 to 50 µm. These are really small channels, it could

create a large flow resistance and it might even be too small for a coupling to a HPLC. Gollub

et al. [131] used the surface micromachining process to form a hollow core ARROW

waveguide using just two layers and photolithography as sacrificial material. The first layer

was a 2 µm thick silicon dioxide film and as the second layer a silicon nitride of 100 nm thick

film was used, the films were deposited by PECVD on the substrate. The optical loss

increased as the waveguide width decreased, see Figure 5.15, the lowest loss was at a width

of 100 µm having a loss of 10 dB/cm. These high values come from the fact that the side

walls were not smooth and some sacrificial material that stayed on the inside of the channel

could have been absorbing light. These losses are really big and not acceptable when using

the ARROW waveguide as a microreactor for photochemical reactions.

Figure 5.15 measured optical loss (dB/cm) as a function of the waveguide width [131]

Page 42: The use of microreactors/optofluidics to study

41

5.3.2 Detection and reaction use

Till now, there are no publications on into the possibility of using the ARROW as a

microreactor, probably because the losses are too high, so the waveguide must be really

short. If by someday the losses are decreased, it could be used as a microreactor. However,

the ARROW waveguide has been used for multiple detection applications.

The ARROW, like all the other named waveguides, can be used as a detection cell for

absorbance, fluorescence and Raman spectroscopy [45], [132]–[136]. Due to the picoliter

and nanolitre volumes the ARROW waveguides can be used for single-molecule detection

[133]. But it is possible to add various functional optical elements on the ARROW chips [124],

[137]–[139]. For example, multimode interference (MMI) splitters based on ARROWs have

been used in on-chip optofluidic interferometers [127], [140]–[143]. Several integrated

optofluidic interferometric devices using the MMI have been demonstrated, like optofluidic

ring resonators [144][145] and Mach–Zehnder interferometers [146][147].

Page 43: The use of microreactors/optofluidics to study

42

6 Conclusions and outlook

To compare the different possible microreactors for photochemical reactions, Table 1

shows the different waveguides with their loss, inner diameter and the used lengths. The

basic requirements when using the waveguides as a micro photoreactor is a low loss of

maximum a few decibels per meter and a possible length of a meter.

The Aerogel and the ARROW do not meet these requirements and would not be

recommended to use as micro photoreactor. However, in the future, both waveguides

would probably be optimized. For the aerogel, the big losses are mostly created by the

faults in the channel formation. When the channels are improved, the losses are decreased

and the length can be extended. Another problem with the ARROW waveguides might be

the inner diameter. The diameter of the waveguide is small and a few nanolitres or even a

few picolitres can enter, which is too small when coupling to a HPLC. The ARROW

waveguides are still relatively new, lots of improvements have already been made and will

probably be made in the future too. Then someday when the losses are low enough and

the length is improved the waveguide can be used as a microreactor. The Bragg fibres do

meet these requirements, but they are very new as a liquid core waveguide. Therefore, they

would need lots of research, time and experiments before they could be used a micro

photoreactor or for longer pathlength spectroscopy experiments.

Table 1 comparison of the different waveguides and their loss, inner diameter and used lengths

Type loss inner diameter used lengths

Micro photoreactor not applicable 10–1000 µm variable Type I Teflon AF ~ dB/m 250 – 500 µm cm - m

Type II Teflon AF ~ dB/m 250 – 500 µm cm - m Aerogel ~ dB/cm ~ 2000 µm 1-5 cm

Bragg ~ dB/km (hollow core) ~ 800 µm cm - m

SC-PCF ~ dB/km 40 to 100 µm cm – 10 m Suspended PCF ~ dB/m 40 to 100 µm cm – 10 m HC-PBF ~ dB/km 40 to 100 µm cm – 10 m

Kagome PCF ~ dB/m 40 to 100 µm cm – 10 m ARROW ~ 10 dB/cm 6 - 50 µm mm - cm

The Teflon AF waveguides do meet the basic requirements when using them for a micro

photoreactor. Both the Type I and the Type II could be used, but I would suggest to use the

Type I. The difference in loss between the two Types is really small, but the light, in the

externally coated waveguide (Type II), is mostly in the glass or quartz capillary. While using

a Type I Teflon AF LCW the light is mostly located in the core. In the future, more

applications for the Teflon AF LCW could be found and more reactions can be done, but I

do not think many improvements can be made for the waveguide.

Page 44: The use of microreactors/optofluidics to study

43

The photonic crystal fibres also meet the basic requirements when using them for a micro

photoreactor. I would recommend the Kagome for the use of a micro photoreactor. The SC-

PCF has a large flow resistance and in the HC-PBF the wavelength of the light has to be

matched to a specific spectral bandgap. The suspended PCF could also work but the light

is mostly in the core and only by evanescent field penetration will it be in the liquid channel.

The kagome PCF has a wide transmission bandwidth and the light is mostly in the liquid

filled core. In the future the PCF might find some new structures, there might be one found

with low losses and have a wide transmission bandwidth.

For the TooCOLD project, I would recommend the Teflon AF type II or the kagome PCF. For

homogenous reactions and homogenous catalytic reactions both the Teflon AF and the PCF

waveguides could be used, all analytes are located in one phase, the solvent and would not

have any complications with the waveguide. For heterogeneous gas-liquid reactions, the

Teflon AF type I LCW would be recommended. The PCF waveguide is not gas-permeable

and would therefore not be very useful. However, the PCF waveguides are recommended

for heterogeneous catalytic reactions. When coating the catalyst on the inside of the Teflon

AF waveguides it would change the refractive index of the cladding and interfere with the

total internal reflection. For the PCF the solid catalyst particles can be placed in the fibre

core without disturbing the light guiding properties.

Page 45: The use of microreactors/optofluidics to study

44

7 The List of References

[1] E. E. Coyle and M. Oelgemöller, “Micro-photochemistry: Photochemistry in microstructured reactors. The new photochemistry of the future?,” Photochem. Photobiol. Sci., vol. 7, no. 11, pp. 1313–1322, 2008, doi: 10.1039/b808778d.

[2] B. P. Mason, K. E. Price, J. L. Steinbacher, A. R. Bogdan, and T. D. McQuade, “Greener approaches to organic synthesis using microreactor technology,” Chem. Rev., vol. 107, no. 6, pp. 2300–2318, 2007, doi: 10.1021/cr050944c.

[3] H. Lu, M. A. Schmidt, and K. F. Jensen, “Photochemical reactions and on-line UV detection in microfabricated reactors,” Lab Chip, vol. 1, no. 1, pp. 22–28, 2001, doi: 10.1039/b104037p.

[4] Y. Matsushita, T. Ichimura, N. Ohba, S. Kumada, K. Sakeda, T. Suzuki, H. Tanibata, and T. Murata, “Recent progress on photoreactions in microreactors,” Pure Appl. Chem., vol. 79, no. 11, pp. 1959–1968, 2007, doi: 10.1351/pac200779111959.

[5] C. Sambiagio and T. Noël, “Flow Photochemistry: Shine Some Light on Those Tubes!,” Trends Chem., vol. 2, no. 2, pp. 92–106, 2020, doi: 10.1016/j.trechm.2019.09.003.

[6] Y. Özbakır, A. Jonáš, A. Kiraz, and C. Erkey, “A new type of microphotoreactor with integrated optofluidic waveguide based on solid-air nanoporous aerogels,” R. Soc. Open Sci., vol. 5, no. 11, 2018, doi: 10.1098/rsos.180802.

[7] J. S. Y. Chen, T. G. Euser, N. J. Farrer, P. J. Sadler, M. Scharrer, and P. S. J. Russell, “Photochemistry in photonic crystal fiber Nanoreactors,” Chem. - A Eur. J., vol. 16, no. 19, pp. 5607–5612, 2010, doi: 10.1002/chem.201000496.

[8] A. M. Cubillas, S. Unterkofler, T. G. Euser, B. J. M. Etzold, A. C. Jones, P. J. Sadler, P. Wasserscheid, and P. S. J. Russell, “Photonic crystal fibres for chemical sensing and photochemistry,” Chem. Soc. Rev., vol. 42, no. 22, pp. 8629–8648, 2013, doi: 10.1039/c3cs60128e.

[9] N. J. Turro, Modern molecular photochemistry. Sausalito: University Science Books,U.S., 1991.

[10] M. Oelgemöller and O. Shvydkiv, “Recent advances in microflow photochemistry,” Molecules, vol. 16, no. 9, pp. 7522–7550, 2011, doi: 10.3390/molecules16097522.

[11] M. Oelgemöller, “Highlights of Photochemical Reactions in Microflow Reactors,” Chem. Eng. Technol., vol. 35, no. 7, pp. 1144–1152, 2012, doi: 10.1002/ceat.201200009.

[12] R. Gorges, S. Meyer, and G. Kreisel, “Photocatalysis in microreactors,” J. Photochem. Photobiol. A Chem., vol. 167, no. 2–3, pp. 95–99, 2004, doi: 10.1016/j.jphotochem.2004.04.004.

[13] N. Wang, X. Zhang, Y. Wang, W. Yu, and H. L. W. Chan, “Microfluidic reactors for photocatalytic water purification,” Lab Chip, vol. 14, no. 6, pp. 1074–1082, 2014, doi: 10.1039/c3lc51233a.

[14] S. Parsons, Advanced Oxidation Processes for Water and Wastewater Treatment. Tunbridge Wells: IWA Publishing, 2004.

[15] E. Yousif and R. Haddad, “Photodegradation and photostabilization of polymers,

Page 46: The use of microreactors/optofluidics to study

45

especially polystyrene: Review,” Springerplus, vol. 2, no. 1, pp. 1–32, 2013, doi: 10.1186/2193-1801-2-398.

[16] A. Y. C. Lin and M. Reinhard, “Photodegradation of common environmental pharmaceuticals and estrogens in river water,” Environ. Toxicol. Chem., vol. 24, no. 6, pp. 1303–1309, 2005, doi: 10.1897/04-236R.1.

[17] D. G. Bradley and D. B. Min, “Singlet oxygen oxidation of foods,” Crit. Rev. Food Sci. Nutr., vol. 31, no. 3, pp. 211–236, 1992, doi: 10.1080/10408399209527570.

[18] D. B. Min and J. M. Boff, “Chemistry and reaction of singlet oxygen in foods,” Compr. Rev. Food Sci. Food Saf., vol. 1, no. 2, pp. 58–72, 2002, doi: 10.1111/j.1541-4337.2002.tb00007.x.

[19] E. Choe and D. B. Min, “Chemistry and Reactions of Reactive Oxygen Species in Foods,” J. Food Sci., vol. 70, no. 9, 2005.

[20] L. Li, N. Y. Gao, Y. Deng, J. J. Yao, K. J. Zhang, H. J. Li, D. Di Yin, H. S. Ou, and J. W. Guo, “Experimental and model comparisons of H2O2 assisted UV photodegradation of Microcystin-LR in simulated drinking water,” J. Zhejiang Univ. Sci. A, vol. 10, no. 11, pp. 1660–1669, 2009, doi: 10.1631/jzus.A0820642.

[21] R. Liu, Q. Jin, G. Tao, L. Shan, Y. Liu, and X. Wang, “LC-MS and UPLC-quadrupole time-of-flight MS for identification of photodegradation products of aflatoxin B1,” Chromatographia, vol. 71, no. 1–2, pp. 107–112, 2010, doi: 10.1365/s10337-009-1354-y.

[22] R. J. McQuitty, S. Unterkofler, T. G. Euser, P. S. J. Russell, and P. J. Sadler, “Rapid screening of photoactivatable metallodrugs: Photonic crystal fibre microflow reactor coupled to ESI mass spectrometry,” RSC Adv., vol. 7, no. 59, pp. 37340–37348, 2017, doi: 10.1039/c7ra06735f.

[23] Y. Özbakir, A. Jonas, A. Kiraz, and C. Erkey, “Aerogels for optofluidic waveguides,” Micromachines, vol. 8, no. 4, 2017, doi: 10.3390/mi8040098.

[24] T. Dallas and P. K. Dasgupta, “Light at the end of the tunnel: Recent analytical applications of liquid-core waveguides,” TrAC - Trends Anal. Chem., vol. 23, no. 5, pp. 385–392, 2004, doi: 10.1016/S0165-9936(04)00522-9.

[25] A. R. Hawkins and H. Schmidt, “Optofluidic waveguides: II. Fabrication and structures,” Microfluid. Nanofluidics, vol. 4, no. 1–2, pp. 17–32, 2008, doi: 10.1007/s10404-007-0194-z.

[26] A. M. Cubillas, M. Schmidt, M. Scharrer, T. G. Euser, B. J. M. Etzold, N. Taccardi, P. Wasserscheid, and P. S. J. Russell, “Ultra-low concentration monitoring of catalytic reactions in photonic crystal fiber,” Chem. - A Eur. J., vol. 18, no. 6, pp. 1586–1590, 2012, doi: 10.1002/chem.201102424.

[27] K. Kurokawa, K. Nakajima, K. Tsujikawa, T. Yamamoto, and K. Tajima, “Ultra-Wideband Transmission Over Low Loss PCF,” J. Light. Technol., vol. 27, no. 11, pp. 1653–1662, 2009, doi: 10.1109/JLT.2009.2014174.

[28] P. K. Dasgupta, Z. Genfa, S. K. Poruthoor, S. Caldwell, S. Dong, and S. Y. Liu, “High-sensitivity gas sensors based on gas-permeable liquid core waveguides and long-path absorbance detection,” Anal. Chem., vol. 70, no. 22, pp. 4661–4669, 1998, doi: 10.1021/ac980803t.

Page 47: The use of microreactors/optofluidics to study

46

[29] X. Cheng, Z. Zhang, and S. Tian, “A novel long path length absorbance spectroscopy for the determination of ultra trace organophosphorus pesticides in vegetables and fruits,” Spectrochim. Acta - Part A Mol. Biomol. Spectrosc., vol. 67, no. 5, pp. 1270–1275, 2007, doi: 10.1016/j.saa.2006.10.018.

[30] R. Manor, A. Datta, I. Ahmad, M. Holtz, S. Gangopadhyay, and T. Dallas, “Microfabrication and characterization of liquid core waveguide glass channels coated with teflon AF,” IEEE Sens. J., vol. 3, no. 6, pp. 687–692, 2003, doi: 10.1109/JSEN.2003.820342.

[31] K. Fujiwara and I. Seiji, “Application of waveguiding in solutions for absorption and fluorescence spectrometry,” trends Anal. Chem., vol. 10, no. 6, pp. 184–190, 1991.

[32] K. Fujiwara, I. Seiji, R.-E. Kojyo, H. Tsubota, and R. L. Carter, “Side-View Type of Waveguide Flow Cells for Fluorimetry as a Detector for Flow Injection Analysis of Lead,” Appl. Spectrosc., vol. 46, no. 6, pp. 1032–1039, 1992, doi: 10.1366/0003702924124286.

[33] J. Li, P. K. Dasgupta, and Z. Genfa, “Transversely illuminated liquid core waveguide based fluorescence detection: Fluorometric flow injection determination of aqueous ammonium/ammonia,” Talanta, vol. 50, no. 3, pp. 617–623, 1999, doi: 10.1016/S0039-9140(99)00148-4.

[34] P. K. Dasgupta, Z. Genfa, J. Li, C. B. Boring, S. Jambunathan, and R. Al-Horr, “Luminescence Detection with a Liquid Core Waveguide Purnendu,” Anal. Chem., vol. 71, no. 7, pp. 1400–1407, 1999, doi: 10.1021/ac981260q.

[35] R. J. Dijkstra, F. Ariese, C. Gooijer, and U. A. T. Brinkman, “Raman spectroscopy as a detection method for liquid-separation techniques,” Trends Anal. Chem., vol. 24, no. 4, pp. 304–323, 2005, doi: 10.1016/j.trac.2004.11.022.

[36] D. Yan, J. Popp, M. W. Pletz, and T. Frosch, “Fiber enhanced Raman sensing of levofloxacin by PCF bandgap-shifting into the visible range,” Anal. Methods, vol. 10, no. 6, pp. 586–592, 2017, doi: 10.1039/c7ay02398g.

[37] T. Frosch, D. Yan, and J. Popp, “Ultrasensitive fiber enhanced UV resonance Raman sensing of drugs,” Anal. Chem., vol. 85, no. 13, pp. 6264–6271, 2013, doi: 10.1021/ac400365f.

[38] Y. Tian, L. Zhang, J. Zuo, Z. Li, S. Gao, and G. Lu, “Raman sensitivity enhancement for aqueous absorbing sample using Teflon-AF 2400 liquid core optical fibre cell,” Anal. Chim. Acta, vol. 581, no. 1, pp. 154–158, 2007, doi: 10.1016/j.aca.2006.07.082.

[39] S. Tanikkul, J. Jakmunee, M. Rayanakorn, K. Grudpan, B. J. Marquardt, G. M. Gross, B. J. Prazen, L. W. Burgess, G. D. Christian, and R. E. Synovec, “Characterization and use of a Raman liquid-core waveguide sensor using preconcentration principles,” Talanta, vol. 59, no. 4, pp. 809–816, 2003, doi: 10.1016/S0039-9140(02)00623-9.

[40] R. J. Dijkstra, C. J. Slooten, A. Stortelder, J. B. Buijs, F. Ariese, U. A. T. Brinkman, and C. Gooijer, “Liquid-core waveguide technology for coupling column liquid chromatography and Raman spectroscopy,” J. Chromatogr. A, vol. 918, no. 1, pp. 25–36, 2001, doi: 10.1016/S0021-9673(01)00721-X.

[41] R. J. Dijkstra, A. N. Bader, G. P. Hoornweg, U. A. T. Brinkman, and C. Gooijer, “On-line coupling of column liquid chromatography and Raman spectroscopy using a liquid core waveguide,” Anal. Chem., vol. 71, no. 20, pp. 4575–4579, 1999, doi:

Page 48: The use of microreactors/optofluidics to study

47

10.1021/ac9902648.

[42] R. Altkorn, M. D. Malinsky, R. P. Van Duyne, and I. Koev, “Intensity considerations in liquid core optical fiber Raman spectroscopy,” Appl. Spectrosc., vol. 55, no. 4, pp. 373–381, 2001, doi: 10.1366/0003702011951939.

[43] M. Holtz, P. K. Dasgupta, and G. Zhang, “Small-volume Raman spectroscopy with a liquid core waveguide,” Anal. Chem., vol. 71, no. 14, pp. 2934–2938, 1999, doi: 10.1021/ac981261i.

[44] K. Hofstadler, R. Bauer, S. Novallc, and G. Heisler, “New Reactor Design for Photocatalytic Wastewater Treatment with TiO2 Immobilized on Fused-Silica Glass Fibers: Photomineralization of 4-Chlorophenol,” Environ. Sci. Technol., vol. 28, no. 4, pp. 670–674, 1994, doi: 10.1021/es00053a021.

[45] H. Schmidt and A. R. Hawkins, “Optofluidic waveguides: I. Concepts and implementations,” Microfluid. Nanofluidics, vol. 4, no. 1–2, pp. 3–16, 2008, doi: 10.1007/s10404-007-0199-7.

[46] Y. Özbakır, A. Jonáš, A. Kiraz, and C. Erkey, “Total internal reflection-based optofluidic waveguides fabricated in aerogels,” J. Sol-Gel Sci. Technol., vol. 84, no. 3, pp. 522–534, 2017, doi: 10.1007/s10971-017-4426-8.

[47] J. E. Saunders, C. Sanders, H. Chen, and H.-P. Loock, “Refractive indices of common solvents and solutions at 1550 nm,” Appl. Opt., vol. 55, no. 4, p. 947, 2016, doi: 10.1364/ao.55.000947.

[48] J. H. Lowry, “Optical characteristics of Teflon AF fluoroplastic materials,” Opt. Eng., vol. 31, no. 9, p. 1982, 1992, doi: 10.1117/12.59910.

[49] M. K. Yang, “Optical properties of Teflon® AF amorphous fluoropolymers,” J. Micro/Nanolithography, MEMS, MOEMS, vol. 7, no. 3, p. 033010, 2008, doi: 10.1117/1.2965541.

[50] R. H. Byrne and E. Kaltenbacher, “Use of liquid core waveguides for long pathlength absorbance spectroscopy: Principles and practice,” Limnol. Oceanogr., vol. 46, no. 3, pp. 740–742, 2001, doi: 10.4319/lo.2001.46.3.0740.

[51] R. Altkorn, I. Koev, R. P. Van Duyne, and M. Litorja, “Low-loss liquid-core optical fiber for low-refractive-index liquids: fabrication, characterization, and application in Raman spectroscopy,” Appl. Opt., vol. 36, no. 34, p. 8992, 1997, doi: 10.1364/ao.36.008992.

[52] Z. Wang, Y. Wang, W. J. Cai, and S. Y. Liu, “A long pathlength spectrophotometric pCO2 sensor using a gas-permeable liquid-core waveguide,” Talanta, vol. 57, no. 1, pp. 69–80, 2002, doi: 10.1016/S0039-9140(02)00008-5.

[53] M. R. Milani and P. K. Dasgupta, “Measurement of nitrogen dioxide and nitrous acid using gas-permeable liquid core waveguides,” Anal. Chim. Acta, vol. 431, no. 2, pp. 169–180, 2001, doi: 10.1016/S0003-2670(00)01331-3.

[54] A. Datta, I. Y. Eom, A. Dhar, P. Kuban, R. Manor, I. Ahmad, S. Gangopadhyay, T. Dallas, M. Holtz, H. Temkin, and P. K. Dasgupta, “Microfabrication and characterization of Teflon AF-coated liquid core waveguide channels in silicon,” IEEE Sens. J., vol. 3, no. 6, pp. 788–795, 2003, doi: 10.1109/JSEN.2003.820343.

[55] C. W. Wu and G. C. Gong, “Fabrication of PDMS-based nitrite sensors using teflon AF

Page 49: The use of microreactors/optofluidics to study

48

coating microchannels,” IEEE Sens. J., vol. 8, no. 5, pp. 465–469, 2008, doi: 10.1109/JSEN.2008.918201.

[56] S. H. Cho, J. Godin, and Y. H. Lo, “Optofluidic waveguides in Teflon AF-coated PDMS microfluidic channels,” IEEE Photonics Technol. Lett., vol. 21, no. 15, pp. 1057–1059, 2009, doi: 10.1109/LPT.2009.2022276.

[57] P. Dress, M. Belz, K. F. Klein, K. T. V. Grattan, and H. Franke, “Physical analysis of teflon coated capillary waveguides,” Sensors Actuators, B Chem., vol. 51, no. 1–3, pp. 278–284, 1998, doi: 10.1016/S0925-4005(98)00196-8.

[58] R. Altkorn, I. Koev, and A. Gottlieb, “Waveguide capillary cell for low-refractive-index liquids,” Appl. Spectrosc., vol. 51, no. 10, pp. 1554–1558, 1997, doi: 10.1366/0003702971939109.

[59] J. Z. Zhang, C. Kelble, and F. J. Millero, “Gas-segmented continuous flow analysis of iron in water with a long liquid waveguide capillary flow cell,” Anal. Chim. Acta, vol. 438, no. 1–2, pp. 49–57, 2001, doi: 10.1016/S0003-2670(01)01031-5.

[60] J. Z. Zhang and J. Chi, “Automated analysis of nanomolar concentrations of phosphate in natural waters with liquid waveguide,” Environ. Sci. Technol., vol. 36, no. 5, pp. 1048–1053, 2002, doi: 10.1021/es011094v.

[61] X. Fan and I. M. White, “Optofluidic microsystems for chemical and biological analysis,” Nat. Photonics, vol. 5, no. 10, pp. 591–597, 2011, doi: 10.1038/nphoton.2011.206.

[62] M. Brzozowski, M. O’Brien, S. V. Ley, and A. Polyzos, “Flow chemistry: Intelligent processing of gas-liquid transformations using a tube-in-tube reactor,” Acc. Chem. Res., vol. 48, no. 2, pp. 349–362, 2015, doi: 10.1021/ar500359m.

[63] M. O’Brien, N. Taylor, A. Polyzos, I. R. Baxendale, and S. V. Ley, “Hydrogenation in flow: Homogeneous and heterogeneous catalysis using Teflon AF-2400 to effect gas-liquid contact at elevated pressure,” Chem. Sci., vol. 2, no. 7, pp. 1250–1257, 2011, doi: 10.1039/c1sc00055a.

[64] A. Polyzos, M. O’Brien, T. P. Petersen, I. R. Baxendale, and S. V. Ley, “The Continuous-Flow Synthesis of Carboxylic Acids using CO2 in a Tube-In-Tube Gas Permeable Membrane Reactor,” Angew. Chemie - Int. Ed., vol. 50, no. 5, pp. 1190–1193, 2011, doi: 10.1002/anie.201006618.

[65] M. O’Brien, I. Baxendale, and S. Ley, “Flow Ozonolysis Using a Semipermeable Teflon AF-2400 Membrane,” Synfacts, vol. 2010, no. 10, pp. 1199–1199, 2010, doi: 10.1055/s-0030-1258661.

[66] S. Ponce, H. Christians, A. Drochner, and B. J. M. Etzold, “An Optical Microreactor Enabling In Situ Spectroscopy Combined with Fast Gas-Liquid Mass Transfer,” Chemie-Ingenieur-Technik, vol. 90, no. 11, pp. 1855–1863, 2018, doi: 10.1002/cite.201800061.

[67] S. Ponce, M. Trabold, A. Drochner, J. Albert, and B. J. M. Etzold, “Insights into the redox kinetics of vanadium substituted heteropoly acids through liquid core waveguide membrane microreactor studies,” Chem. Eng. J., vol. 369, no. March, pp. 443–450, 2019, doi: 10.1016/j.cej.2019.03.103.

[68] S. Ponce, H. Christians, J. Albert, A. Drochner, and B. J. M. Etzold, “A new concept for

Page 50: The use of microreactors/optofluidics to study

49

a microreactor with intense light / mater and gas / liquid interaction Time ( min ),” 2018, pp. 5–6.

[69] B. Yalizay, Y. Morova, K. Dincer, Y. Ozbakir, A. Jonas, C. Erkey, A. Kiraz, and S. Akturk, “Versatile liquid-core optofluidic waveguides fabricated in hydrophobic silica aerogels by femtosecond-laser ablation,” Opt. Mater. (Amst)., vol. 47, pp. 478–483, 2015, doi: 10.1016/j.optmat.2015.06.024.

[70] G. Eris, D. Sanli, Z. Ulker, S. E. Bozbag, A. Jonás, A. Kiraz, and C. Erkey, “Three-dimensional optofluidic waveguides in hydrophobic silica aerogels via supercritical fluid processing,” J. Supercrit. Fluids, vol. 73, pp. 28–33, 2013, doi: 10.1016/j.supflu.2012.11.001.

[71] L. Xiao and T. A. Birks, “Optofluidic microchannels in aerogel,” Opt. Lett., 2011, doi: 10.1364/ol.36.003275.

[72] B. Yalizay, Y. Morova, Y. Ozbakir, A. Jonas, C. Erkey, A. Kiraz, and S. Akturk, “Optofluidic waveguides written in hydrophobic silica aerogels with a femtosecond laser,” Integr. Opt. Devices, Mater. Technol. XIX, vol. 9365, p. 936518, 2015, doi: 10.1117/12.2077132.

[73] N. Hüsing and U. Schubert, “Aerogels—Airy Materials: Chemistry, Structure, and Properties,” Angew. Chemie Int. Ed., vol. 37, no. 1/2, pp. 22–45, 1998, doi: 10.1002/1521-3773(19980202)37:1/2<22::aid-anie22>3.3.co;2-9.

[74] A. C. Pierre and G. M. Pajonk, “Chemistry of aerogels and their applications,” Chem. Rev., vol. 102, no. 11, pp. 4243–4265, 2002, doi: 10.1021/cr0101306.

[75] A. Du, B. Zhou, Z. Zhang, and J. Shen, “A special material or a new state of matter: A review and reconsideration of the aerogel,” Materials (Basel)., vol. 6, no. 3, pp. 941–968, 2013, doi: 10.3390/ma6030941.

[76] A. Soleimani Dorcheh and M. H. Abbasi, “Silica aerogel; synthesis, properties and characterization,” J. Mater. Process. Technol., vol. 199, no. 1, pp. 10–26, 2008, doi: 10.1016/j.jmatprotec.2007.10.060.

[77] Q. Chen, D. Long, L. Chen, X. Liu, X. Liang, W. Qiao, and L. Ling, “Synthesis of ultrahigh-pore-volume carbon aerogels through a ‘reinforced-concrete’ modified sol-gel process,” J. Non. Cryst. Solids, vol. 357, no. 1, pp. 232–235, 2011, doi: 10.1016/j.jnoncrysol.2010.09.049.

[78] S. A. Al-Muhtaseb and J. A. Ritter, “Preparation and properties of resorcinol-formaldehyde organic and carbon gels,” Adv. Mater., vol. 15, no. 2, pp. 101–114, 2003, doi: 10.1002/adma.200390020.

[79] S. Zhao, W. J. Malfait, A. Demilecamps, Y. Zhang, S. Brunner, L. Huber, P. Tingaut, A. Rigacci, T. Budtova, and M. M. Koebel, “Strong, Thermally Superinsulating Biopolymer-Silica Aerogel Hybrids by Cogelation of Silicic Acid with Pectin,” Angew. Chemie - Int. Ed., vol. 54, no. 48, pp. 14282–14286, 2015, doi: 10.1002/anie.201507328.

[80] B. T. Mekonnen, M. Ragothaman, C. Kalirajan, and T. Palanisamy, “Conducting collagen-polypyrrole hybrid aerogels made from animal skin waste,” RSC Adv., vol. 6, no. 67, pp. 63071–63077, 2016, doi: 10.1039/c6ra08876g.

[81] L. Xiao, M. D. W. Grogan, S. G. Leon-Saval, R. Williams, R. England, W. J. Wadsworth,

Page 51: The use of microreactors/optofluidics to study

50

and T. A. Birks, “Tapered fibres embedded in silica aerogel,” Opt. InfoBase Conf. Pap., vol. 34, no. 18, pp. 2724–2726, 2009, doi: 10.1364/ol.34.002724.

[82] J. Fricke, “Aerogels - highly tenuous solids with fascinating properties,” J. Non. Cryst. Solids, vol. 100, no. 1–3, pp. 169–173, 1988, doi: 10.1016/0022-3093(88)90014-2.

[83] Y. Özbakir, A. Jonáš, A. Kiraz, and C. Erkey, “Photocatalytic transformation in aerogel-based optofluidic microreactors,” Proc. Spie, vol. 10535, no. February 2018, p. 84, 2018, doi: 10.1117/12.2292309.

[84] P. Russell, “Photonic crystal fibers,” Science (80-. )., vol. 299, no. 5605, pp. 358–362, 2003, doi: 10.1126/science.1079280.

[85] X. Tao, Wearable photonics based on integrative polymeric photonic fibres. Woodhead Publishing Limited, 2005.

[86] R. Yew, S. K. Karuturi, H. H. Tan, and C. Jagadish, Nanostructured Photoelectrodes via Template-Assisted Fabrication, 1st ed., vol. 97. Elsevier Inc., 2017.

[87] B. Gandhi, A. K. Skukla, and G. . Pandey, “Design of 1 x 4 All Optical Splitter Based on 2D Photonic Crystal,” Springer Proc. Phys., vol. 194, pp. 551–557, 2017, doi: 10.1007/978-981-10-3908-9_69.

[88] H. Qu and M. Skorobogatiy, “Liquid-core low-refractive-index-contrast Bragg fiber sensor,” Appl. Phys. Lett., vol. 98, no. 20, 2011, doi: 10.1063/1.3592758.

[89] P. Yeh and A. Yariv, “Bragg reflection waveguides,” Opt. Commun., vol. 19, no. 3, pp. 427–430, 1976, doi: 10.1016/0030-4018(76)90115-2.

[90] P. Yeh, A. Yariv, and C. Hong, “Electromagnetic propagation in periodic stratified media. I. General theory,” J. Opt. Soc. Am., vol. 67, no. 4, pp. 423–438, 1977.

[91] P. Yeh, A. Yariv, and E. Maron, “Theory of Bragg Fiber.,” J Opt Soc Am, vol. 68, no. 9, pp. 1196–1201, 1978, doi: 10.1364/JOSA.68.001196.

[92] K. J. Rowland, S. Afshar, A. Stolyarov, Y. Fink, and T. M. Monro, “Bragg waveguides with low-index liquid cores,” Opt. Express, vol. 20, no. 1, p. 48, 2012, doi: 10.1364/oe.20.000048.

[93] K. Kuriki, O. Shapira, S. D. Hart, G. Benoit, Y. Kuriki, J. F. Viens, M. Bayindir, J. D. Joannopoulos, and Y. Fink, “Hollow multilayer photonic bandgap fibers for NIR applications,” Opt. Express, vol. 12, no. 8, p. 1510, 2004, doi: 10.1364/opex.12.001510.

[94] B. Temelkuran, S. D. Hart, G. Benoit, J. D. Joannopoulos, and Y. Fink, “Wavelength-scalable hollow optical fibres with large photonic bandgaps for CO2 laser transmission,” Nature, vol. 420, no. 6916, pp. 650–653, 2002, doi: 10.1038/nature01275.

[95] K. Milenko, D. J. J. Hu, P. P. Shum, and T. R. Wolinski, “Hollow-core Bragg fiber for bio-sensing applications,” Acta Phys. Pol. A, vol. 118, no. 6, pp. 1205–1208, 2010, doi: 10.12693/APhysPolA.118.1205.

[96] Z. Liu and H.-Y. Tam, “Fabrication and Sensing Applications of Special Microstructured Optical Fibers,” Sel. Top. Opt. Fiber Technol. Appl., 2018, doi: 10.5772/intechopen.70755.

[97] G. Antonopoulos, F. Benabid, T. A. Birks, D. M. Bird, J. C. Knight, and P. S. J. Russell,

Page 52: The use of microreactors/optofluidics to study

51

“Experimental demonstration of the frequency shift of bandgaps in photonic crystal fibers due to refractive index scaling,” Opt. Express, vol. 14, no. 7, p. 3000, 2006, doi: 10.1364/oe.14.003000.

[98] G. A. Mahdiraji, D. M. Chow, S. R. Sandoghchi, F. Amirkhan, E. Dermosesian, K. S. Yeo, Z. Kakaei, M. Ghomeishi, S. Y. Poh, S. Y. Gang, and F. R. M. Adikan, “Challenges and solutions in fabrication of silica-based photonic crystal fibers: An experimental study,” Fiber Integr. Opt., vol. 33, no. 1–2, pp. 85–104, 2014, doi: 10.1080/01468030.2013.879680.

[99] R. F. Cregan, B. J. Mangan, J. C. Knight, T. A. Birks, P. S. J. Russell, P. J. Roberts, and D. C. Allan, “Single-Mode Photonic Band Gap Guidance of Light in Air,” Science (80-. )., vol. 285, no. 5433, pp. 1537–1539, 1999, doi: DOI: 10.1126/science.285.5433.1537.

[100] J. M. Fini, “Microstructure fibres for optical sensing in gases and liquids,” Meas. Sci. Technol., vol. 15, no. 6, pp. 1120–1128, 2004, doi: 10.1088/0957-0233/15/6/011.

[101] A. S. Webb, “Suspended-core holey fiber for evanescent-field sensing,” Opt. Eng., vol. 46, no. 1, p. 010503, 2007, doi: 10.1117/1.2430505.

[102] T. M. Monro, S. Warren-Smith, E. P. Schartner, A. Franois, S. Heng, H. Ebendorff-Heidepriem, and S. Afshar, “Sensing with suspended-core optical fibers,” Opt. Fiber Technol., vol. 16, no. 6, pp. 343–356, 2010, doi: 10.1016/j.yofte.2010.09.010.

[103] T. G. Euser, J. S. Y. Chen, M. Scharrer, P. S. J. Russell, N. J. Farrer, and P. J. Sadler, “Quantitative broadband chemical sensing in air-suspended solid-core fibers,” J. Appl. Phys., vol. 103, no. 10, 2008, doi: 10.1063/1.2924408.

[104] F. Benabid, “Hollow-core photonic bandgap fibre: New light guidance for new science and technology,” Philos. Trans. R. Soc. A Math. Phys. Eng. Sci., vol. 364, no. 1849, pp. 3439–3462, 2006, doi: 10.1098/rsta.2006.1908.

[105] P. J. Roberts, F. Couny, H. Sabert, B. J. Mangan, D. P. Williams, L. Farr, M. W. Mason, A. Tomlinson, T. A. Birks, J. C. Knight, and P. St. J. Russell, “Ultimate low loss of hollow-core photonic crystal fibres,” Opt. Express, vol. 13, no. 1, p. 236, 2005, doi: 10.1364/opex.13.000236.

[106] F. Benabid, J. C. Knight, G. Antonopoulos, and P. S. J. Russell, “Stimulated Raman scattering in hydrogen-filled hollow-core photonic crystal fiber,” Science (80-. )., vol. 298, no. 5592, pp. 399–402, 2002, doi: 10.1126/science.1076408.

[107] A. Hartung, J. Kobelke, A. Schwuchow, J. Bierlich, J. Popp, M. A. Schmidt, and T. Frosch, “Low-loss single-mode guidance in large-core antiresonant hollow-core fibers,” Opt. Lett., vol. 40, no. 14, p. 3432, 2015, doi: 10.1364/ol.40.003432.

[108] F. Couny, F. Benabid, and P. S. Light, “Large-pitch kagome-structured hollow-core photonic crystal fiber,” Opt. Lett., vol. 31, no. 24, p. 3574, 2006, doi: 10.1364/ol.31.003574.

[109] G. O. S. Williams, J. S. Y. Chen, T. G. Euser, P. S. J. Russell, and A. C. Jones, “Photonic crystal fibre as an optofluidic reactor for the measurement of photochemical kinetics with sub-picomole sensitivity,” Lab Chip, vol. 12, no. 18, pp. 3356–3361, 2012, doi: 10.1039/c2lc40062f.

[110] S. Yiou, P. Delaye, A. Rouvie, J. Chinaud, R. Frey, G. Roosen, P. Viale, S. Février, P. Roy,

Page 53: The use of microreactors/optofluidics to study

52

J.-L. Auguste, and J.-M. Blondy, “Stimulated Raman scattering in an ethanol core microstructured optical fiber,” Opt. Express, vol. 13, no. 12, p. 4786, 2005, doi: 10.1364/opex.13.004786.

[111] K. Nielsen, D. Noordegraaf, T. Sørensen, A. Bjarklev, and T. P. Hansen, “Selective filling of photonic crystal fibres,” J. Opt. A Pure Appl. Opt., vol. 7, no. 8, pp. L13–L20, 2005, doi: 10.1088/1464-4258/7/8/l02.

[112] L. Xiao, W. Jin, M. S. Demokan, H. L. Ho, Y. L. Hoo, and C. Zhao, “Fabrication of selective injection microstructured optical fibers with a conventional fusion splicer,” Opt. Express, vol. 13, no. 22, p. 9014, 2005, doi: 10.1364/opex.13.009014.

[113] F. M. Cox, A. Argyros, and M. C. J. Large, “Liquid-filled hollow core microstructured polymer optical fiber,” Opt. Express, vol. 14, no. 9, p. 4135, 2006, doi: 10.1364/oe.14.004135.

[114] S. M. Atiqullah, A. Palit, M. I. Reja, J. Akhtar, S. Fatema, and R. Absar, “Detection of harmful food additives using highly sensitive photonic crystal fiber,” Sens. Bio-Sensing Res., vol. 23, no. March, p. 100275, 2019, doi: 10.1016/j.sbsr.2019.100275.

[115] Y. Sun, X. Yu, T. N. Nam, P. Shum, and C. K. Yien, “Long path-length axial absorption detection in photonic crystal fiber,” Anal. Chem., vol. 80, no. 11, pp. 4220–4224, 2008, doi: 10.1021/ac800417n.

[116] G. O. S. Williams, T. G. Euser, P. S. J. Russell, and A. C. Jones, “Spectrofluorimetry with attomole sensitivity in photonic crystal fibres,” Methods Appl. Fluoresc., vol. 1, no. 1, 2013, doi: 10.1088/2050-6120/1/1/015003.

[117] M. Azkune, T. Frosch, E. Arrospide, G. Aldabaldetreku, I. Bikandi, J. Zubia, J. Popp, and T. Frosch, “Liquid-Core Microstructured Polymer Optical Fiber as Fiber-Enhanced Raman Spectroscopy Probe for Glucose Sensing,” J. Light. Technol., vol. 37, no. 13, pp. 2981–2988, 2019, doi: 10.1109/JLT.2019.2908447.

[118] S. Wolf, T. Frosch, J. Popp, M. W. Pletz, and T. Frosch, “Highly sensitive detection of the antibiotic ciprofloxacin by means of fiber enhanced Raman spectroscopy,” Molecules, vol. 24, no. 24, 2019, doi: 10.3390/molecules24244512.

[119] D. Yan, T. Frosch, J. Kobelke, J. Bierlich, J. Popp, M. W. Pletz, and T. Frosch, “Fiber-Enhanced Raman Sensing of Cefuroxime in Human Urine,” Anal. Chem., vol. 90, no. 22, pp. 13243–13248, 2018, doi: 10.1021/acs.analchem.8b01355.

[120] S. Unterkofler, R. J. McQuitty, T. G. Euser, N. J. Farrer, P. J. Sadler, and P. S. J. Russell, “Microfluidic integration of photonic crystal fibers for online photochemical reaction analysis,” Opt. Lett., vol. 37, no. 11, p. 1952, 2012, doi: 10.1364/ol.37.001952.

[121] A. M. Cubillas, M. Schmidt, T. G. Euser, N. Taccardi, S. Unterkofler, P. S. J. Russell, P. Wasserscheid, and B. J. M. Etzold, “In Situ Heterogeneous Catalysis Monitoring in a Hollow-Core Photonic Crystal Fiber Microflow Reactor,” Adv. Mater. Interfaces, vol. 1, no. 5, pp. 1–5, 2014, doi: 10.1002/admi.201300093.

[122] H. A. MacLeod, Thin-Film Optical Filters. Tucson, Arizona, USA, 2001.

[123] M. Mann, U. Trutschel, C. Wachter, L. Leine, and F. Lederer, “Directional coupler based on an antiresonant reflecting optical waveguide,” Opt. Lett., vol. 16, no. 11, pp. 805–807, 1991.

[124] G. Testa, G. Persichetti, and R. Bernini, “Liquid core ARROW waveguides: A promising

Page 54: The use of microreactors/optofluidics to study

53

photonic structure for integrated optofluidic microsensors,” Micromachines, vol. 7, no. 3, 2016, doi: 10.3390/mi7030047.

[125] M. A. Duguay, Y. Kokubun, T. L. Koch, and L. Pfeiffer, “Antiresonant reflecting optical waveguides in SiO2-Si multilayer structures,” Appl. Phys. Lett., vol. 49, no. 1, pp. 13–15, 1986, doi: 10.1063/1.97085.

[126] T. L. Koch, U. Koren, G. D. Boyd, P. J. Corvini, and M. A. Duguay, “Antiresonant reflecting optical waveguides for III-V integrated optics,” Electron. Lett., vol. 23, no. 5, pp. 244–245, 1987.

[127] R. Bernini, G. Testa, L. Zeni, and P. M. Sarro, “A 2 × 2 optofluidic multimode interference coupler,” IEEE J. Sel. Top. Quantum Electron., vol. 15, no. 5, pp. 1478–1484, 2009, doi: 10.1109/JSTQE.2009.2020996.

[128] R. Antonello and R. Oboe, “MEMS Gyroscopes for Consumers and Industrial Applications,” Microsensors, no. June 2011, 2011, doi: 10.5772/17689.

[129] R. Bernini, S. Campopiano, and L. Zeni, “Silicon micromachined hollow optical waveguides for sensing applications,” IEEE J. Sel. Top. Quantum Electron., vol. 8, no. 1, pp. 106–110, 2002, doi: 10.1109/2944.991405.

[130] D. Yin, D. W. Deamer, H. Schmidt, J. P. Barber, and A. R. Hawkins, “Integrated optical waveguides with liquid cores,” Appl. Phys. Lett., vol. 85, no. 16, pp. 3477–3479, 2004, doi: 10.1063/1.1807966.

[131] A. H. Gollub, D. O. Carvalho, T. C. Paiva, and M. I. Alayo, “Hollow core ARROW waveguides fabricated with SiOxNy films deposited at low temperatures,” Phys. Status Solidi Curr. Top. Solid State Phys., vol. 7, no. 3–4, pp. 964–967, 2010, doi: 10.1002/pssc.200982889.

[132] R. Bernini, E. De Nuccio, A. Minardo, L. Zeni, and P. M. Sarro, “Integrated silicon optical sensors based on hollow core waveguide,” Silicon Photonics II, vol. 6477, no. May 2010, p. 647714, 2007, doi: 10.1117/12.700410.

[133] D. Yin, D. W. Deamer, H. Schmidt, J. P. Barber, and A. R. Hawkins, “Single-molecule detection sensitivity using planar integrated optics on a chip,” Opt. Lett., vol. 31, no. 14, p. 2136, 2006, doi: 10.1364/ol.31.002136.

[134] D. Yin, J. P. Barber, A. R. Hawkins, and H. Schmidt, “Highly efficient fluorescence detection in picoliter volume liquid-core waveguides,” Appl. Phys. Lett., vol. 87, no. 21, pp. 1–3, 2005, doi: 10.1063/1.2135378.

[135] P. Measor, L. Seballos, D. Yin, J. P. Barber, A. R. Hawkins, J. Zang, and H. Schmidt, “Integrated ARROW Waveguide for Molecule Specific Surface-enhanced Raman Sensing,” IEEE, no. 6, pp. 181–182, 2006.

[136] S. Campopiano, R. Bernini, L. Zeni, and P. M. Sarro, “Microfluidic sensor based on integrated optical hollow waveguides,” Opt. Lett., vol. 29, no. 16, p. 1894, 2004, doi: 10.1364/ol.29.001894.

[137] C. Grillet, P. Domachuk, V. Ta’Eed, E. Mägi, J. A. Bolger, and B. J. Eggleton, “Compact tunable microfluidic interferometer,” Opt. Express, vol. 12, no. 22, pp. 5440–5447, 2004.

[138] A. Nitkowski, A. Baeumner, and M. Lipson, “On-chip spectrophotometry for bioanalysis using microring resonators,” Biomed. Opt. Express, vol. 2, no. 2, p. 271,

Page 55: The use of microreactors/optofluidics to study

54

2011, doi: 10.1364/boe.2.000271.

[139] F. Prieto, B. Sepúlveda, A. Calle, A. Llobera, C. Domínguez, A. Abad, A. Montoya, and L. M. Lechuga, “An integrated optical interferometric nanodevice based on silicon technology for biosensor applications,” Nanotechnology, vol. 14, no. 8, pp. 907–912, 2003, doi: 10.1088/0957-4484/14/8/312.

[140] L. K. Chin, A. Q. Liu, Y. C. Soh, C. S. Lim, and C. L. Lin, “A reconfigurable optofluidic Michelson interferometer using tunable droplet grating,” Lab Chip, vol. 10, no. 8, pp. 1072–1078, 2010, doi: 10.1039/b920412a.

[141] P. Domachuk, C. Grillet, V. Ta’Eed, E. Mägi, J. Bolger, B. J. Eggleton, L. E. Rodd, and J. Cooper-White, “Microfluidic interferometer,” Appl. Phys. Lett., vol. 86, no. 2, 2005, doi: 10.1063/1.1849415.

[142] D. Ozcelik, J. W. Parks, T. A. Wall, M. A. Stott, H. Cai, J. W. Parks, A. R. Hawkins, and H. Schmidt, “Optofluidic wavelength division multiplexing for single-virus detection,” Proc. Natl. Acad. Sci. U. S. A., vol. 112, no. 42, pp. 12933–12937, 2015, doi: 10.1073/pnas.1511921112.

[143] F. Yu, K. Yamamoto, X. Piao, and S. Yokoyama, “Multimode interference waveguide switch of electro-optic polymer with tapered access waveguides,” Phys. Procedia, vol. 14, pp. 25–28, 2011, doi: 10.1016/j.phpro.2011.05.006.

[144] G. Testa, C. Collini, L. Lorenzelli, and R. Bernini, “Planar silicon-polydimethylsiloxane optofluidic ring resonator sensors,” IEEE Photonics Technol. Lett., vol. 28, no. 2, pp. 155–158, 2016, doi: 10.1109/LPT.2015.2487680.

[145] G. Testa, Y. Huang, P. M. Sarro, L. Zeni, and R. Bernini, “Integrated silicon optofluidic ring resonator,” Appl. Phys. Lett., vol. 97, no. 13, 2010, doi: 10.1063/1.3496027.

[146] R. Bernini, G. Testa, L. Zeni, and P. M. Sarro, “Integrated optofluidic Mach-Zehnder interferometer based on liquid core waveguides,” Appl. Phys. Lett., vol. 93, no. 1, pp. 1–4, 2008, doi: 10.1063/1.2957031.

[147] G. Testa, Y. Huang, P. M. Sarro, L. Zeni, and R. Bernini, “High-visibility optofluidic Mach–Zehnder interferometer,” Opt. Lett., vol. 35, no. 10, p. 1584, 2010, doi: 10.1364/ol.35.001584.