Sie sind auf Seite 1von 32

Metal clad symmetric waveguides: properties and their use for the determination of dc Kerr coefficients of organic films.

Jonathan Cornish Project report 2011-2012 Supervisors: Marek Szablewski and Graham Cross.

Abstract
The properties of fabricated metal clad symmetric waveguides are examined in terms of the variance of structure, cover layer thickness, film thickness and film thickness uniformity. The materials used as film layers are PMMA (polymethylmethacrylate) doped with MOR2 and IPDI (polydicyanovinyl isophoronediisocyanate). Silver is used as the cover and cladding layers. Glass is used as the base layer and for coupling light from in one structure type (base coupled), whilst air is used for coupling light from in the other structure type (free space coupled). Analysis of their ATR spectra at 632.8nm is carried out, allowing the achievement of up to 83.40.1% light coupling, single mode waveguiding, film height uniformity analysis, and dc Kerr effect analysis. Free space coupled structures with silver cover layers of ~24nm are found to be of most use for dc Kerr anaylsis. Microscopic flaws in film layer quality restrict dc Kerr analysis to only one of many fabricated waveguides. Upper bounds for to and and therefore a range for the dc Kerr coefficient are found for 0.8050.006% w/w doped MOR2 in isotropic PMMA. These bounds ) ) are ( and ( for and respectively, leading to a range for the dc Kerr coefficient of ( ) . This range implies that doping MOR2 into isotropic PMMA at 0.8050.006% w/w does not significantly change the dc Kerr coefficient compared to undoped isotropic PMMA.

Acknowledgements
I would very much like to thank Dr. Szablewski for being always supportive and helpful with regard to carrying out the project, but more so for just being easy to talk to and having a good sense of humour. I would like to thank Dr. Cross for always being available for discussion with regard to the operation of waveguides, and again for having a good sense of humour. I thank the lab technicians as they always made borrowing equipment easy, and especially Duncan McCallum, for his constant help with regard to making my experiments safe. My housemates were supportive as always, and were always open to letting me talk in detail about how my project was going, while even at points seeming genuinely interested, for these reasons and more, I thank them.

Contents
1 Introduction 2 Theory 2.1 Waveguide theory 2.1.1 Ray optics model 2.1.2 Effective refractive index and propagation constant 2.1.3 EM field model 2.2 Waveguide coupling techniques 2.3 NLO origins and effects 2.4 Organic NLO materials 3 Experimental 3.1 Fabrication 3.1.1 Preparation of solutions 3.1.2 Waveguide structures and cleaning 3.1.3 Deposition of silver layers 3.1.4 Formation of thin films 3.2 ATR technique 3.3 Film thickness measurements 3.4 Absorbance spectra 3.5 dc Kerr effect measurements 4 Results and discussion 4.1 Absorbance spectra 4.2 Unsuccessful waveguides 4.3 Successful waveguides 4.4 Film thickness measurements and code generated ATRs 4.5 Variance of structure type 4.6 Variance of cover layer thickness 4.7 Single mode waveguides 4.8 dc Kerr effect measurements 5 Conclusion

1. Introduction
Telecommunications sees the widespread commercial use of waveguides. This is due to the low loss with which optical light can transfer information over long distances compared to electronic signals in copper wire [1]. Developments into increasing bandwidth for fibre-optic communication were needed and were continuously being made up until relatively recently [2]. Electro-optics has taken over research with regard to waveguides as it found its home in the advent of laser technology. Being able to modulate the frequency and/or phase of laser light is essential for mode locking and Q-switching in lasers, which is bought about with a NLO (non linear optical) material, with a waveguide used to couple the laser light to it[3,4]. Due to the high sensitivity of metal clad waveguide mode angles and intensities to anomalies in the film or cover layer quality, single mode waveguides can also be used as biosensing devices [5]. Waveguides trap light to propagate along one direction within the guiding layer. Ways to couple light to a waveguide include end fire coupling[6], tapered film coupling[7], grating coupling[8] and prism coupling[9]. However in 2003 using metal clad symmetric waveguides, a new method to couple light to a waveguide from free space was investigated[10]. This report looks to experimentally explore the coupling of such metal clad waveguides whilst also briefly investigating the theory that allows them to operate. The experimental procedure for analysing waveguides will be to obtain their Attenuated-Total-Reflectance spectra (ATR) [adam], as this method allows a very sensitive measure of the mode angles of the waveguides. ATR spectra allow experimental analysis for coupling by varying different properties of the waveguides. This report will analyse the variance cover layer thickness, film layer thickness and structure type on the ATR spectra the waveguides. Through the variance of these different parameters, a single mode waveguide will be attempted to be fabricated and coupling efficiency will be optimised. ATR spectra can further be used for sensitive thickness measurements of the film layer [balthazar], but are mostly for the investigation of NLO effects [chrome]. Once the ATR spectra of fabricated waveguides have been obtained, they will be used for the measurement of the NLO properties of the guiding materials. The materials used will be organics as these types of materials have typically large NLO susceptibilities, fast NLO response times and can be spin-coated allowing ease of fabrication [11]. In the literature organic materials used in conjunction with waveguiding can be seen in electro-optic devices [dandilion], but also in other devices such as for lighting [excelsior] and for photovoltaic cells [flutter]. The NLO effect investigated in this project will be the dc Kerr effect, a third order NLO effect. Although typically third order effects are of little importance for consideration of materials for integration into devices, an appreciation of the third order polarisability of a material allows a more discerning approach to appropriate material choice [tyrant].

2. Theory
4

2.1 Waveguide theory The different types of waveguide differ mostly on how they couple light to the film layer; not in the fundamentals of how they operate[11]. Waveguides confine visible (and near visible) electromagnetic waves to propagate in one direction, with the electric and magnetic fields confined to oscillate inside the film layer of the waveguide. EM waves propagating with no magnetic field in the propagation direction are labelled transverse magnetic (TM) modes, whilst those that propagate with no electric field in the propagation direction are called transverse electric (TE) modes. This project is concerned with general slab waveguides, and as such will discuss the guiding theory behind these waveguides but it should be noted that this theory can be applied, without much deviation, to any other waveguide type. Slab waveguides consist of three layers, a cover layer, a film layer and a substrate layer. The light is coupled by a number of methods from outside the waveguide through the cover layer and into the film layer, which then guides the light along it. There are two ways to mathematically analyse the process of waveguiding. One is to treat light as rays and to solve the problem in terms of optical geometry. The second is to consider light as EM fields and therefore solve Maxwells equations inside the waveguide. 2.1.1 Ray optics model Slab waveguides, as can be seen in Fig. 1, operate by repeated total internal reflection at the film-cover and film-substrate layer boundaries respectively. Exceeding the critical angle is not the only requirement for waveguiding to occur. Only waves that constructively interfere with their repeated reflections can waveguide. Only certain angles of incidence allow this constructive interference; these are called the waveguide modes. Fig. 2 shows constructively interfering waves being guided without loss. From Fig. 2 it can be seen that is subject to certain conditions if it is to be an allowed mode of the waveguide. After two reflections, i.e. from the point to in figure 2, it is clear that the two rays must be in phase with one another so that constructive interference can occur. This implies: ( )

where is the mode order ( ), is the phase change between the ray travelling from points a and b, and is given by [12]:

Figure 2.1.11: A diagram to show the basic symmetric slab waveguide, showing repeated total internal reflection at both film layer boundaries, with i having to be greater than C for this to occur. The waveguide is symmetric, meaning that the cover layer and the substrate layer have the same value for permittivity. The coordinate system used throughout this report is that quoted here; propagation along , waveguide thickness along and width of device along . This diagram was self-made.

Figure 2.1.12 [5]: A diagram showing a ray optics model of the operation of a waveguide. Only rays that constructively interfere along the waveguide are allowed modes of the waveguide. and are the phase changes induced in the ray at reflection. This project only considers symmetric waveguides, such that .

where

is the thickness of the film, The phase shift,

and

is the refractive index of the film.

, can be found to be [12]:

[(

with being the refractive index of the cover layer and being an indicator for whether TE or TM modes are being considered, with =0 for TE and =1 for TM. Subbing in 2.1.13 and 2.1.12 into 2.1.11 leads to the full transcendental equation:

[(

) ( )

from which it can now readily be seen how the mode angle, , and . 2.1.2 Effective refractive index and propagation constant

depends only on

The discrete angular modes of a waveguide govern the phase velocity at which the light within the waveguide travels. This can be seen in equation 1.4, as the term acts as a refractive index that is dependent on mode angle. This term is called the effective refractive index of a mode, , and is defined thus: ( where is the propagation constant assigned to each mode, and is defined as: ( where is the phase velocity of the light guided by a particular mode. ) )

2.1.3 EM field model The problem of a waveguide can be approached using Maxwells equations. Assuming no free charges exist within the waveguide system, one can use Maxwells equations to arrive at the wave equations for the electric field and magnetic field of light propagating through the waveguide: ( ( ) )

and where is assumed to be unity for these materials as they are not appreciably magnetic. Using the same coordinate system seen in figure 2.1.11, it can be safely assumed that for in a slab waveguide, when the electric field (equation 2.1.31) or magnetic field (equation 2.1.32) is polarised along , then: ( ) ( )

Equations 2.1.33 and 2.1.34 lead to Maxwells equations being divided into two sets; one in which there is only an E-field along and H-fields along and (TE), and one in which there is only an H-field along and E-fields along and (TM).
7

Figure 2.1.31 [5] (modified): The E field profile for the TE modes m=0 and m=1 of a metal clad waveguide configuration. TM modes exhibit similar behaviour, only that the H field has a larger field in the substrate and cover layers, and thus have a slightly different shape to the E field profile. Substituting in a plane wave solution to solve the wave equations under these conditions yields sinusoidal solutions for (TE) or (TM) along in the film layer and exponentially decaying functions for the field in the cover and substrate layers, as seen in figure 2.1.31. Applying the boundary conditions that phase, amplitude and the gradient of the field are conserved across the boundaries between layers leads to eigenvalue equations for symmetric and anti-symmetric TE and TM modes. Applying the condition that the wave propagates in phase between two total internal reflections leads to allowing the rearrangement of any of the eigenvalues equations to be of the form [jackhammer]: (( ( ( )

where

) and

) . Note that this equation is exactly

the same as equation (2.1.14). 2.2 Waveguide coupling techniques As mentioned previously, there are several ways to couple light to a waveguide. The prism coupling technique is most similar to the coupling techniques investigated in this report, so it is useful to investigate for analogies to be drawn, and for the power of the free space coupling technique to be seen. The technique setup can be seen in figure 2.21. The prism coupling method relies on the evanescent wave created in the air gap at the point where the incident light ray hits the edge of the prism. This wave allows the light to be transferred to the guiding layer and thus be coupled to the waveguide. The air gap then acts as a cover layer of the guiding layer where total internal reflection can repeatedly occur. Careful
8

Figure 2.21 [10]: A diagram of the prism coupling method, where n3 is the refractive index of the prism and 3 is the incident angle of the light to the guiding layer of the waveguide. control of the air gap thickness allows for maximum light coupling from prism to guiding layer to be achieved. From solving Maxwells equations for this system, it is seen that only modes with effective refractive index less than the refractive index of the prism are allowed: ( )

Another limit is that the effective refractive index of modes is that they must lead to solutions of Maxwells equations where the light propagates in the film layer but is evanescent in the air gap and substrate layer [12], with the allowed range being: ( )

In analogy to the prism coupling layout we can see the metal clad symmetric waveguide setup in figure 2.22. When discussing light propagation properties of the various layers of the metal clad symmetric waveguide, it is necessary for theory relating to the refractive index of a material to be examined further. From equations 2.1.31 or 2.1.32 it can be seen that ( )

However, equation 2.1.35 contains , not , so considering with regard to propagation properties instead of allows a more thorough approach. When considering it must be noted that it has a real and complex part, and therefore so does the refractive index: ( ) ( )

where is the real part and is the complex part of the relative permittivity, and is the real part of the refractive index and is the extinction coefficient. The complex parts of and must be considered in this project due to metals tending to have | |>| |.

Figure 2.22: This diagram shows the free space coupling arrangement with the permittivity of each layer shown. It is analogous to figure 2.21, apart from instead of an actual prism, air acts as the prism layer. The silver cover layer is analogous to the air gap layer in figure 2.21, with the guiding and substrate layers being identical, other than that the substrate and the guiding layer have the same permittivity 2. This diagram was self made. Although the real and complex part of are functions of the wavelength of the propagating wave, the only wavelength considered in this project is that of a HeNe laser (632.8nm). For silver at the HeNe wavelength, and [micheal]. Substituting in these values into equation 2.24 yields and . Numerically solving equation 2.1.35 for as negative for the cover and substrate layers leads to a much broader range of allowed [10]: ( )

This result is significant because it allows the refractive index of the film investigated to be arbitrarily small. It also allows the study of modes where the effective refractive index is less than 1. For the free space coupling setup the prism used is air, thus only modes with an effective index less than that of air are allowed, as the condition seen in equation 2.21 still holds [jackhammer]. 2.3 NLO origins and effects When an electromagnetic field is incident onto any material, the bound charges that make up that material respond to it. For most materials, these bound charges are charges found in atoms. The Electric part of the EM field is absorbed by these charges and they then are displaced. The relationship between the applied field and the charge displacement induced in the material is the polarisation of the material, . For optically linear materials the relationship between and the applied E-field is linear, related simply by the permittivity of free space and the linear susceptibility of the material, (1). The energy absorbed is then emitted radiatively or non-radiatively. If the material emits on radiative modes, and is homogeneous and isotropic, the field then travels through the material in the same direction it entered. However the wave travels
10

slower through the medium than through a vacuum, due to the time spent through absorption and re-emission. The ratio of the speed of light in a vacuum to that of the speed of light in a material is that materials refractive index, . If the charges in the material emit on non-radiative modes, through heat dissipation, then some energy from the travelling EM wave is absorbed in the material. The absorption of the material is then the ratio of light exiting the material to the amount that entered it, divided by the materials thickness. The refractive index and absorption of a material are linear properties, but can be influenced by higher order effects. Non-linear properties of a material come about when the response of the charges within a material respond to an applied field is higher than first order. This occurs for any material if either the incident light field strength or an applied potential is high enough, but for some materials NLO effects can be seen due to the nature of its bound charge, not only due to field strength. The polarization of a bound charge in a material can be expressed as a Taylor expansion to see the whole range of response orders: ( )

where is the permanent dipole moment of the bound charge, is its linear polarisability and and are the second and third order polarisability (also called first and second order hyperpolarisability). Macroscopic polarisation of the material can be written in a similar manner:
( ) ( ) ( )

where is the permanent polarisation of the material, ( ) , ( ) , and ( ) are the linear, second order and third order susceptibilities respectively. These E-fields can oscillate or be constant, so different kinds of second and third order effects exist depending on the frequencies of the fields. The dc Kerr effect, investigated in this project, is the variance of the refractive index of a material proportionate to the square of an applied electric field. In order to measure this effect, any second order effects must be taken into account as they are typically far stronger than third order effects [George]. However, ensuring that samples used are centrosymmetric prohibits second order effects from manifesting as only non-centrosymmetric materials will show even orders of susceptibility[13], whereas all materials show odd orders of susceptibility. The dc Kerr effect makes a material birefringent if it otherwise wouldnt have been such that the change in refractive index between incident light perpendicular and parallel to an applied field is given by: ( )

where is the difference between the refractive indices, is the Kerr coefficient for a material and is the magnitude of an applied dc electric field.
11

The change in refractive index in the direction of the film layer of a waveguide when subjected to a dc electric field is given by [marvellous] (for full derivation see appendix): ( )

where for TM modes and for TE modes. The quadratic electro-optic coefficient relates, under isotropic conditions, to the third order susceptibility ( ) :
( )

( The dc Kerr coefficient for a material can be gained from ( ) and from: (

The Kerr effect can be observed using the ATR spectra obtained from this project, as they are sensitive to small changes in refractive index of the film layer [14]. Once a mode dip in an ATR spectrum has been found, choosing an angle, , at which the gradient of the dip in reflectance is approximately linear, it can be shown (see appendix for full proof) that a change in refractive index in the film constitutes a measurable change in reflectance [marvellous]: ( ) where is the change in reflectance, the linear gradient the change in the refractive index of the film. 2.4 Organic NLO materials In non organic crystal structures there is a rigorous approach as to the crystal groups that are non-centrosymmetric and which are not, from analysis on the lattice groups and their structures [13]. With organic materials however, there is not such a rigorous approach. Using guesthost organic systems allows one to investigate the effects of altering the Kerr coefficient of a material by doping it with small molecules. The guest-host system of PMMA (poly(methyl methacrylate)) doped with the chromophore MOR2 is chosen for this project for this reason. The other organic material chosen for this experiment is the side-chain polymer IPDI (polydicyanovinyl isophoronediisocyanate). is given by ( and ) is

12

Figure 2.41: The three organic materials used in this project. a) Is the side-chain polymer IPDI (polydicyanovinyl isophoronediisocyanate). b) Is the host polymer PMMA (polymethylmethacrylate). c) Is the guest small molecule MOR2. From figure 2.41 it can be seen that IPDI and MOR2 have a section of their molecular build that is connected by conjugated bonds and section that is not. This change in mobility of the electrons within the bonds is what leads to these materials having high hyperpolarisabilities [reference here would be nice] and is harder to achieve in inorganic material structures [all day mate]. It should also be noted that as the electrons that oscillate subject to incident light are mobile through conjugated bonds, they are a lot more free to move than in the inorganic case and thus allows them a faster response time [all bloody day].

3. Experimental
3.1 Fabrication All of the following procedures for fabrication were conducted in a clean room with full clean room equipment worn: goggles, gloves, full body suit, clean shoes and a clean hat. Due to the very sensitive nature of waveguides to contaminates, two waveguides were made of every different type listed in this section. 3.1.1 Preparation of solutions The two different waveguiding organic systems had to be prepared in solution such that they could be spin-coated to form the film layer of the waveguides. The glass vials that the solutions were made in were first extensively cleaned. This entailed washing in dicholoromethane (DCM) and wiping with dust free dry medical wipes, then washing in isopropanol (IPA), then washing and sonicating in deionised water for 20 minutes, before finally blow drying with dry oxygen free nitrogen. The solutions, once made, were stirred magnetically over night and covered in tin toil to avoid photo degradation [probably].

13

MOR2 PMMA Amount Solution Name of mass mass of solvent conc. Doping conc. soln. (0.0001)g (0.0001)g (0.1)ml % w/w % w/w PMMA1 0.0240 2.0102 10.0 21.5(0.2) 1.194(0.005) PMMA2 0.0134 1.6642 7.0 25.4(0.3) 0.805(0.006) PMMA3a 0.0084 1.9010 7.0 28.9(0.3) 0.442(0.005) PMMA3b 0.0090 2.0401 10.0 21.7(0.2) 0.441(0.005) PMMA4 0.0050 2.2900 7.0 34.7(0.3) 0.218(0.004) PMMA5 0.0000 2.1454 10.0 22.7(0.2) 0 Figure 3.1.11: A table showing all appropriate characteristics and their errors of the solutions made for the guest-host system of MOR2 doped in PMMA. The solvent used was DMF. Name of soln. IPDI1 IPDI2 PDCV-IPDI mass (0.0001)g 0.5820 0.4222 Amount of solvent (0.1)ml 4.0 2.5 Solution conc. % w/w 29.20.7 17.90.7

Figure 3.1.12: A table showing the appropriate characteristics and their errors of the solutions made for the side-chain polymer PDCV-IPDI. The solvents used were cyclohexanone for IPDI1 and DMF for IPDI2. To investigate the effect of varying different parameters of the film layer of the waveguides, several solutions were made comprising of the two organic systems used in this project. The solutions relating to the guest-host system of MOR2 doped in PMMA can be seen in figure 3.1.11, while the solutions relating to the side-chain polymer PDCV-IPDI can be seen in figure 3.1.12. 3.1.2 Waveguide structures and cleaning ISO 8037 glass slides were cut into thirds using a glass cutter and a straight edge to form the base for which the waveguides were constructed. The slides were cleaned in the same manner that the glass vials were cleaned in section 3.1.1. All washing and drying was conducted by holding the edge of the slides with fingers with nitrile gloves on. The surfaces of the slides were never touched except when wiping. Two different structures were fabricated for these waveguides. The free space coupled method [10], and a base coupled structure that is similar to prism coupled structures seen in the literature [9]. The specific layouts of these structures can be seen in figure 3.1.21. In order for the metal cladding layers to act as electrodes they had to be deposited in a pattern such that they overlapped in the centre of the slide but nowhere else, as shown in figure 3.1.22. For a contact to be made with the first deposited silver layer, the film layer had to be removed. This was done by carefully soaking the edge of the structure in chloroform for approximately 10 seconds, then gently wiping away the film with a lint free dry medical wipe.

14

Figure 3.1.21: Free space coupled and base coupled structures are shown, with the arrows being the incident light to the waveguides. The nature of the various layers is discussed in full in this section. Right at the end, put in some labels for the cover layers.

Figure 3.1.22: The specific structure of the waveguides in order for the substrate and cover layers to act as electrodes. The film layer on the left edge would have to be removed for a conducting contact to be made. 3.1.3 Deposition of silver layers The initial layers of silver were deposited on the centre of the glass slides using a BOC Edwards evaporator. This evaporator achieved a pressure around 10-5 Torr (1 Torr = 1/760 Atm), by utilising first a mechanical backing pump to achieve around 10-1 Torr, then an oil diffusion pump to raise the vacuum higher. Silver was thermally evaporated by use of a current of 25(1)A being passed through a molybdenum boat containing a silver coil. For 5 seconds the covering baffle was closed to allow imperfections to be cooked off. It was then opened to allow deposition of silver onto the glass slide. Deposition thicknesses used for fabrication were 11.2, 18.0 and 24.0 (0.1)nm for the cover layer, whilst thicknesses mostly of more than 200nm were used for the substrate layer. A quartz crystal vibration detector was utilised to measure the thickness of the deposited layer. Controlling the thickness accurately was challenging due to there being a delay between the baffle being closed and the thickness meter showing an invariant value, thus several thicknesses evaporated differed slightly to these cover layer values. The deposition rate of the silver is proportional to the magnitude of the current flowing through the molybdenum boat[reference mother fuckeerrrrr]. At a current of 25(1)A
15

the deposition rate was observed to be approximately 0.1nms-1. This was the slowest deposition speed that could be obtained, as below this rate the amount of current flowing through the boat fluctuated too wildly. The slowest and most steady deposition rate are preferential because this ensures a common density of silver perpendicular and parallel to deposition does it?. A slight flaw of vapour deposition is that the deposition will always be at its highest rate at the shortest distance between the boat and the sample[reference that shit again, yo]. This is anticipated in the design of the evaporator, with the distance between the sample and the boat large compared to the size of either. 3.1.4 Formation of thin films The thin polymer films formed in this project were spin coated. To ensure as little contamination as possible, the slides were immediately placed from the evaporator to the spin coater after the first silver layer was deposited. The spin coater had an environment of nitrogen gas constantly pumped through it to further prevent contamination to the waveguide. The slide was held in place by vacuum. Millipore 0.50m hydrophilic PTFE membrane with 1.0m APFB glass fibre prefilter filters were used when depositing around 0.5ml of the solutions through syringes onto the thin silver cover layers for spin coating. The drops were deposited directly onto the middle of the slide, with the pre-programmed spin session started immediately after the drops fell to ensure even consistency of solution after spinning. The spin coater used was the Laurel WS-400 Lite series spin processor. Final spin speed and solution concentration have been shown to be the most important factors for thin film thickness [15], so the exact amount of solution deposited each time did not need to be carefully measured. 0.5ml seemed approximately enough to give a full coverage whilst not being too much to waste the solution. It was found that 1300rpm for 15.0s seemed appropriate for allowing the variance of film thickness by solution strength, apart for when attempting to fabricate a single mode waveguide, where 1600rpm was used. Post spin coating the slides were immediately placed in an oven at 80(10)C which was then immediately placed under vacuum for at least two hours. This ensured the solvent was evaporated, leaving only the organic material. The last few steps were done as quickly as possible such that there was as little contamination as possible amongst the waveguide layers. All moving of the glass slides at their various stages of housing the waveguides were done delicately but quickly with tweezers at the very edge, making sure not to touch the central part where the overlapping of the silver cladding was located. After completing fabrication, the waveguides were stored in small slide containers and then wrapped in aluminium foil to avoid photo degradation. 3.2 ATR technique As explained in the theory section, modes of the waveguide will be excited only if the light incident is at the correct angle. The method to find these angles was the ATR method.
16

Figure 3.21: The experimental set up used to obtain angular reflectance spectra. From figure 3.21, the arrangement of rotating the sample degrees and rotating the photodiode degrees can be seen. The laser used was a polarized HeNe 632.8nm 5mw laser. The polariser used that was calibrated such that at 0 the E-field of light passing through it is along . Thus rotating the laser until a maximum intensity was found when the polariser was set to gives TM mode propagation in the waveguide. Setting the polariser to and finding the maximum intensity ensures that the H-field is along and therefore gives TE mode propagation in the waveguide. The laser was initially set up so that the beam was directly normal to the polarizer and waveguide. This was accomplished by observing the reflected beam returning to the laser cavity exactly. A set square was used to measure the height at which the laser beam was with regard to the optical bench at various points along the lasers path to ensure that there was no pitch angle between the laser and the sample, and that the beam would always be, as much as possible, incident on the same part of the waveguide. The mount on which the sample was placed allowed for initial easy manoeuvrability and screws to tighten such that the sample could be firmly held in place at any pitch or perpendicular displacement with respect to the beam. This was necessary as probing the waveguides for the best areas was common, where often some parts of the waveguides would show light coupling while others would not, due to film quality variations. The beam was positioned so that it passed exactly though the axis of rotation of the stage, and the photodiode was rotated around the same axis of rotation as the sample by being attached to the goniometer, which for simplicitys sake is not shown in figure 3.21. The angular range measured for was from to . To investigate deviation in measured current from the photodiode across the angular range used for the ATR spectra, a waveguide was placed in the mount backward, acting as a
17

mirror. Once proper alignment had been achieved, it was found that the very slight variances in measured photodiode current were only due to blemishes in the substrate layer of the waveguide. Noting the reading on the multimeter attached to the photodiode every half a degree allowed angular spectra of current recorded by the photodiode to be constructed. To obtain the reflectance spectra from the current spectra, the current readings were simply divided by the highest current reading obtained, as this then amounted to unity reflection. The laser had to be left on for several hours to reach and stabilize a maximum power. The light was on in the lab during all sessions to ensure that a persons pupils were small in case of accidental direct viewing on the laser, but as the class of laser was 3B the damage caused by direct viewing is extremely low [reference this surely]. 3.3 Film thickness measurements A Taylor-Hodson tally step machine was used to measure the thicknesses of the films of the waveguides. This was carried out by carefully scratching off a small section of the polymer film at a point on the waveguide, then subjecting that small section to the tally step to measure the thickness of the film. The spin coating technique for polymer solutions is known to give approximately uniform film height [reference it] but the edges of the slides were appreciably different in consistency and thickness due to build up there, so the edges were avoided for tally step measurements. 3.4 Absorption spectra To ensure no absorption processes involved with the organic materials interfered with the angular reflectance spectra, two samples were prepared as above but with no silver layers. The solutions used for these samples were PMMA3a and IPDI2. Absorption spectra were found for both samples using a Shimadzu UV/VIS spectrometer. 3.5 dc Kerr effect measurements As the Kerr effect is a very weak and quadratically dependant on an applied dc voltage, a power generator producing voltages up to was used. As such high voltages were used, the mount had to be completely shrouded. The mount on top of the rotating stage was enveloped in a shrouded environment consisting of insulating plastic. The front was transparent and had a horizontal slit cut out as to not interfere with the incident and reflected laser beam. The plastic shroud had connections to crocodile clips inside the environment which were then attached to the electrodes of the waveguides. The transparent front part was also removable to allow positioning of the sample on the mount, but safety switches ensured a break in the circuit if the lid was off. From equation 2.35, to maximise for a specific , needs to be maximised, and ( ) needs to be minimised. Minimising ( ) is achieved by using the lowest
18

order mode available, while is maximised by finding the steepest mode dip in the ATR spectra. Thus specific modes from each ATR spectra will give the most sensitive measurement of the change in refractive index of a film layer, and these were found for each waveguide. As commercial grade slabs of PMMA are known to have a low Kerr coefficient of ~ at 632.8nm [tout], too low to be seen with the setup in this report, analysis can be carried out by varying the doping of MOR2 in PMMA, to see if the Kerr coefficient can be bought up to detectable levels of ~ . The Kerr coefficient of isotropic IPDI at 632.8nm can also be found if it is larger than ~ , as it is currently unknown. and can also be analysed for these systems to an accuracy of ~ , and given that other polymers such as polyimide have and of ~ [marvellous], it does not seem unreasonable that analysis can be carried out.

4. Results and Discussion


4.1 Absorbance spectra To ensure no absorption effects altered the ATR spectra such that reflectivity could be altered without light being coupled to the waveguide, the absorbance spectra of solutions PMMA3a and IPDI2 were obtained. Absorbance 2

PMMA3a
1.6 1.2 0.8 0.4 0 300 350 400 450 500 550 Wavelength (nm) 600 650 700 IPDI2

Figure 4.11: The absorbance spectra of PMMA3a and IPDI2. There is no appreciable absorption at 632.8nm. The HeNe laser used operated at 632.8nm, and from figure 4.11 it can be seen there is negligible absorption at this waveglength for the host, dopant and side-chain materials. Thus absorption effects can be ignored with regard to the ATR spectra. 4.2 Unsuccessful waveguides Many waveguides initially fabricated did not show ATR spectra with reflectivity dips, indicating that no light was being coupled to the guiding layer.
19

The first few waveguides were constructed with a cover layer of around 50nm. This is clearly thicker than the skin depth of silver at 632.8nm for this environment, as no coupling occurred at all. It seemed prudent to investigate the effect varying the cover layer thickness had on the ATR spectra following from this, as the literature typically only uses thicknesses of ~24nm [ref]. More often the failure of waveguides was down to the quality of the film layer. In some cases it could be seen that the film layer was inhomogeneous in thickness to an unacceptable extent, indeed for IPDI despite many waveguides being made when cyclohexanone was used as a solvent for its solution, none showed light coupling. Once DMF was used as the solvent for it however, a smoother film could be seen as the comparison in figure 4.21 shows. Subsequently light coupling was observed for an IPDI film layer.

Figure 4.21: Photos of a waveguide using IPDI2 (left) and one made with IPDI1 (right). It can be seen that in the waveguiding section of the samples (where the silver layers overlap), that the IPDI2 film is a lot smoother than the IPDI1 film. 4.3 Successful waveguides Although many fabricated waveguides were unsuccessful in producing ATR spectra that correspond to waveguiding, enough were successful such that suitable analysis can be made. The successful waveguides fabricated can be seen in figure 4.31. Waveguide name WG1 WG2 WG3 WG4 WG5 WG6 WG7 WG8 WG9 WG10 Solution Structure type Cover layer thickness (0.1nm) 11.2 18.0 24.0 11.2 24.1 23.9 24.1 23 23.8 24.4 Substrate thickness (0.1nm) 208.5 208.5 211.2 320.0 217.9 209.4 204.4 239.4 218.1 236.6

PMMA3a PMMA3a PMMA3a PMMA3b PMMA3b PMMA1 PMMA1 PMMA2 PMMA4 PMMA5

Base coupled Base coupled Base coupled Free space coupled Free space coupled Free space coupled Free space coupled Free space coupled Free space coupled Free space coupled
20

WG12 IPDI2 Free space coupled 11.4 211.1 Figure 4.31: A table showing the successful fabricated waveguides. For film thickness approximations see section 4.31. The substrate thickness for each waveguide fabricated varies considerably, however as seen in section 4.2 that silver seemed to act as a mirror even when only 50nm thick , so to ensure no substrate effects interfered with the investigation, only a minimum thickness of 200nm seemed necessary for the silver substrate layer. 4.4 Film thickness measurements and code generated ATRs Only one film thickness measurement was made for each waveguide measured, but as from section 3.3 it is seen that as long as edge is not used to measure the thickness, the thickness measured and the thickness of the film at the waveguiding section should be approximately equal. As the TE modes and TM modes ATR spectra of the waveguides were not collected at the exact same point on each waveguide, it can be seen whether the films are approximately uniform in height. In all experimentally achieved waveguide ATR spectra, there was never a difference in the number of modes seen for the TE and TM spectra, except for WG10. The TM modes were mostly seen at slightly higher angles than TE modes, which is in line with theory [reference]. Thus although there are slight variations in film thickness across the waveguide, these variations are not enough to distort the shape of a dip in an ATR spectrum enough such that the equations seen in section 2.3 would not hold. Analysing the spectrum of WG10 allows more specific analysis, as the permittivity of PMMA is known to be 2.217 at 632.8nm [4.42], and Taylor-Hodson tally step machine was used to measure the film thickness at a point on the film layer, giving a thickness value of 3.250.01m. The TM and TE mode dips from the experimental ATR spectra are very close together, implying a good uniformity of the film layer for WG10. However, for high angles of incidence the laser beam was spread over a much larger area of the waveguide, and therefore is subject to more changes in thickness. Thus the dip after 70 seen in the theorised plot is barely seen in the experimental plot. The theorised dips in figure 4.41 are not exactly in the same place as they are in the experimental dips, but they are very close. Comparing the angular difference for each mode from theory and experiment allows a corresponding difference in film thickness to be found from equation 2.1.35, as tabulated in figure 4.42. The average difference in film height from figure 4.42 is calculated as 0.64 m with a standard deviation of 0.62 m. This value and its associated error seem rather high given that spin coated polymer layers are categorised to be approximately uniform in height seen in some section or paper. This can be accounted for by recognising that the experimental ATR

21

Reflectance 1 0.8 0.6 0.4 0.2 0 0 10 20 30 40 50 Angle of incidence () 60 70 Theory TE Theory TM Expt. TM Expt. TE

Figure 4.41: A plot of the experimental and theoretical ATR spectra for WG10. Film thickness of 3.25m was used for the theoretical plots. The experimental TE and TM ATR spectra were obtained from different points on the waveguide. The error bars are shown, but are all smaller that the shape that contains them. Corresponding difference in film height (m) TE3 6.40 0.54 TM3 5.34 0.16 TE2 1.30 1.57 TM2 2.03 0.18 TE1 1.34 1.22 TM1 0.89 0.15 Figure 4.42: A table showing the difference in coupling mode angle for the theoretical ATR spectra for WG10 using the film layer thickness as 3.25m. The equation 2.1.35 was used to generate the corresponding film thicknesses for each angular mode separation. The errors associated with the mode angle difference and therefore film height difference are very small compared differences amongst the film height differences themselves, so they have not been displayed. spectra are distorted by the film height changing across the spectra due to the spread of the incident laser light increasing for higher angles as previously mentioned, which cannot be modelled by the code. This difficulty is overcome by using a laser with as small a beam width as possible. Thus while it is definitely fair to say that there is an observed non uniformity in film height across the waveguide, the value of 0.64m is an overestimate. 4.5 Variance of structure type The glass base had to be extremely clean for the based coupled waveguides otherwise the incident laser light would be unacceptably dispersed. It was thought that the base coupled structure would have an advantage over free space coupled structure in that the thin silver layer would not be exposed, and so would be less susceptible to degradation, however over
22

Mode number

Absolute difference in mode angle ()

Solution strength % w/w 21.5 0.2 21.7 0.2 22.7 0.2 25.4 0.3 28.9 0.3

Normalised solution strength (0.01) 0.62 0.63 0.65 0.73 0.83

Number of modes between 0 and 40 TE TM 1 1 1 1 2 2 2 2 8 8

Structure type

free space free space free space free space base coupled

34.7 0.3 1.00 5 5 free space Figure 4.51: A table showing the difference in the number of modes seen in waveguides for increasing strength of film solution with reference to their structure type. For simplicity of analysis the solution strengths were normalised with respect to the highest value. the time frame of this project of approximately 20 weeks, the free space coupled structures showed no signs of degradation when exposed to air and being carefully moved around. Total internal reflection within the glass base layer reduced the incident laser power to the waveguide by around a half, so this led to the error on reflectance being doubled. Doubling the error on reflectance halves the sensitivity of waveguide to dc Kerr measurements and thus the base coupled structures are disadvantaged to the free space coupled structures from this standpoint. Due to the refraction of the laser light through the glass base layer, the base coupled ATR spectra angle of incidence to the waveguide only reaches a maximum of around 40, whereas the free space coupled ATR spectra reach the full 73 obtainable with this setup. As the same amount of angular steps was taken for each, the base coupled structure allows better angular resolution for the lower order modes. From equation 2.37 it is seen that higher order modes are more sensitive to changes in refractive index of the film layer, so these are of more use for Kerr shift measurements. The largest difference between the structure types is the amount of allowed modes that can be supported. As the refractive index of glass is higher than air, there should be a larger mode density for the base coupled structure than from the free space structure, as there is a larger allowed mode range, which can be seen from equation 2.21. Given that all waveguide film layers were spin coated at 1300rpm, the strength of the solution used for a waveguide is approximately proportional to the film thickness for that waveguide [yep]. In figure 4.51 it is seen that there are more modes over the same angular range for the base coupled structure than for any of the other structures, regardless of solution strength. Even despite small discrepancies in the relation between solution strength and film thickness, it is clear that the base coupled structure has a higher mode density than the free space structure, as predicted from the restriction shown in equation 2.21. From a theoretical standpoint the base coupled structure has some qualities convenient for the observation of dc Kerr shifts, as they have a higher mode density which
23

leads to steeper mode dips and they also have better angular resolution. However, minimising ( ) is important as seen from 3.5, so the limit of 40 of the ATR spectra observable for the base coupled structure renders them not as useful as the free space coupled structure. In practise it was also more difficult to achieve deeper, more uniform mode dips from the base coupled structures. Many base coupled structured waveguides fabricated showed coupling of less than 10%, and usually only a few modes were seen rather than the amount expected. These problems were due to the glass base layer needing to be extremely clean and uniform on both surfaces, which although was sometimes achieved, no difference in cleaning technique seemed to have a direct influence on the quality of the waveguides. Free space coupled waveguides showed deeper, if not slightly less steep, angular mode dips, consistently achieving coupling of 40% or more. Indeed 83.40.1% coupling for the 11.2(0.1)nm cover layer seen in figure 4.61 is surprisingly high given how easily coupling is affected by contaminants leading to deviations in a smooth waveguide surface. 4.6 Variance of cover layer thickness Silver cover layers of 50(0.1)nm were determined to be too thick to allow the light through, as several waveguides were fabricated with both structures but no mode dips were seen in their ATR spectra. Varying the silver cover layer from 11.2 to 24.1(0.1)nm shows a large difference in the shape of the mode dips seen in the ATR spectra, as can be seen in figure 4.61. Although the TE mode dips have not been shown for clarity, their shapes were almost identical to those seen in figure 4.61. The width of the mode dips for WG4 are clearly much larger than those of WG5, even to the extent that there is very close to a linear plot from the bottom of the dip to the beginning of the next. The different shapes of the dips can be explained by the nature of how the light couples to the film layer from free space. When the metal cover layer is thick such that there is only a small field tunnelled through to the film layer, propagation through the film layer becomes very lossy even for small deviations away from a mode angle. If the cover layer is thinner and there is a larger field in the film layer, then more lossy modes can be sustained for longer, such that the mode dip shape is wider. There is a critical value for cover thickness where the coupling is largest, where after which the dip will still continue to become wider but will not get any deeper [ref]. This critical cover thickness value varies depending on the thickness and refractive index for each particular waveguide, and is clearly closer to 11.2nm than 24.1nm from figure 4.61. Although this is not shown in figure 4.61 due to slight variations in film thickness occurring even when the same solutions are used at the same spin speed, changing the cover layer thickness should not alter the position of the mode angles [ref]. Varying the cover layer thickness for the base structured waveguides seemed to have less of an obvious effect to that of the free space coupled structures. Figure 4.62 shows that although the WG1 ATR spectra, with the thinnest cover layer of 11.20.1nm, had the deeper
24

Reflectance 1 0.8 0.6 0.4 0.2 0 0 10 20 30 40 50 Incident angle () 60 70 80 WG4 TM WG5 TM

Figure 4.61: The ATR spectra of WG4 and WG5 for TM modes only. Reflectance 1 0.8 0.6 0.4 0.2 0 10 12 14 16 18 Incident angle () 20 22 24 WG1 TE WG2 TE WG3 TE

Figure 4.62: The TE ATR spectra of WGs 1, 2 and 3 between 11 and 24. Although the mode dips were not perfectly aligned in the experimental data, they have been shifted here for ease of mode dip shape analysis. These implemented shifts were no more than by 2. mode dips as expected, there was little variance between the shapes of the dips for WG2 and WG3. WG1 is also seen to have the sharper dips of the three, which is unexpected. A reasonable explanation for these differences from expectation is that of non uniformity of the glass base layer, cover layer and/or film layer. 4.7 Single mode waveguides Several waveguides were fabricated at 1600rpm with the most dilute solutions shown in figures 3.1.11 and 3.1.12 to try and achieve a film layer thin enough that would support
25

only a single mode. These very thin film layer waveguides did not show any mode dips in their ATR spectra except for WG12. Reflectance 1 0.8 0.6 0.4 WG12 TE 0.2 0 WG12 TM

10

20

30 40 Incident angle ()

50

60

70

Figure 4.71: The single TE mode and TM mode ATR spectra for WG12. Although it is possible that there could be other modes in figure 4.71 that are not being seen by the experiment due to contamination, this seems very unlikely due to how smooth the spectra are until the dip is seen for both the TE and TM case. Also the dips seen are wide as expected due to the cover layer for WG12 being only 11.4(0.1)nm. Thus it is fair to say that WG12 is a single mode waveguide, and that the material IPDI can be made into a successful waveguide, just requiring DMF as the solvent and not cyclohexanone. 4.8 dc Kerr effect measurements Unfortunately only one fabricated waveguide could be investigated for the observation of the dc Kerr effect. This was due to their being a current flowing from the cover layer to the substrate layer of all other waveguides, such that no voltage could be applied across the film layer of the waveguide. This seems to have been caused by very slight deformations in the film layer of the waveguides, not noticeable when looked at by eye. As this problem was unanticipated and due to the time constraints on this project, it was not possible to fabricate more waveguides with thicker films in order to overcome this setback. Thicker film layers were not initially investigated due to them having many more modes, thus making their ATR spectra a lot harder to investigate or even to see at all, due to the angular resolution of the goniometer being 0.25. WG8 however did allow a voltage of up to 200.0(0.1)V across it. The most suitable modes were located and their gradients were found, as can be seen in figure 4.81.

26

Reflectance 1 0.8 0.6 0.4 0.2 0 55 57 59 61 63 Incident angle () 65 67 69 WG8 TM

WG8 TE

Figure 4.81: The TE and TM mode dips with the lowest values for ( ) in the ATR spectra for WG8. The places where the absolute value for are highest are where are taken, and they are marked with vertical lines. These values are and . No change in intensity was measured at the angles seen in figure 4.81 from 0.00.1V all the way up to 200.00.1V for TE or TM modes. This allows an upper bound of the dc Kerr coefficient for this material and its associated error to be calculated. Taking in equation 2.35 to be the error on the photodiode leads to the upper bound on to be ( ) ) and to be ( . From equation 2.36, these two bounds give a range for for this film as( ) . It seems fair to say that this level of MOR2 doping has not substantially changed the Kerr constant for isotropic PMMA, even despite there being possible density differences between the film layer in WG8 and the commercial grade PMMA it is being compared to in section 3.5. Calculating these bounds carried the assumption that the refractive index of the film layer was the same as that of PMMA, just that it was given a 2% error. Given the doping was only 0.8050.006% w/w it seems reasonable that the refractive index of the film layer wouldnt change by more than 2% of its value undoped [the last one prob]. The film height, , was that obtained from the Tally-step method, and was given an error of in light of section 4.4. The errors on and are very high, but would be reduced if better angular resolution could be achieved. 5. Conclusion The properties of metal clad symmetric waveguides were examined in terms of the variance of structure, cover layer thickness, film thickness and film thickness uniformity. Varying these parameters allowed optimisation of the devices for several different uses. The
27

uses achieved/analysed in this report were; the successful fabrication and demonstration of a single mode waveguide, achievement of a maximum coupling efficiency of 83.40.1%, insight into the film thickness nonuniformity associated with spin coating, and analysis of the dc Kerr coefficients for the organic materials used. Film layer quality often hampered waveguide analysis; coupling efficiency is greatly reduced by contaminates and voltages could not be sustained across the electrodes of all fabricated waveguides except one. Upper bounds for to and and therefore a range for the dc Kerr coefficient were found for 0.8050.006% w/w doped MOR2 in isotropic PMMA. These bounds were ( ) ) and ( for and respectively, leading to a range for the dc Kerr coefficient of ( ) . This range implies that doping MOR2 into isotropic PMMA at 0.8050.006% w/w does not significantly change the dc Kerr coefficient compared to undoped isotropic PMMA. Greater angular resolution would have allowed thicker films to be analysed with regard to the dc Kerr effect, increasing the probability that film quality would be sufficient to sustain voltages across the metal clad electrodes, whilst also allowing higher sensitivity of the waveguides to changes in refractive index, allowing more refined electro optic measurements of the film layers.

References
[1] Kapron et al., Radiation losses in glass optical waveguides, Applied Physics Letters 17, 423 (1970) [2] K. Satzke et al., Ultrahigh-bandwidth (42 GHz) polarisation-independent ridge waveguide electroabsorption modulator based on tensile strained InGaAsP MQW, IEEE Electronics Letters 31 (23), 2030 (1995) [3] J. J. Degnan, Optimization of passively Q-switched lasers, IEEE Journal Quantum Electronics 31 (11), 1890 (1995) [4] L. E. Hargrove, R. L. Fork, and M. A. Pollack, Locking of HeNe laser modes induced by synchronous intracavity modulation, Applied Physics Letters 5, 4 (1964) [5] N. Skivesen, Metal-clad waveguide sensors, PhD Thesis, Riso National Laboratory (2005) [6] D. Marcuse and E. A. J. Marcatili, Excitation of Waveguides for Integrated Optics with Laser Beams, Bell System Technical Journal 50, 43 (1971) [7] M. L. Dakss et al., Grating coupler for efficient excitation of optical guided waves in thin films, Applied Physics Letters 16, 523 (1970) [8] P. K. Tien and R. J. Martin, Experiments on light waves in a thin tapered film and a new light wave coupler, Applied Physics Letters 18, 398 (1971) [9] D. Sarid, P. J. Cressman and R. L. Holman, High efficiency prism coupler for optical waveguides, Applied Physics Letters 33, 514 (1978)

28

[10] H. Li et al., Free-space coupling of a light beam into a symmetrical metal-cladding optical waveguide, Applied Physics Letters 83, 2757 (2003) [adam] S. B. Mendes et al., Broad-band attenuated total reflection spectroscopy of a hydrated protein film on a single mode planar waveguide, Langmuir 12, 3374 (1996) [Balthazar] L. Lvesque et al., Precise thickness and refractive index determination of polyimide films using attenuated total reflection, Applied Optics 33, 8036 (1994) [chrome] Y. Jiang et al., Low voltage electro-optic polymer light modulator using attenuated total internal reflection, Optics & Laser Technology 33, 417 (2001) [dandelion] G. F. Lipscomb et al., Poled electro-optic waveguide formation in thin-film organic media, Applied Physics Letters 52, 1031 (1988) [excelsior] H. Y. Lin et al., Improvement of the outcoupling efficiency of an organic lightemitting device by attaching microstructured films, Optics Communications 275, 464 (2007) [flutter] S. Rhle et al., Optical waveguide enhanced photovoltaics, Optics Express 16, 21801 (2008) [tyrant] P. N. Prasad and D. J. Williams, Introduction to nonlinear optical effects in molecules and polymers, Wiley, New York (1991) [11] H. S. Nalwa, Organic Materials for Third-Order Nonlinear Optics, Advanced Materials 5, 341 (1993) [12] P. K. Tien, Intergrated optics and new wave phenomena in optical waveguides, Rev. Mod. Phys 49, 361 (1977) [jackhammer] H. Lu et al., Study of ultrahigh-order modes in a symmetrical metal-cladding optical waveguide, Applied Physics Letters 85, 4579 (2004) [micheal] P. B. Johnson and R. W. Christy, Optical constants of the noble metals, Physical Review B 6, 4370 (1972) [13] M. Melnichuk and L. T. Wood, Direct Kerr electro-optic effect in non-centrosymmetric materials, Physical Review A 82, 13821 (2010) [Probably] Y. Ren, M. Szablewski and G. H. Cross, Waveguide photodegradation of nonlinear optical organic chromophores in polymeric films, Applied Optics 39, 2499 (2000) [George] J. M. Hales and J. W. Perry, Organic and polymeric 3rd order nonlinear optical materials and device applications, in S.-S. Sun and L. Dalton (eds.), Introduction to Organic Electronic and Optoelectronic Materials and Devices, CRC Press, Orlando, FL (2008). [marvellous] J. Zhou et al., Determination of dc Kerr coefficients of polymer films with prism-optical waveguide configuration, Applied Physics Letters 88, 21106 (2006)

29

[14] Y. Jiang et al., Improved attenuated-total-reection technique for measuring the electrooptic coefcients of nonlinear optical polymers, J. Opt. Soc. Am. 17, 805 (2000) [15] C. J. Lawrence, The mechanics of spin coating of polymer films, Phys. Fluids 31, 2786 (1988) [4.41] J. P. Cresswell, Waveguiding in electrooptic Langmuir-Blodgett films, PhD thesis, University of Durham (1992) [4.42] S. N. Kasarova et al., Analysis of the dispersion of optical plastic materials, Optical Materials 29, 1481 (2007) [tout] K. S. Kim et al., Kerr effect in solid polymethylmethacrylate and polyethene, Journal of Applied Physics 54, 449 (1983) [last one prob] N. A. Hackman, Nonlinear optical characterisation of organic chromophores and aspects of molecular aggregation, PhD thesis, University of Durham (2001)

30

Appendix

How you worked out errors in section 4.4 lol I mean Error propagation on theta: The angular error used for the free spaced coupled method is 0.25. This is due to 2 having an accuracy of 0.5, and so therefore the error on has half this value. This follows from error propagation analysis such that if

then,

where

is the error on

and

is the error on

and

is a constant.

The angular error for the base coupled method requires more calculation. Given that the incident angle on the glass, , is related to the incident angle of the light on the waveguide by ( where on , is the refractive index of the glass and , will be given as ( ) )

is assumed to be unity , then the error

where ( )

and

is the error on .

The error on and are propagated from the above error analysis rule seen in the above section, but also the analysis such that if is a function of and that ( ) then |

31

Errors for upper limits on s11 and s12 The equation used to calculate the upper limit on s11 and s12 is as follows: ( )

where all the variables have been previously defined, except is the chosen angle to use for either the TM or TE mode, and is used as the error on the photodiode. All the variables in this equation have associated errors except for . The methods used to propagate these errors are the same as those used as those for propagating the errors for theta in the above section.

32

Das könnte Ihnen auch gefallen