Sie sind auf Seite 1von 18

Invited Paper

Silicon Photonics The Early Years


Graham T. Reed*, William R Headley, C E Jason Png Advanced Technology Institute, University of Surrey, Guildford, Surrey, UK GU2 7XH
ABSTRACT
The purpose of this paper is to set the scene for what promises to be an outstanding conference. To this end the paper will survey the early work in silicon photonics from the late 1980s to the mid 1990s. This was when the more fundamental studies of basic building blocks were carried out, such as study of the silicon optical waveguide itself, the contributions to loss and improvement of waveguiding devices. Issues such as how to achieve modulation, and how to implement a modulator, the criteria for single mode propagation will also be covered, as well as work on the beginnings of optical circuits in silicon and SOI. The focus will be upon pure silicon, usually, but not exclusively in the form of Silicon on Insulator (SOI), as opposed to work on compounds such as SiGe or SiC. Much of this work still resonates with work being carried out today, because the move to smaller and more efficient devices means that some of these issues must be revisited in order to achieve optimal device performance. Hence the paper will provide a summary of the early work on silicon photonics, and attempt to relate it to some of the issues being studied today.

1. INTRODUCTION
Ever since the earliest research on optical circuits, dating back to the 1970s, there have been visions of an optical superchip (see for example [1]), containing a variety of integrated optical components to carry out light generation, modulation, manipulation, detection and amplification (Figure 1). The early work was associated with ferroelectric materials such as Lithium Niobate (LiNbO3), and III-V semiconductors such as the Gallium Arsenide (GaAs) and Indium Phosphide (InP) based systems. LiNbO3 was interesting almost solely due to the fact that it possesses a large electrooptic coefficient enabling optical modulation via the Pockels effect. Alternatively the III-V compounds were interesting because of the relative ease of laser fabrication, and the prospect of optical amplification and electronic integration. However, the dominance of silicon as the semiconductor of choice for electronics applications eventually led to the investigation of silicon photonic circuits, primarily because of the potential attraction of integration with electronics in a cost effective manner. Such research began in the mid 1980s and has continued ever since. The global research effort was modest at first, but has increased rapidly in recent years. What deterred more significant progress during the early years was that firstly the indirect bandgap of silicon means that light emission from silicon is very inefficient, and nontraditional techniques must be investigated if a silicon laser is to be realised. Secondly, the centro-symmetric crystal structure of silicon means that it does not exhibit a linear electro-optic (Pockels) effect. Since this is the traditional means of implementing an optical modulator in an optical waveguide based device, modulation of the refractive index of silicon must be carried out in another way. This paper discusses the early research on silicon photonics from the mid 1980s to the mid 1990s. Whilst a considerable amount of work had been carried in related materials (for example Si3N4 [2-5], SiON [6-14], SiO2 [15-19], SiGe [20-24], and SiC [25-27]) this paper focuses on single-crystal silicon devices and the ultimate goal of monolithic silicon integration. Whilst these other materials are compatible with standard CMOS technology, they potentially add complexity to the processing. Therefore the goal of this paper is to summarise the highlights not only of the advancements made towards realising an integrated, monolithic silicon optical circuit, but of the individual photonic components that comprise it. A brief history of optical integration in silicon is given first. This is followed by a discussion of several of the important components that would comprise an integrated device or circuit, these being waveguides, modulators, switches, and detectors. However, since research on light sources in silicon is reviewed elsewhere in this conference, silicon light sources are explicitly excluded from this paper.

g.reed@surrey.ac.uk; phone +44 1483 68 9122; fax +44 1483 68 9404

Optoelectronic Integration on Silicon II, edited by Joel A. Kubby, Ghassan E. Jabbour, Proceedings of SPIE Vol. 5730 (SPIE, Bellingham, WA, 2005) 0277-786X/05/$15 doi: 10.1117/12.596921

2. MONOLITHIC INTEGRATED DEVICES


There can be no doubt that silicon is the dominant semiconductor material for electronic applications. It is an inexpensive, well understood material, with a high quality native oxide, and excellent thermal and electrical properties. It also provides strong optical confinement for the telecommunication wavelengths when incorporated into waveguiding structures such as Silicon-On-Insulator (SOI). Seemingly, this would make a fully-integrated silicon optoelectronic telecommunications chip an obvious choice. However, because of its indirect bandgap, silicon is a poor light-emitting material. Hence, a fully integrated monolithic light emitting device cannot be fabricated in conventional ways, and therefore researchers have studied a variety of ways to modify the structure of silicon in order to force it to emit light [28]. To compound the issue, silicon is a centro-symmetric crystal such that the Pockels effect is non-existent. Hence, switching and modulation have to be achieved by other means. Nonetheless work was pursued on defining a fully integrated monolithic optoelectronic superchip as originally proposed by Abstreiter [29] for silicon hybrid integration and modified by Soref to present a hybrid device based upon a silicon bench, utilising silicon optical waveguides [30] (see figure 1). More specifically work was pursued upon devices and simple photonic circuits that comprised some of the building blocks of the superchip, as well as steps towards future integration. It is interesting however, that in the light of the progress that has been made in silicon photonics, that such a superchip can now much more realistically be considered as a target for monolithic integration, notwithstanding many of the very significant challenges that remain. The basic premise of a monolithic superchip is that it incorporates all of the necessary components to detect, route, convert, encode, reroute, amplify or even create optical signals that could subsequently be interrogated by electrical circuitry. In Sorefs vision of such a chip, optical fibres were proposed as a means to butt-couple to waveguides on the chip. The fibres were to be supported by high-precision v-grooves etched into the silicon substrate [30]. It is again interesting to note that passive alignment of fibres to waveguides on a silicon photonic circuit remains a challenge today, partly due to the more recent trend to smaller cross sectional dimensions of waveguides and devices. The optical signals would then either be detected by a SiGe photodiode for use in electronic interrogation, or passed to other devices for optical processing by means of waveguides. Due to losses and/or penalties in other parts of the system, the light could also be amplified, or perhaps re-shaped and re-encoded via an optical modulator before re-transmission. Similarly new electronic signals could be encoded onto optical carrier via such a modulator. With recent work in Raman amplification and lasers [31, 32], and fast modulators [33, 34] we can now look forward to monolithic versions of these devices. Furthermore, the vision also included traditional electronic components to control, drive and time these functions. Optical and electronic integration has taken place only in a very simple way (e.g. [33], [34], [35]), and remains a challenge. However, recent initiatives such as the DARPA EPIC programme [36] will also begin to address these issues.

Figure 1: Silicon-based optoelectronic integrated circuits (OEIC) superchip [30].

Silicon photonics research can be traced back to the mid to late 1980s. Research efforts can be categorised into two general areas: a) Work in silicon photonic devices, or work that sought to overcome the deficiencies of silicon and hence preserve the vision of monolithic integration (such as using the plasma dispersion effect to realise an optical modulator);

Proc. of SPIE Vol. 5730

and b) the investigation of innovative methods to create hybrid devices to move towards integrated hybrid optoelectronic devices. It is interesting to note that the optimal route is still an issue of intense debate, but it is undoubtedly true that any monolithic solution will need to provide significant advantages over the hybridised approach in order to succeed. In other words, monolithic integration for its own sake is insufficient justification. Hybridsation on any level, however, will usually lead to greater complexity, higher loss and in many cases, higher fabrication costs. Also, many of the materials required for hybrid integration (such as III-Vs) arent compatible with standard silicon ULSI processing methods. Therefore additional fabrications steps and/or facilities may be required to integrate the necessary components for a given application. However, depending on photonic circuit design, the same argument could be levelled at a monolithic solution. Nevertheless from the viewpoint of scientific elegance, or more pragmatically cost, monolithic integration is desirable. Therefore, we have focussed in this paper, on devices suitable for monolithic integration, although in passing we note some of the significant work in hybridisation in the these early years [3, 5, 7, 12, 15, 16, 18, 19, 29, 30, 37-50].

3. DEVELOPMENT OF COMPONENTS FOR MONOLITHIC PHOTONIC SILICON CIRCUITS


This section considers the building blocks of the silicon photonic circuit that were studied in the early years of silicon photonic research.

3.1

WAVEGUIDE DEVELOPMENT

One of the most critical components of any integrated optical system is the optical waveguide itself. A significant amount of research on the planar waveguide was undertaken in the late sixties and through the seventies and eighties [5158]. In the mid-eighties Soref et al. demonstrated single-crystal silicon waveguides [59, 60]. Whilst these first devices were fabricated using highly doped silicon substrates, other substrate configurations were employed subsequently, such as Silicon on Sapphire (SOS) [61], and SOI [62, 63]. In the late eighties and early nineties, silicon waveguides were beginning to reap the benefits of the development of Separation by IMplantated OXygen (SIMOX) substrates, as well as Bond and Etch-back SOI (BESOI) [64, 65]. Whilst originally developed for the semiconductor industry to prevent latch-up [66], these materials were an obvious choice as waveguiding substrate materials. Several early assessments of SIMOX and BESOI as waveguiding systems were made [62, 67-70], and some of the initial work yielded very large losses [71], but the loss was rapidly improved to respectable levels [72]. In 1989, Davies et al. [69] measured a loss of 4 dB/cm for optical waveguides fabricated in SIMOX. Multiple layer waveguiding structures using SIMOX technology were also demonstrated [63, 70]. Kurdi et al. predicted a loss for a 0.2m thick planar waveguide with a 0.5m buried oxide thickness to have a loss of less than 1 dB/cm [73]. At the University Of Surrey research was also underway to reduce the initial high loss. Initially, Weiss et al. [71] determined a losses as high as 30dB/cm from a 2 m thick planar waveguide. Further efforts were made by Rickman et al. [72] to reduce the loss by investigating the thickness of the buried oxide layer. Their results showed that a buried oxide layer thickness of greater than 0.4 m was necessary to prevent loss due to substrate coupling for a silicon layer of several microns. Reduction of the optical loss progressed quickly. In 1991, Schmidtchen et al. [74] reported a loss of 0.4 dB/cm for a silicon rib waveguide. By 1994, this loss was reduced to a level indistinguishable from pure silicon, as reported by Rickman et al. [75] for TE polarised light for a wavelength of 1.5 m. This measurement demonstrated that silicon was not only a viable waveguiding material, but that the propagation loss was not going to be a serious issue in the development of the technology. In these early experiments, the majority of the work was conducted on relatively large waveguides, of the order of several microns in cross-sectional dimensions (e.g [76-78]). However, there were also preliminary reports of very small waveguides being measured. Typically the loss was very high as discussed above (e.g. [71]), due in part to insufficiently well confined waveguides, and/or significant surface roughness. It is also interesting to note the there were views expressed at the time (early to mid 1990s) that very small waveguides were unlikely to be useful, due in part to the high loss, but also due to the fact that coupling from optical fibres was very lossy. It is therefore interesting to note that

Proc. of SPIE Vol. 5730

current research is actively targeting miniaturisation to micro- and nanophotonic circuits. The coupling of light to these small waveguides remains an issue today, especially for very small, sub-micron, waveguides [79-84]. In most applications, a single mode optical waveguide is required. In order to achieve the necessary criteria of singlemode behaviour in planar SOI based waveguides, the thickness of the overlayer needs to be on the order of a few hundred nanometres. To avoid the need to fabricate such small devices, the rib waveguide structure was used in most applications. One of the early efforts to determine the geometrical constraints required to enable realisation of large cross-sectional single-mode rib waveguides was made by Petermann et al. [85] in the late seventies. In 1991, Soref et al. [24] refined the work of Petermann by normalising his equations to determine the necessary cross-sectional dimensions for a single-mode rib waveguide. The resulting equation is now widely used:

a r 2 (assuming 2b n12 n 2 1 ) 0.3 + 2 b 1 r

(1)

Where a, b, r, n1, and n2 are defined in figure 2(a). 0 is the free space wavelength of light. A plot of this equation as a function of the rib height factor, b, for the SiO2-Si-SiO2 configuration is shown in figure 2(b). A similar plot for air-clad waveguides was also given in the paper. Experimental verification of this work in SOI was carried out by Rickman et al [75, 76]. It is interesting to note that the geometrical requirements for single mode waveguides have also been revisited in recent years, associated with the reduction in device dimensions, and the need to also consider the polarisation dependence of sub-micron waveguides [86]. With these fundamental issues resolved, the work on silicon photonics moved towards more complex devices and circuits, although refinements of the fundamentals as applied to specific applications or device requirements, have been a feature of the field of research [87-89].

(a)

(b)

Figure 2. (a) Dimensions for a single mode rib waveguide. (b) The a/b ratio as a function of b to determine the single mode condition for the Si02-Si-SiO2 rib configuration [24].

3.2

MODULATORS

Without a strong linear electrooptical effect, silicon was left un-investigated as a switching/modulating material for years. However, in 1987 Soref and Bennett [90] investigated the potential modulation mechanisms available in silicon. They showed that electric field based effects are very small but that an injection of free carriers (the plasma dispersion effect), has potential as a modulation mechanism. They studied results in the scientific literature to evaluate the change in refractive index, n, due to experimentally produced absorption curves for a wide range of electron and hole densities, over a range of wavelengths. In particular they focussed on the communications wavelengths of 1.3m and 1.55m. Interestingly their results were in good agreement with the classical Drude-Lorenz model, but only for electrons. For

Proc. of SPIE Vol. 5730

holes they noted a (N)0.8 dependence. As well as quantifying the changes in refractive index with carrier density, they also quantified the changes in absorption [90]. They produced the following extremely useful empirical expressions, which are now used almost universally to evaluate changes due to injection or depletion of carriers in silicon: At 0=1.55m: n=ne+nh=[8.8x10 Ne+ 8.5x10 (Nh)
-22 -18 0.8

(2) (3)

=e+h = 8.5x10 -18 Ne + 6.0x10 -18 Nh where: ne = change in refractive index resulting from change in free electron carrier concentrations. nh = change in refractive index resulting from change in free hole carrier concentrations. e = change in absorption resulting from change in free electron carrier concentrations. h = change in absorption resulting from change in free hole carrier concentrations.

Equation 2 implies a change in refractive index in excess of 1.6 x 10-3 for a change in electron and hole density of 5 x 1017 cm-3, the latter relatively easy to achieve. Thus a realistic modulation mechanism was confirmed in silicon, and the challenge for designers became one of implementing this modulation mechanism both efficiently and at high speed. However, the plasma dispersion effect is not an ideal modulation mechanism because it is derived from a change in the absorption spectrum of silicon via a Kramers-Kronig coupling, and hence any change in refractive index must also suffer an associated change in absorption. For a similar change in electron and hole density of 5 x 1017 cm-3, equation 3 implies a change in absorption coefficient of = 7.25 cm-1, or 31.5dB/cm. The latter sounds like an enormous loss, but modulators are typically of the order of 500m in length, so the active loss is of the order 1.5dB. However, this result also shows that the plasma dispersion effect has significant potential as an absorption-based modulator, or a Variable Optical Attenuator (VOA). Similar equations were determined for 0=1.3m. Thus silicon now had a method to produce relatively fast switches and modulators thereby increasing the likelihood of fully-integrated monotlithic superchip. The earliest designs of modulators were based upon silicon guiding layers fabricated on doped silicon substrates (to form the lower waveguide boundary), the latter having a reduced refractive index via the plasma dispersion effect. Later Silicon-on-Insulator (SOI) waveguiding structures became more popular due to the possibility of strong optical confinement. The first plasma dispersion modulation device in silicon was proposed by Soref et al. [91]. This p+-n-n+ modulator, shown in figure 3, was based on a single-mode, submicron silicon rib waveguide. It was found by modelling that the interaction length of the modulator required for a -radian phase shift was less than 1 mm. The corresponding loss was less than 1 dB at = 1.3 m for both TE and TM polarisation modes. The authors noted that to a first approximation, the modulator was polarisation independent. It is interesting to note the buried block of SiO2 below the waveguide, acting as the lower waveguide boundary. This is similar to the SOI structure used almost universally today in silicon photonics, but the Figure 3: Proposed SOI p+-n-n+ channel retained contact to the buried n+ substrate allowed vertical current flow.
waveguide electro-optical phase modulator [91].

In the late 1980s, Friedman et al. [92] proposed and theoretically analysed phase modulators in a series of transistor structures integrated into rib waveguides, as shown in figure 4. The devices were based around MOSFETs, utilising injection of single carrier types (holes or electrons), or dual injection field effect transistors (DIFET). The injected charge in the DIFET was to be controlled by the junction field effect in which a voltage variable depletion width controlled the effective cross-sectional area of the conducting channel. The authors predicted effective refractive index changes of the order of 1 103, for applied gate voltages of 10 20 V. They also proposed two gate devices which offered better

Proc. of SPIE Vol. 5730

overlap of the modal field with injected carriers than a single gate device, although fabrication of two gates is impractical for potential SOI structures. They concluded that because the refractive index changes produced occurred over thin layers, the devices would be optimal in small waveguides, of the order of 0.1 m height. At the time this was regarded as unreasonably small, but as the current trend to smaller cross sectional dimensions continues, and nanophotonic devices are more seriously considered, some of these early devices are being reconsidered. It is also interesting to note that some of these devices bare a resemblance to the recent Intel modulators that achieved a bandwidth in excess of 1GHz [33].

Figure 4: Cross-section views of the proposed dual injection field effect transistors (DIFET) [92]. In 1987, Lorenzo et al. [93] reported the first 2 2 electro-optical switch in silicon operating at a wavelength of 1.3 m, shown in figure 5. The switch was fabricated as a vertical 4.6m p+n diode, in a silicon on doped silicon configuration. The injection of holes into the intersection of the crossing waveguides caused a change in refractive index, resulting in some limited switching. For an input current density of J = 1.26 kA/cm2 applied at the p+n junction, the device experienced an on state which switched 50% of the optical power from the output straight-through channel port 3 to the output cross channel port 4. Although this device was not optimised, it reaffirmed the feasibility of the plasma dispersion effect.

(a) Perspective view

(b) Active midsection

Figure 5: 2x2 guided-wave Si optical switch by injection of free carriers [93].

Proc. of SPIE Vol. 5730

Treyz et al. [94] reported silicon waveguide intensity modulators operating in the range of 1.3 to 1.55 m. A schematic diagram is shown in figure 6. The intrinsic region was grown to a thickness of 7.7 m and had a p-type doping concentration of less than 5 1015 cm3. The p+ layer had a doping concentration greater than 5 1019 cm3 and a thickness of 0.5 m. By applying a forward bias to the device, free carriers were injected into the p-type silicon (guiding) region of the device. The maximum modulation depth achieved was 6.2 dB (76%) for a high current density of 3.4 103 A/cm2. The authors measured the 90% 10% response time to be less than 50ns. The devices were multimode in both the vertical and horizontal directions. This device was further developed in another paper where the same authors demonstrated a similar device configured as a phase modulator [95]. The device was incorporated into one arm of a Mach-Zehnder Interferometer (MZI) to convert phase modulation into amplitude modulation. The devices also had response times of 50 ns (> 7MHz) with active region losses estimated to be 0.75 2.0 dB, compared to 0.4 1.0 dB [91] above. In common with the modulator of Lorenzo et al. [93], this waveguide experienced a high static loss due to the relatively weak optical confinement in the vertical direction and interaction of guided modes with the doped silicon substrate. Such interaction in the doped region results in optical absorption. The trend to fabrication of SOI based structures eliminated this problem, but also meant that vertical structures were much more difficult to fabricate, as a substrate contact could not be included. However, Jackson et al. [96, 97] later produced a similar vertical modulator based on SOI, utilizing a buried n+ contact at the bottom of the waveguide as shown in Figure 7. It is also interesting to note that the cross-sectional dimensions of this device are large, of the order of several microns. This was typical of modulation devices in the early and mid 1990s, largely to ensure good coupling from telecommunications optical fibres. However, more latterly in the late 1990s and early 2000s, there has been a trend toward much smaller dimensions, of the order of 1 m and below.
1 m 0.5 m 0.2 m 0.2 m Al p+ Oxide

p-type Silicon

n + Silicon Substrate
Figure 6. Schematic diagram of the active region. Rib height was 4.6 m with widths of 13 and 24 m. Diode lengths were 500 and 1000 m [94]. Figure 7. Vertical modulator demonstrated experimentally by Jackson et al. [96, 97]

In 1993, Soref et al, [30] also proposed devices to act as in line attenuators, and/or modulators, the forerunner to the Variable Optical Attenuators (VOAs) discussed frequently today. They coupled this discussion with another device, the Mode Displacement Modulator [98], which used the change in refractive index to displace the optical mode downwards and hence implement modulation/attenuation. They also identified direct carrier absorption as another mechanism to implement attenuation, the usual technique used to implement VOAs today [99]. Once again we see that the early work laid solid foundations for todays devices Work carried out at the University of Surrey by Tang et al. in 1994, in support of an earlier simulation paper in 1993 [100], produced an highly efficient phase modulator in a large single mode rib waveguide (Figure 8). This work was about improving the overlap between injected carriers and the propagating optical mode. For example, the authors showed that it was possible to obtain a 30% increase in the concentration of injected carriers into the waveguiding region of an SOI rib waveguide phase modulator [101, 102] by changing the sidewall angle of the rib from vertical to 54.7 degrees, as depicted in 8. The separation distance of the n+ injecting regions remained constant for both the vertical and angled rib waveguides. The device was first designed at the University of Surrey and fabricated in a University

Proc. of SPIE Vol. 5730

Table 1: A comparison of theoretical and experimental performance [100- 102].

Parameter
Current density, J (A/cm2) Active optical loss (dB) Current required for a -radian phase shift, I (mA) Refractive index change, n Peak carrier concentration, N (cm3) Figure of merit, ( cm/A)

Modelled device
120 1.37 4 9.24 104 2.5 1017 17.9

Experiment al device
175 1.3 7 1.55 103 4.6 1017 20.6
3.3

Anode
4.0

y z x
t = 3.2 0.5

p+

All dimensions in m

Cathode
n
+

Intrinsic Silicon
0.25 0.25

Cathode
n+
0.5

0.4

SiO 2

Figure 8: Three terminal phase modulator with angled rib walls [100-102].

of Surrey/University of Southampton collaboration [102]. The device was modelled using the 2-dimensional (2D) semiconductor device package MEDICI. Table 1 shows the comparison between the predicted theoretical and experimental performance of two modulators with different cross sectional areas [100-102]. From the table, it can be seen that despite the differences between the two modulator dimensions, the figure of merit is in reasonable agreement, and is a direct consequence of optimisation via the modelling exercise. At the time when these devices were fabricated, the typical current densities were of the order of kA/cm2 (e.g. [93]). Therefore, with an experimental drive current of 7 mA, and current density of 175 A/cm2, this device represented an improvement in the current density of approximately an order of magnitude. Modulation bandwidths were in the range 5 20 MHz for different device variants. Huang et al. [103] proposed and modelled a novel guided-wave modulator in 1993 which utilises an IMPATT type diode, to generate carriers by avalanche multiplication (Figure 9). The authors argued that the modulation speed of the injection-type of device is limited by the minority-carrier lifetime, (~s), and a depletion-type device can only provide modulation over a sub-micrometer device size. Hence they based their design on avalanche generation. Their device utilised two regions. The first was a small and high-field avalanche region where impact ionization occurs and carriers were generated. The second was a long and low-field drift region where generated carriers cause the refractive index to be modulated. Modulation speed is improved as the drifting carriers are majority carriers. The proposed modulator structure of n+-p+-p-p+ based on SOI is reproduced as below. The simulated doping concentrations for the n+ and p+ cladding were 1.5 1019 cm3 and 1019 cm3 respectively; while the avalanche p+ region was 3 1018 cm3. The authors modelled their devices using the SUPREM-IV process simulator [104], and found the switching time of their proposed device to be less than 1 ns, which implies a modulation bandwidth in the gigahertz regime. However, the predicted current density was two orders of magnitude higher than an injection device (~ 110 kA/cm2), and very lossy as lateral confinement was achieved via doped regions, although presumably an etched rib is also viable. Another switching device that employed the plasma dispersion effect was introduced in 1994 by Liu et al. [105]. The device was a 1x2 digital optical switch that had p-n junctions placed in each arm of a y-junction just after the junction itself (see figure 10). When a current was applied to forward bias one of the p-n junctions, the plasma dispersion effect would cause the index to change. This would cause a mode that was originally guided to cut off, thereby causing that waveguide to disappear. For maximum switching, an applied current of 290 mA was required. This corresponded to a 22.3 dB extinction ratio between the arms. The switching time was reported to be less than 0.2 ps, staggeringly fast for a waveguide device of 3.6m in height and 6m wide, switching such a large current. Unfortunately however, no switching data was provided, and no subsequent reports of the device appeared.

Proc. of SPIE Vol. 5730

Figure 9. Schematic diagram and doping profile of the SOI n+p+-p-p+ modulator [103]

Figure 10. A 1x2 digital optical switch using the plasma dispersion effect [105].

3.3

OTHER MODULATOR CONFIGURATIONS

Other forms of switching and modulation were envisaged during the time span covered in this paper, such as thermal switches and Micro Optical Electrical Mechanical Systems (MOEMS). Since these devices are inherently slower than devices based upon the plasma dispersion effect, they are covered in less detail in this paper, but a selection of example devices are given. Another interesting area is light-by-light control of switching and modulating devices. There are fewer examples of these devices, but a selection is also given here. Whilst interesting, they too would add to the complexity of any monolithic circuit (see [106-110]). As compared to the plasma dispersion effect, thermal modulation is attractive because it does not induce absorption associated with the phase shift, unless the metal used for heating is very close to the optical mode. However, thermal modulation is generally a slower mechanism. In 1991, Treyz [95] demonstrated one of the first thermo-optical devices on single-crystal silicon. He fabricated a Mach Zehnder Interferometer (MZI) with one arm coated with NiCr to act as a heater. The waveguide had a height of 1.1 m with an etch depth of 0.25 m. The length of the heater varied from 250 500 m and had two widths of 4 and 13 m (corresponding to 3 and 10 m waveguide widths). He was able to measure a 40% modulation depth on devices of both widths for an applied power of 150 mW for linearly polarized 0=1.319 m light. He was also able to measure the frequency dependence of his device. At a frequency of 20 kHz he reported an impressive modulation depth of 95%, albeit rather slow. In 1992 Cocorullo and Rendina demonstrated a thermo-optic modulator in silicon [111], albeit not in a guided wave configuration. They used a 300 m silicon wafer as a Fabry Perot etalon, by shining light with a wavelength of 1.523 m at normal incidence, and heating the entire assembly in a temperature controlled chamber. The cavity exhibited a finesse of 1.5, but due to slow response times of the environmental chamber, response times could not be reported. The thermo-optic coefficient of silicon was determined during this experiment to be 1.86 x 10-4/K, a figure now widely used for thermo-optic modulators. The same group produced a 700 kHz thermo-optic device in 1995, this time utilising a polysilicon heater on top of a large silicon strip waveguide, on a doped silicon substrate [112]. Once again the device was realised as a Fabry-Perot resonator, in the strip guide 4.5 m high, 15 m wide, and 100m long, with etched facets. Maximum modulation depth was ~60% for voltage pulses each providing ~0.2 J. In late 1995, Fischer et al. [113] reported the first 1x2 and 2x2 SOI rib waveguide switches operated using the thermooptic effect. Switching is achieved by heating one of the arms of an (MZI) as depicted in figure 11(a). The top schematic is the 1x2 switch and the bottom is the 2x2 switch. Balanced splitting of the beam was achieved by a Multi-Mode Interference (MMI) coupler. The phase shifter comprised a titanium heating element placed on top of the waveguide as shown in figure 11(b). Their devices required 85 mW of power to obtain a change in the output intensity of the 1x2 and

Proc. of SPIE Vol. 5730

2x2 switches of 7 dB and 8 dB respectively. They also obtained cross-talk values of -6 dB (1x2 switch) and -5 dB (2x2 switch) for 0=1.3 m. No values were given for the switching time, however.

(a)

(b)

Figure 11. (a) Schematic of a 1x2 (top) and 2x2 (bottom) switch fabricated by Fischer et al. (b) Cross-sectional view of (left) the waveguide dimensions used in the switches and (right) the phase shifter region. [113] One of the more interesting switches/modulators of the time was the Micro OptoElectronic Mechanical System (MOEMS) device proposed by Watts et al. [106]. Figure 12 shows the micro-cantilever beam (left) and micro-bridge (right) devices. By applying a voltage between the substrate and the bridge/cantilever deflection of the waveguide occurs due to the electrostatic force. This deflection causes guided light to be coupled out of the waveguide and into the substrate. It was predicted that losses of 60 dB could be achieved with such devices. Also, for modulation voltages less than 10V and lengths less than several hundred microns, modulation response times on the order of tens of microseconds could be realized. The cantilever style of device design was realised by Eng et al. in 1995 [114]. For a 4780 m cantilever beam they were able to obtain a contrast ratio of 12 dB for an applied voltage of 10V and approximately 40 dB for 20V. Whilst these devices are too Figure 12. MOEMS switches and modulators as slow for high-speed telecommunications devices, they may proposed by Watts et al. [106] prove useful for some system reconfiguration or sensor applications [115]. In 1986, Normandin [107] demonstrated the first all-optical switch in silicon. He demonstrated all-optical logic gates with a better than 20 dB contrast ratio. The switching properties of a similar device were investigated by Colbourne and Jessop in 1988 [108]. These devices induced free carriers by applying above bandgap laser pulses to a silicon waveguide. These free carriers induce an effective index change causing a below bandgap optical beam passing through the waveguide to refract. Thus, by pulsing the laser inducing the free carries, the below bandgap beam is modulated as well. A typical device is depicted in figure 13. Soref and Lorenzo also demonstrated a similar device in 1989 [109]. In their configuration, however, they used Figure 13. Light-by-light switching in silicon [108]. an SOI strip waveguide to confine the light. The photogenerated carriers were then used to absorb the light instead of refracting the beam as in the case of the previously discussed device. They achieved a 3.1% modulation depth for an applied 640 W/cm2 (0= 800 nm) pulse. Such devices are interesting as they may reduce monolithic fabrication steps by removing the need for contact doping, but may increase fabrication costs due to alignment and required circuitry of an external light source. However, these devices also bear some resemblance to the recent work of Almeida et al.

10

Proc. of SPIE Vol. 5730

[110], who used guided light (=1.535.6 m) to modulate a ring resonator via two photon absorption, and simultaneously overcome the alignment issues.

3.4 DETECTORS
Whilst a majority of this article has focused on devices comprised of single-crystal silicon, detectors are a necessary exception. Silicon is transparent at the telecommunication wavelengths. If monolithic single-crystal silicon devices are to be realised then methods such as implanting the silicon with germanium must be investigated. By implanting with large enough concentrations of germanium, the band-edge of the resulting alloy will be pushed further into the infrared wavelengths compared to undoped silicon. The result is an optically absorbent material at the telecommunications wavelengths. In order to move the bandgap far enough, however, a large amount of germanium is required. For example, a germanium fraction of greater than 30% is required to absorb 1.3 m light, but more than 85% is required for 1.55 m [30]. However it is impossible to include sufficient Ge in bulk silicon due to both solid solubility limits and/or prohibitive losses. Consequently many authors have considered multilayered structures. Furthermore, such a process may make the concept of an integrated photodetector prohibitively expensive. Nonetheless research was carried out to design efficient photodetectors that were compatible with standard silicon CMOS fabrication techniques. This is also an area of current research, with some early work influencing todays research. Before considering such devices, it is worth mentioning one detector comprised entirely of silicon [116]. It consisted of a p-n junction with a very highly doped region to narrow the bandgap of silicon, a heavily debated phenomenon at the time. The result was photodiode that was sensitive to 1.3 m light with an approximately 20 ps time resolution. Not surprisingly the device had a very low quantum efficiency of 10-7. However the device demonstrated that an all silicon detector could be achieved and whilst it efficiency was poor, it made for a respectable high-speed single-photon detector. There may also be some value in considering such an approach for power monitoring applications. Some of the earliest photodetectors with silicon as its primary material was proposed by researchers at AT&Ts Bell Laboratories [117119] in the mid-eighties. For example, Luryi et al. [119] proposed a Strained Layer Superlattice (SLS) detector consisting of GexSi1-x/Si interwoven layers (see figure 14). Figure 14. Cross-sectional view of an SLS photodetector in silicon [120]. Bandgap narrowing occurred due to the effect of the strained alloy layers. As a result of this narrowing, the bandgap is shifted so that absorption occurs beyond a wavelength of 1.3 m wavelength. The authors proposed both p-i-n detectors and APDs. The p-i-n devices were proposed to be 0.5 mm long and <10 m wide to maintain capacitance below 0.5 pF. In the APD design, the photogenerated electrons in the SLS are amplified by an avalanche region created in the silicon waveguide cladding. Luryi utilised a germanium fraction to be x 0.6 by considering the trade-off between the stability of the device and its absorption characteristics. Such a device was fabricated by Temkin et al. [120] in 1986. They reported an internal quantum efficiency () of ~40% at 0=1.3 m. For the device in figure 14, there were 20 wells of Ge0.6Si0.4, each 60 in thickness, separated by 290 thick Si barriers. If the Ge fraction was increased to 0.8 the well thickness had to be reduced to less than 20 . Leakage current ranged from 0.4A to as much as 3A for Ge fraction of 0.4 and 0.8 respectively. They also obtained a frequency bandwidth in excess of 1 GHz. Several months later they reported a similar device but with gain via an avalanche effect [121]. They investigated devices each having different a different Ge fraction (0.2-0.8). However they determined that a Ge content of 0.6 had the best efficiency for 1.3 m light. With this device they also measured a modulation bandwidth-gain product in the region of 3.2-3.7 GHz.

Proc. of SPIE Vol. 5730

11

In 1990, Kesan et al. [122] developed a similar detector with the exception of the avalanche photodiode. They, instead, used a reversed-biased P-I-N structure to collect the photogenerated carriers. Their device was based on a 28 period superlattice comprising 40 Si0.4Ge0.6/210 Si layers. What made their device important was that it was integrated with a silicon rib waveguide on SIMOX material (see figure 15). They measured a leakage current of 10-30 pA/m2 (@ 15V reverse bias potential) for a 260x260 m2 detector footprint. They also measured ~50% for 0=1.1 m, but this reduced to negligible proportions beyond 1.25m. They determined their waveguide loss to be on the order of 1-2 dB/cm at 0=1.3 m. Their frequency response bandwidth was on the order of 1-2 GHz, similar to the result obtained by Temkin. In the period 1992-93, Jalali et al. [123], [124] reported photodetectors similar to that of Kesan above. They reported and external quantum efficiency of ~5% @ 1Gbit/s for 0=0.96 m in 1992. Subsequent devices were designed for butt coupling to a fibre as an optical fibre interconnect as shown in figure 16. They used a rapid thermal chemical vapour deposition technique to grow their superlattice. Their device was based on a 14 period superlattice comprising 140 Si0.29Ge0.71/430 Si layers. They reported on an improved detector operating at 0=1.32 m with a frequency response of 1.5 Gbit/s, with an external quantum efficiency of 7%, and subsequently, a detector with a dark current of only 27 pA/m2 whilst still operating at 0=1.32 m . They were also reported an internal quantum efficiency of 33% based on estimates of coupling efficiency.

Figure 15. An integrated Si rib waveguidephotodetector on SIMOX [122].

Figure 16. Schematic of a waveguide P-I-N detector for use in an optical interconnect [123].

Figure 17. An integrated single-mode rib waveguide with a photodetector [125].

12

Proc. of SPIE Vol. 5730

In 1994, Splett et al. [125] [35] reported another version of the p-i-n SiGe MQW detector (figure 17) this time integrated with a single mode rib waveguide. They achieved an external quantum efficiency of up to 12% at 0=1.282 m. They also achieved a frequency bandwidth of 2 GHz and had a dark current of 200 nA, and an external quantum efficiency of 11% (internal quantum efficiency of 40%). Later that year they reported on a device [126] with a waveguide to detector coupling efficiency of 67%. They were also able to obtain 6.1 GHz operation, limited by capacitance. The reports of integrated detectors in the period of investigation have been remarkable. Growth and processing methods have greatly assisted in allowing high quality devices to be fabricated. It is interesting to note that the choice of detector approach is still debated, although this early work made very significant progress, and established a solid foundation for future work.

4. CONCLUSIONS
This paper has attempted to provide an overview of research in silicon photonics from the mid 1980s to the mid 1990s, in particular noting work towards fully-integrated monolithic silicon optical devices, or the so-called superchip. A brief history of the components that may comprise such a chip has been provided whilst focusing on the highlights made during this period. In providing such a summary it is impossible to be comprehensive, as a surprising range and depth of work was carried out, not only in single crystal silicon, but also in related technologies. However it is hoped that the reader has obtained a sense of the history of the formative years of silicon photonics and the range of work that was carried out. These ten years have shown incredible innovation and provided a solid platform upon which more recent work has developed. It is interesting to note that much of the early work has been revisited in recent years as trends in silicon photonics technology have changed. A number of recent developments have utilised devices that are similar to devices produced in the early years, both deliberately and accidentally. Therefore it is noteworthy that short-term research is not necessarily the most appropriate. It is of course even more noteworthy that silicon photonics is now progressing at a significant rate. Whilst the progress in the decade studied in this paper was significant, things are moving much more quickly now, partly due to the increased interest in silicon photonics, as well as the significantly increased resources worldwide. Consequently silicon photonics can look forward to an intense period of advanced study, and perhaps more of the devices studied in the early years will again prove instrumental.
ACKNOWLEDGEMENTS

The authors are grateful to the Intel Corporation and Bookham Technology for funding, and to Dr Richard Soref for helpful discussions.

REFERENCES
1. 2. 3. 4. Miller, S.E., Integrated optics: an Introduction. B.S.T.J., 1969. 48(7): p. 2059-2069. Hickernell, F.S., and C.T. Seaton. Channelized optical waveguides on silicon. in Proc. SPIE: Integration and Packaging of Optoelectronic Devices, 1986. 703, p 164 - 174. Henry, C.H. Recent advances in integrated optics on silicon. in EFOC/LAN 90, 27-29 June. 1990. Munich, West Germany. Henry, C.H., G.E. Blonder, and R.F. Kazarinov, Glass waveguides on silicon for hybrid optical packaging. Twelfth Conference on Optical Fiber Communications (OFC) and Topical Meetings on Integrated and GuidedWave Optics, Feb 6-9 1989, 1989. 7(10): p. 1530-1539. Nagata, T., T. Tanaka, K. Miyake, H. Kurotaki, S. Yokoyama, and M. Koyanagi, Micron-size optical waveguide for optoelectronic integrated circuit. JJAP Proc. of the 1993 International Conference on Solid State Devices and Materials (SSDM '93), Aug 29-Sep 02 1993, 1994. 33(1B): p. 822-826. Aarnio, J., P. Heimala, M. Del Giudice, and F. Bruno, Birefringence control and dispersion characteristics of silicon oxynitride optical waveguides. Electron. Lett., 1991. 27(25): p. 2317-18. Zurhelle, D., J.P. Schmidt, R. Hoffman, D. Sander, and J. Muller. Coupling structures for active and passive integrated opto-electronic components and circuits on silicon. in Proc. SPIE: Optoelectronic Integrated Circuit Materials, Physics, and Devices, 6-9 Feb. 1995. San Jose, CA, USA. Gehler, J., A. Brauer, W. Karthe, and M. Jager, Antiresonant reflecting optical waveguides in strip configuration. App. Phys. Lett., 1994. 64(3): p. 276-8.

5.

6. 7.

8.

Proc. of SPIE Vol. 5730

13

9.

10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20.

21. 22.

23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34.

Muller, J., U.-P. Dahms, M. Mahnke, and S. Wunderlich. Refractive index relaxation in PECVD- and LPCVDSiON-waveguides on silicon substrates. in Proc. of ECIO'95 - 7th European Conference on Integrated Optics, 3-6 April. 1995. Delft, Netherlands: Delft University Press. Wunderlich, S., J.P. Schmidt, and J. Muller, Integration of SiON waveguides and photodiodes on silicon substrates. Appl. Opt., 1992. 31(21): p. 4186-9. Gleine, W., and J. Muller, Laser trimming of SiON components for integrated optics. J. of Light. Tech., 1991. 9(11): p. 1626-1629. Bezzaoui, H. and E. Voges, Integrated optics combined with micromechanics on silicon. Sensors and Actuators A (Physical), 1991. A29(3): p. 219-23. Tu, Y.-K., J.-C. Chou, and S.-P. Cheng, Single-mode SiON/SiO/sub 2//Si optical waveguides prepared by plasma-enhanced chemical vapor deposition. Fiber and Integrated Optics, 1995. 14(2): p. 133-9. Bossi, D.E., J.M. Hammer, and J.M. Shaw, Optical properties of silicon oxynitride dielectric waveguides. Appl. Opt., 1987. 26(4): p. 609-11. Najafi, S.I., S. Honkanen, and A. Tervonen. Recent progress in glass integrated optical circuits. in Proc. SPIE: Integrated Optics and Microstructures II, 26-28 July 1994. San Diego, CA, USA. Allen, J.J., S.P. Shipley, T.M. Ong, and N. Nourshargh, All-silica integrated optical components. GEC J. Res. Inc. Marconi Rev. & Plessey Res. Rev., 1993. 10(2): p. 96-100. Shimizu, T., S. Nakamura, K. Ueki, and I. Ohyama, Silica-based waveguide devices. Furukawa Review, 1992(11): p. 50-54. Habara, K. and T. Matsunaga. Silica glass waveguide planar lightwave circuit applications in photonic switching. in Proc. SPIE: 1992 Topical Meeting on Photonic Switching, 1-3 July. 1993. Minsk, Byelorussia. Grant, M.F. Integrated optical waveguide devices on silicon for optical communications. in IEE Colloquium `Planar Silicon Hybrid Optoelectronics' (Digest No.1994/198), 24 Oct. 1994. London, UK: IEE. Splett, A., J. Schmidtchen, B. Schuppert, and K. Petermann. Integrated optical channel waveguides in silicon using SiGe-alloys. in Proc. SPIE: Physical Concepts of Materials for Novel Optoelectronic Device Applications II: Device Physics and Applications, 28 Oct.-2 Nov. 1991. Aachen, Germany. Weiss, B.L., Z. Yang, and F. Namavar, Wavelength dependent propagation loss characteristics of SiGe/Si planar waveguides. Electron. Lett., 1992. 28(24): p. 2218-20. Kesan, V.P., P.G. May, F.K. LeGoues, and S.S. Iyer, Si/SiGe heterostructures grown on SOI substrates by MBE for integrated optoelectronics. Journal of Crystal Growth: J. Cryst. Growth: Sixth International Conference on Molecular Beam Epitaxy, 27-31 Aug., 1991. 111(1-4): p. 936-42. Namavar, F. and R.A. Soref, Optical waveguiding in Si/Si/sub 1-x/Ge/sub x//Si heterostructures. J. of Appl. Phys., 1991. 70(6): p. 3370-2. Soref, R.A., J. Schmidtchen, and K. Petermann, Large Single-Mode Rib Waveguides in GeSi-Si and Si on SiO2. IEEE J. of Quantum Elect., 1991. 27(8): p. 1971-1974. Jackson, S.M., G.T. Reed, and K.J. Reeson, Waveguiding in epitaxial 3C-silicon carbide on silicon. Electron. Lett., 1995. 31(17): p. 1438-9. Liu, Y.M. and P.R. Prucnal, Low-loss silicon carbide optical waveguides for silicon-based optoelectronic devices. IEEE Photonics Technol. Lett., 1993. 5(6): p. 704-7. Tang, X., K. Wongchotigul, and M.G. Spencer, Optical waveguide formed by cubic silicon carbide on sapphire substrates. Appl. Phys. Lett., 1991. 58(9): p. 917-18. Iyer, S.S., and Y.-H. Xie, Light Emission from Silicon. Science, 1993. 260(5104): p. p.40-46. Abstreiter, G., Engineering the future of electronics. Physics World, 1992. 5(3): p. 36-9. Soref, R.A., Silicon-based optoelectronics. Proc. IEEE, 1993. 81(12): p. 1687-706. Boyraz, O., and B. Jalali, Demonstration of a silicon Raman laser. Opt. Express, 2004. 12(21): p. 5269-5273. Liu, A., H. Rong, M. Paniccia, O. Cohen, and D. Hak, Net optical gain in a low loss silicon-on-insulator waveguide by stimulated Raman scattering. Opt. Express, 2004. 12(18): p. 4261-4268. Liu, A., R. Jones, L. Liao, D. Samara-Rubio, D. Rubin, O. Cohen, R. Nicolaescu, and M. Paniccia, A high-speed silicon optical modulator based on a metal-oxide-semiconductor capacitor. Nature, 2004. 427(6975): p. 615-18. Png, C.E., S.P. Chan, S.T. Lim, and G.T. Reed, Optical phase modulators for MHz and GHz modulation in silicon-on-insulator (SOI). J. Lightwave Technol., 2004. 22(6): p. 1573-82.

14

Proc. of SPIE Vol. 5730

35.

36. 37.

38. 39. 40. 41.

42. 43.

44. 45. 46. 47. 48.

49. 50. 51. 52. 53. 54. 55. 56. 57.

Splett, A., Th. Zinke, B. Schuppert, K. Petermann, H. Kibbel, H.-J. Herzog, and H. Presting. Integrated optoelectronic waveguide-detectors in SiGe for optical communications. in Proc. SPIE: Photodetectors and Power Meters II, 11-12 July. 1995. San Diego, CA, USA. http://www.darpa.mil/mto/solicitations/baa04-15/s/epic_web.pd. Armiento, C.A., M.J. Tabasky, J. Chirravuri, M.A. Rothman, A.N.M.M. Choudhury, A.J. Negri, A.J. Budman, T.W. Fitzgerald, V.J. Barry, and P.O. Haugsjaa. Hybrid optoelectronic integration of transmitter arrays on silicon waferboard. in Proc. SPIE: Integrated Optoelectronics for Communication and Processing, 3-4 Sept. 1992. Boston, MA, USA. Ayliffe, P., J. Parker, S. Bertolini, T. Clapp, M. Geear, P. Harrison, and R. Peall, The hybrid integration of optical and electronic components on silicon. Int. J. Optoelectron., 1994. 9(2): p. 179-91. Blackie, G.N. and I.R. Croston, Silicon optohybrids for advanced optoelectronic multi-chip modules. GEC J. Res. Inc. Marconi Rev. & Plessey Res. Rev., 1993. 10(2): p. 106-10. Fincato, A., S. Lorenzotti, P. Nugent, G. Parafioriti, and G. Randone, Glass on silicon technology for optical interconnections and optoelectronic hybrid integration. Inf. MIDEM, 1994. 24(3): p. 151-60. Hornak, L.A., S.K. Tewksbury, and T.W. Weidman. Toward cointegration of optical interconnection networks within silicon multichip systems. in Advances in Optical Information Processing V, Apr 21-24 1992. Orlando, FL, USA: Publ by Int Soc for Optical Engineering, Bellingham, WA, USA. Hunsperger, R.G., Integrated Optics: Theory and Technology. 3rd ed. Springer Series in Optical Scinces, ed. T. Tamir. Vol. 33. 1991, Berlin: Springer-Verlag. 347. Iezekiel, S., E.A. Soshea, M.F.J. O'Keefe, and C.M. Snowden. Glass-silicon substrate for hybrid optoelectronic packaging. in IEE Colloquium `Planar Silicon Hybrid Optoelectronics' (Digest No.1994/198), 24 Oct. 1994. London, UK: IEE. Jones, C.A., K. Cooper, M.W. Nield, and J.D. Rush. Hybrid integration using silicon optical motherboards. in IEE Colloquium on `Microengineering and Optics' (Digest No.1994/043), 17 Feb. 1994. London, UK: IEE. Kobayashi, M. and K. Kato, Hybrid optical integration technology. Trans. Inst. Electron. Inf. Commun. Eng. CI, 1994. J77C-I(5): p. 340-51. Miyashita, T., S. Sumida, and S. Sakaguchi. Integrated optical devices based on silica waveguide technologies. in Proc. SPIE: Integrated Optical Circuit Engineering VI, 8-9 Sept. 1988. Boston, MA, USA. Runge, K., M. Bagheri, and J. Young, High performance hybrid circuit modules for lightwave systems operating at data rates of 10 Gbit/s and higher. Electron. Lett., 1991. 27(3): p. 267-70. Solgaard, O., N.C. Tien, M. Daneman, M.-H. Kiang, A. Friedberger, R.S. Muller, and K.Y. Lau. Precision and performance of polysilicon micromirrors for hybrid integrated optics. in Proc. SPIE: MicroOptics/Micromechanics and Laser Scanning and Shaping, 7-9 Feb. 1995. San Jose, CA, USA. Schuppert, B., J. Schmidtchen, A. Splett, U. Fischer, T. Zinke, R. Moosburger, and K. Petermann, Integrated optics in silicon and SiGe-heterostructures. J. Lightwave Technol., 1996. 14(10): p. 2311-2323. Valette, S., J.P. Jadot, P. Gidon, S. Renard, A. Fournier, A.M. Grouillet, H. Denis, P. Philippe, and E. Desgranges, Si-based integrated optics technologies. Solid State Technol., 1989. 32(2): p. 69-75. Polky, J.N. and G.L. Mitchell, Metal-clad planar dielectric waveguide for integrated optics. J. Opt. Soc. Amer., 1974. 64(3): p. 274-279. Nezval, J., WKB Approximation for optical modes in a periodic planar waveguide. Opt. Commun., 1982. 42(5): p. 320-322. de Ruiter, H.M., Limits on the propagation constants of planar optical waveguide modes. Appl. Opt., 1981. 20(5): p. 731-2. Tamir, T., Leaky waves in planar optical waveguides. Nouvelle Revue d'Optique, 1975. 6(5): p. 273-284. Payne, F.P. Generalized transverse resonance model for planar optical waveguides. in Optical Communication, ECOC '84: Tenth European Conference. 1984. Stuttgart, Austria. Kawachi, M., M. Yasu, and T. Edahiro, Fabrication of SiO2-TiO2 glass planar optical waveguides by flame hydrolysis deposition. Electron. Lett., 1983. 19(15): p. 583-584. Hanaizumi, O., M. Miyagi, and S. Kawakami, Low radiation loss y-junctions in planar dielectric optical waveguides. Opt. Commun., 1984. 51(4): p. 236-238.

Proc. of SPIE Vol. 5730

15

58. 59. 60. 61. 62.

63.

64. 65. 66. 67.

68. 69.

70. 71. 72. 73. 74.

75. 76. 77. 78. 79.

Jerominek, H., Z. Opilski, and J. Kadziela, Some elements of integrated optics circuits based on planar gradient glass waveguides. Optica Applicata, 1983. 13(2): p. 159-68. Soref, R.A., and J.P. Lorenzo, All-Silicon Active and Passive Guided-Wave Components for =1.3 and 1.6 m. IEEE J. of Quantum Elect., 1986. QE-22(6): p. 873-879. Soref, R.A. and J.P. Lorenzo, Single-crystal silicon: a new material for 1.3 and 1.6 m integrated-optical components. Electron. Lett., 1985. 21(21): p. 953-954. Albares, D.J. and R.A. Soref. Silicon-on-sapphire waveguides. in Proc. SPIE: Integrated Optical Circuit Engineering IV, 1987, 704, p. 24 -25. Cortesi, E., F. Namavar, and R.A. Soref. Novel silicon-on-insulator structures for silicon waveguides. in 1989 IEEE SOS/SOI Technology Conference (Cat. No.89CH2796-1), 3-5 Oct. 1989. 1989. Stateline, NV, USA: IEEE. Namavar, F., E. Cortesi, R.A. Soref, and P. Sioshansi. On the formation of thick and multiple layer SIMOX structures and their applications. in Ion Beam Processing of Advanced Electronic Materials Symposium, 25-27 April 1989. 1989. San Diego, CA, USA: Mater. Res. Soc. Reed, G.T., L. Jinhua, C.K. Tang, L. Chenglu, P.L.F. Hemment, and A.G. Rickman, Silicon on insulator optical waveguides formed by direct wafer bonding. Mat. Sci. Eng. B: Solid, 1992. B15(2): p. 156-159. Evans, A.F., D.G. Hall, and W.P. Maszara, Propagation loss measurements in silicon-on-insulator optical waveguides formed by the bond-and-etchback process. Appl. Phys. Lett., 1991. 59(14): p. 1667-9. Colinge, J.-P., Silicon On Insulator Technology: Materials to VLSI. 2nd ed. 1997: Kluwer Academic Publishers. Mohd Kassim, N., H.P. Ho, T.M. Benson, and D.E. Davies. Assessment of SIMOX material by optical waveguide losses. in ESSDERC 90. 20th European Solid State Device Research Conference, 10-13 Sept. 1990. Nottingham, UK: Adam Hilger. Weiss, B.L. and G.T. Reed, The transmission properties of optical waveguides in SIMOX structures. Opt. Quant. Electron., 1991. 23(8): p. 1061-5. Davies, D.E., M. Burnham, T.M. Benson, N.M Kassim, and M. Seifouri. Optical waveguides and SIMOX characterisation. in 1989 IEEE SOS/SOI Technology Conference (Cat. No.89CH2796-1), 3-5 Oct. 1989. 1989. Stateline, NV, USA: IEEE. Soref, R.A., E. Cortesi, F. Namavar, and L. Friedman, Vertically integrated silicon-on-insulator waveguides. IEEE Photonic. Tech. L., 1991. 3(1): p. 22-4. Weiss, B.L., Reed, G.T., Toh, S.K., Soref, R.A., and F. Namavar, Optical waveguides in SIMOX structures. IEEE Photonic. Tech. L., 1991. 3(1): p. 19-21. Rickman, A., G.T. Reed, B.L. Weiss, F, Namavar, Low-loss planar optical waveguides fabricated in SIMOX material. IEEE Photonic. Tech. L., 1992. 4(6): p. 633-635. Kurdi, B.N. and D.G. Hall, Optical waveguides in oxygen-implanted buried-oxide silicon-on-insulator structures. Opt. Lett., 1988. 13(2): p. 175-7. Schmidtchen, J., A. Splett, B. Schuppert, and K. Petermann. Low loss integrated-optical rib-waveguides in SOI. in 1991 IEEE International SOI Conference, Oct 1-3 1991. 1992. Vail Valley, CO, USA: Publ by IEEE, Piscataway, NJ, USA. Rickman, A.G., G.T. Reed, and F. Namavar, Silicon-on-insulator optical rib waveguide loss and mode characteristics. J. Lightwave Technol., 1994. 12(10): p. 1771-1776. Rickman, A.G., and G.T. Reed, Silicon-on-insulator optical rib waveguides: loss, mode characteristics, bends and y-junctions. IEE Proc.-Optoelectron., 1994. 141(6): p. 391-393. Tang, C.K., A.K. Kewell, G.T. Reed, A.G. Rickman, and F. Namavar, Development of a library of low-loss silicon-on-insulator optoelectronic devices. IEE Proc.-Optoelectron., 1996. 143(5): p. 312-315. Rickman, A.G., G.T. Reed, and F. Namavar, Silicon-on-insulator optical rib waveguide circuits for fibre optic sensors. Proc. SPIE: Distributed and Multiplexed Fiber Optic Sensors III, 8-9 Sept., 1993. 2071: p. 190-6. Sure, A., T. Dillon, J. Murakowski, C. Lin, D. Pustai, and D.W. Prather, Fabrication and characterization of three-dimensional silicon tapers. Opt. Express, 2003. 11(26): p. 3555-3561.

16

Proc. of SPIE Vol. 5730

80.

81. 82. 83. 84. 85. 86.

87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100.

101. 102. 103. 104.

Lu, Z., P. Yao, S. Venkataraman, D. Pustai, C. Lin, G. Schneider, J. Murakowski, S. Shi, and D.W. Prather. Fiber-to-waveguide evanescent coupler for planar integration of silicon optoelectronic devices. in Proc. SPIE: Photonics Packaging and Integration IV, Jan 29 2004. 2004. San Jose, CA, United States. Masanovic, G.Z., V.M.N. Passaro, and G.T. Reed, Dual grating-assisted directional coupling between fibers and thin semiconductor waveguides. IEEE Photonic. Tech. L., 2003. 15(10): p. 1395-1397. Almeida, V.R., R.R. Panepucci, and M. Lipson, Nanotaper for compact mode conversion. Opt. Lett., 2003. 28(15): p. 1302-1304. Lipson, M., Overcoming the limitations of microelectronics using Si nanophotonics: Solving the coupling, modulation and switching challenges. Nanotechnology, 2004. 15(10): p. 622-627. Shoji, T., T. Tsuchizawa, T. Watanabe, K. Yamada, and H. Morita, Low loss mode size converter from 0.3 m2 Si wire waveguides to singlemode fibres. Electron. Lett., 2002. 38(25): p. 1669-70. Petermann, K., Properties of optical rib-guides with large cross-section. Archiv fur Electronik und Ubertragungstechnik (Germany), 1976. 30: p. 139-140. Chan, S.P., C.E. Png, S.T. Lim, G.T. Reed, and V.M.N. Passaro. Single mode, polarisation independent waveguides in silicon-on-insulator. in Proceedings of International Conference on Group IV Photonics, Oct. 2004. Hong Kong. Pogossian, S.P., L. Vescan, and A. Vonsovici, The single-mode condition for semiconductor rib waveguides with large cross section. J. Lightwave Technol., 1998. 16(10): p. 1851-1853. Powell, O., Single-mode conditions for silicon rib waveguides. J. Lightwave Technol., 2002. 20(10): p. 18511855. Powell, O., Erratum: Single-mode condition for silicon rib waveguides (Journal of Lightwave Technology (Oct. 2002) 20 (1851-1855)). J. Lightwave Technol., 2003. 21(3): p. 868. Soref, R.A., and B.R. Bennett, Electrooptical Effects in Silicon. IEEE J. Quantum Elect., 1987. QE-23(1): p. 123-129. Soref, R.A. and B.R. Bennett. Kramers-Kronig analysis of electro-optical switching in silicon. in Proc.SPIE: Integrated Optical Circuit Engineering IV, 16-17 Sept. 1986. 1987. Cambridge, MA, USA. Friedman, L., R.A. Soref, and J.P. Lorenzo, Silicon double-injection electro-optic modulator with junction gate control. J. Appl. Phys., 1988. 63(6): p. 1831-9. Lorenzo, J.P. and R.A. Soref, 1.3 m electro-optic silicon switch. Appl. Phys. Lett., 1987. 51(1): p. 6-8. Treyz, G.V., P.G. May, and J.-M. Halbout, Silicon Mach-Zehnder waveguide interferometers based on the plasma dispersion effect. Appl. Phys. Lett., 1991. 59(7): p. 771-3. Treyz, G.V., Silicon Mach-Zehnder waveguide interferometers operating at 1.3 m. Electron. Lett., 1991. 27(2): p. 118-120. Jackson, S.M., G.T. Reed, C.K. Tang, A.G.R. Evans, J. Clark, C. Aveyard, and F. Namavar, Optical beamsteering using integrated optical modulators. J. Lightwave Technol., 1997. 15(12): p. 2259-2263. Jackson, S.M., P.D. Hewitt, G.T. Reed, C.K. Tang, A.G.R. Evans, J. Clark, C. Aveyard, and F. Namavar, Novel optical phase modulator design suitable for phased arrays. J. Lightwave Technol., 1998. 16(11): p. 2016-2019. Pirnat, T. and L. Friedman, Electro-optic mode-displacement silicon light modulator. J. Appl. Phys., 1991. 70(6): p. 3355-9. Kotura, Inc., 2630 Corporate Place Monterey Park, CA 91754. www.lightcross.com. Tang, C.K., G.T. Reed, A.J. Wilson, and A.G. Rickman. Simulation of a low loss optical modulator for fabrication in SIMOX material. in Silicon-Based Optoelectronic Materials Symposium, 12-14 April 1993. 1993. San Francisco, CA, USA: Mater. Res. Soc. Tang, C.K., G.T. Reed, A.J. Wilson, and A.G. Rickman, Low-loss, single-mode, optical phase modulator in SIMOX material. J. Lightwave Technol., 1994. 12(8): p. 1394-1400. Tang, C.K. and G.T. Reed, Highly efficient optical phase modulator in SOI waveguides. Electron. Lett., 1995. 31(6): p. 451-2. Huang, H.C. and T.C. Lo, Simulation and analysis of silicon electro-optic modulators utilizing the carrierdispersion effect and impact-ionization mechanism. J. Appl. Phys., 1993. 74(3): p. 1521-1528. Synopsys, Inc. 700 East Middlefield Road, Mountain View, CA 94043, USA.

Proc. of SPIE Vol. 5730

17

105. 106.

107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118.

119. 120. 121. 122.

123. 124.

125. 126.

Liu, Y.L., E.K. Liu, S.L. Zhang, G.Z. Li, and J.S. Luo, Silicon 1x2 digital optical switch using plasma dispersion. Electron. Lett., 1994. 30(2): p. 130-131. Watts, R., A.L. Robinson, and R.A. Soref. Electromechanical optical switching and modulation in micromachined silicon-on-insulator waveguides. in 1991 IEEE International SOI Conference, Oct 1-3 1991. 1992. Vail Valley, CO, USA: Publ by IEEE, Piscataway, NJ, USA. Normandin, R., All-optical, fiber-optic modulator and logic gate using nonlinear refraction and absorption. Opt. Lett., 1986. 11(11): p. 751-3. Colbourne, P.D. and P.E. Jessop, Recovery time for a silicon waveguide all-optical switch. Electron. Lett., 1988. 24(6): p. 303-4. Soref, R.A., and J.P. Lorenzo, Light-by-Light Modulation in Silicon-on-Insulator Waveguides. OSA Integrated Guided Wave Optics '89, February 1989. MEE1: p. 86-89. Almeida, V.R., C.A. Barrios, R.R. Panepucci, and M. Lipson, All-optical control of light on a silicon chip. Nature, 2004. 431(7012): p. 1081-1084. Cocorullo, G. and I. Rendina, Thermo-optical modulation at 1.5 m in silicon etalon. Electron. Lett., 1992. 28(1): p. 83-85. Cocorullo, G., M. Iodice, I. Rendina, and P.M. Sarro, Silicon thermooptical micromodulator with 700-kHz - 3dB bandwidth. IEEE Photonic. Tech. L., 1995. 7(4): p. 363-365. Fischer, U., T. Zinke, and K. Petermann. Integrated optical waveguide switches in SOI. in Proceedings of the 1995 IEEE International SOI Conference, Oct 3-5 1995. 1995. Tucson, AZ, USA: IEEE, Piscataway, NJ, USA. Eng, T.T.H., S.Y.S. Sin, S.C. Kan, and G.K.L. Wong, Micromechanical optical switching with voltage control using SOI movable integrated optical waveguides. IEEE Photonic. Tech. L., 1995. 7(11): p. 1297-1299. Churenkov, A.V., Silicon micromechanical optical waveguide for sensing and modulation. Sensor. Actuat. APhys., 1996. 57(1): p. 21-27. Ghioni, M., A. Lacaita, G. Ripamonti, and S. Cova, All-silicon avalanche photodiode sensitive at 1.3 m with picosecond time resolution. IEEE J. Quantum Elect., 1992. 28(12): p. 2678-2681. Luryi, S., A. Kastalsky, and J.C. Bean, New infrared detector on a silicon chip. IEEE T. Electron Dev., 1984. ED-31(9): p. 1135-9. Kastalsky, A., S. Luryi, J.C. Bean, and T.T. Sheng. Single-crystal Ge/Si infrared photodetector for fiber-optics communications. in Extended Abstracts, Spring Meeting - Electrochemical Society. 1985. Toronto, Ont, Can: Electrochemical Soc, Pennington, NJ, USA. Luryi, S., T.P. Pearsall, H. Temkin, and J.C. Bean, Waveguide infrared photodetectors on a silicon chip. IEEE Electron Dev. L., 1986. ED-7(2): p. 104-107. Temkin, H., T.P. Pearsall, J.C. Bean, R.A. Logan, and S. Luryi, GexSi1-x strained-layer superlattice waveguide photodetectors operating near 1.3 m. Appl. Phys. Lett., 1986. 48(15): p. 963-5. Temkin, H., J.C. Bean, T.P. Pearsall, N.A. Olsson, and D.V. Lang, High photoconductive gain in GexSi1-x/Si strained-layer superlattice detectors operating at =1.3 m. Appl. Phys. Lett., 1986. 49(3): p. 155-7. Kesan, V.P., et al. Integrated waveguide-photodetector using Si/SiGe multiple quantum wells for long wavelength applications. in 1990 International Electron Devices Meeting, Dec 9-12 1990. 1990. San Francisco, CA, USA: Publ by IEEE, Piscataway, NJ, USA. Jalali, B., A.F.J. Levi, F. Ross, and E.A. Fitzgerald, SiGe waveguide photodetectors grown by rapid thermal chemical vapour deposition. Electron. Lett., 1992. 28(3): p. 269-271. Jalali, B., L. Naval, A.F. Levi, and P. Watson. GeSi infrared photodetectors grown by rapid thermal CVD. in Rapid Thermal and Laser Processing, Sep 24-25 1992. 1993. San Jose, CA, USA: Publ by Int Soc for Optical Engineering, Bellingham, WA, USA. Splett, A., T. Zinke, K. Petermann, E. Kasper, H. Kibbel, H.-J. Herzog, and H. Presting, Integration of waveguides and photodetectors in SiGe for 1.3 m operation. IEEE Photonic. Tech. L., 1994. 6(1): p. 59-61. Splett, A. and K. Petermann. Ultimate performance of SiGe/Si multi-quantum-well waveguide-photodetector combinations. in Proceedings of ECOC'94. 20th European Conference on Optical Communications, 25-29 Sept. 1994. 1994. Firenze, Italy: Ist. Int. Comun.

18

Proc. of SPIE Vol. 5730

Das könnte Ihnen auch gefallen