optical interconnections in digital systems—status and prospects

7
Optical interconnections in digital systems— Status and prospects By D. Z. Tsang I n computing, trends toward paral- lel processing, multiprocessor computers, higher operating speeds, and larger as well as denser very large scale integrated (VLSI) circuits have provided an impetus for the development of advanced inter- connection technology. Multiproces- sor computer designs tend to require both a larger number and a higher density of interconnections with a higher degree of connectivity than single-processor computers. The com- bination of higher operating speed, longer distances, and higher density leads to problems in maintaining sig- nal integrity because of parasitic react- ances, impedance mismatches, cross- talk, dispersion, and frequency- dependent skin effect losses. Also, as chip designs become larger and dens- er, the so-called pinout problem, in which the number of input/output pins and pads is limited by the density of wirebonds, is of increasing concern to VLSI circuit designers. Optical interconnections have been of interest at all levels in digital com- puters for applications between main- frames, modules, boards, 1 chips, 2-4 and even points within a chip. 5 They are viewed both as a technology to be used in conjunction with electrical interconnections and electronic logic circuits and as an enabling technology for optical computers. These inter- connections are expected to provide advantages in situations that require high speed, high interconnection density, high connectivity, long dist- ance, and the flexibility of optics. The basic characteristics and advantages of optical interconnections, the cur- rent status of work in the field, and the prospective performance of opti- cal interconnection systems are des- cribed here. Present high-speed digital systems are limited by the packaging and electrical interconnection technol- ogy. 6 While the fastest transistors have switching times of the order of 10 ps, the fastest commercial integrated cir- cuits (GaAs logic) are limited by package parasitics to rise and fall times of the order of 100 ps. Signal propa- gation within circuit boards results in some additional deterioration. The electrical backplane, into which boards are inserted, can only transfer rise times of 1 ns between boards. In effect, interconnection limitations re- strict current clock cycle times to the order of 10 ns, although reduced times are promised in the next gener- ation of machines. Optical interconnection configurations The recent interest in optical inter- connections was stimulated in 1984 by Goodman et al., 2 who presented ideas for the incorporation of optics into VLSI systems. Since then, many others have studied various imple- mentations of optical interconnec- tions, all of which can be partitioned into a source, a channel, and a receiver (Figure 1). The source can be a directly mod- ulated semiconductor laser or light- emitting diode or a cw laser coupled to an external modulator. The exter- nal modulator can be electro-optic, absorptive, reflective, or even a spatial light modulator, the latter of which can accommodate two-dimensional arrays of optical channels. A special case of the absorptive modulator is the self-electro-optic effect device (SEED) being pursued at A T & T . 7 The optical channel or channels can consist of free space with lenses or holograms acting to direct the light. It is also possible to use a guided-wave channel, examples of which include single or multimode optical fibers or waveguides. Polymer waveguides, photolithographically defined on a circuit board, are being pursued by Sullivan 8 and Hartman et al. 9 The receiver is generally a photo - diode or photoconductor connected to bias and preamplifier circuitry. A variation is to use a modulator as a combined detector and switching el- ement, an example of which is the FIGURE 1. Alternatives for the source, channel, and re- ceiver components of optical interconnection systems. OPTICS & PHOTONICS NEWS • OCTOBER 1990 23

Upload: d-z

Post on 08-Oct-2016

215 views

Category:

Documents


2 download

TRANSCRIPT

Optical interconnections in digital systems— Status and prospects

By D. Z. Tsang

In computing, trends toward paral­lel processing, multiprocessor computers, higher operating

speeds, and larger as well as denser very large scale integrated (VLSI) circuits have provided an impetus for the development of advanced inter­connection technology. Multiproces­sor computer designs tend to require both a larger number and a higher density of interconnections with a higher degree of connectivity than single-processor computers. The com­bination of higher operating speed, longer distances, and higher density leads to problems in maintaining sig­nal integrity because of parasitic react­ances, impedance mismatches, cross­talk, dispersion, and frequency-dependent skin effect losses. Also, as chip designs become larger and dens­er, the so-called pinout problem, in which the number of input/output pins and pads is limited by the density of wirebonds, is of increasing concern to VLSI circuit designers.

Optical interconnections have been of interest at all levels in digital com­puters for applications between main­frames, modules, boards,1 chips, 2 - 4

and even points within a chip. 5 They are viewed both as a technology to be used in conjunction with electrical interconnections and electronic logic circuits and as an enabling technology for optical computers. These inter­connections are expected to provide advantages in situations that require high speed, high interconnection density, high connectivity, long dist­ance, and the flexibility of optics. The

basic characteristics and advantages of optical interconnections, the cur­rent status of work in the field, and the prospective performance of opti­cal interconnection systems are des­cribed here.

Present high-speed digital systems are limited by the packaging and electrical interconnection technol­ogy. 6 While the fastest transistors have switching times of the order of 10 ps, the fastest commercial integrated cir­cuits (GaAs logic) are limited by package parasitics to rise and fall times of the order of 100 ps. Signal propa­gation within circuit boards results in some additional deterioration. The electrical backplane, into which boards are inserted, can only transfer rise times of 1 ns between boards. In effect, interconnection limitations re­strict current clock cycle times to the order of 10 ns, although reduced times are promised in the next gener­

ation of machines.

Optical interconnection configurations

The recent interest in optical inter­connections was stimulated in 1984 by Goodman et al.,2 who presented ideas for the incorporation of optics into VLS I systems. Since then, many others have studied various imple­mentations of optical interconnec­tions, all of which can be partitioned into a source, a channel, and a receiver (Figure 1).

The source can be a directly mod­ulated semiconductor laser or light-emitting diode or a cw laser coupled to an external modulator. The exter­nal modulator can be electro-optic, absorptive, reflective, or even a spatial light modulator, the latter of which can accommodate two-dimensional arrays of optical channels. A special case of the absorptive modulator is the self-electro-optic effect device (SEED) being pursued at A T & T . 7

The optical channel or channels can consist of free space with lenses or holograms acting to direct the light. It is also possible to use a guided-wave channel, examples of which include single or multimode optical fibers or waveguides. Polymer waveguides, photolithographically defined on a circuit board, are being pursued by Sullivan 8 and Hartman et al.9

The receiver is generally a photo­diode or photoconductor connected to bias and preamplifier circuitry. A variation is to use a modulator as a combined detector and switching el­ement, an example of which is the

FIGURE 1. Alternatives for the source, channel, and re­ceiver components of optical interconnection systems.

OPTICS & PHOTONICS NEWS • OCTOBER 1990 23

S E E D , or to include a wavelength-se­lective filter in front of the detector, as is under consideration at I B M ( A T & T and Bellcore are also pursuing this for other purposes).

In any of the combinations, one source is combined with one of the channels and one of the receivers to form the optical interconnection. The central idea behind all of the combi­nations is to minimize the length of electrical line, preferably a short wire, between the digital chip or driver and the source, so that the signal is con­verted to light before significant deg­radation. A similar argument applies to the receiver. For the single optical channel this has many advantages. Impedance matching, reflections, and ground planes should be less of a problem than with high-speed electri­cal interconnections. Loss is indepen­dent of length and dispersion is neg­ligible for distances of interest. The optical channel itself has low crosstalk because photons do not interact, al­though stray light from scattering, re­flection, or diffraction onto the de­tector can be an issue. The channel has high bandwidth, and with free-space interconnections, no physical contact is needed. The disadvantages include the cost of the sources and detectors and loss of signal due to limited opto-electronic conversion efficiency, and coupling and align­ment problems.

Additional flexibility becomes ap­parent when considering multiple channels. With free-space channels, beams can cross with no interference and several channels can coexist in the same space. With guided-wave optics, waveguides can cross at an angle with little penalty. Holographic optical

channels promise a very high number and density ofinterconnections. Two-dimensional imaging techniques can be applied for arrays of highly parallel interconnections. In addition, many researchers are studying techniques for reconfigurable optical intercon­nections in which the computer archi­tecture can be optimized for specific problems. 2 , 1 0 Other researchers are studying computer architectures in which an optical implementation would be far easier to construct than its electrical counterpart.

The optical channel has received most of the attention from research­ers, but the entire system and the system performance must be consid­ered. Conversion or modulation effi­ciency of the optical source, coupling and alignment, overall link efficiency, and interfaces to the digital world must be considered. The eventual cost of each link will be very important because, conservatively estimated, a single computer may have 100-100,000 links. Packaging and reliabil­ity of a system will also be crucial to the acceptance of optical interconnec­tions, as will the need for bit error rates ranging from 10-12 to 1 0 - 1 5 .

Experimental optical interconnection systems

Several experimental optical inter­connections between digital circuits have been demonstrated. For exam­ple, fiber-optic interconnections have been installed in a multiprocessor computer by Lane and coworkers 1 1 in a collaborative effort between Honey­well and Thinking Machines. Using only commercially available transmit­ters and receivers, as well as proven GaAs digital multiplexers and demul­

tiplexers, these researchers replaced 1024 wires by two optical fibers to form one leg of a hypercube architec­ture. Successful computer operation was demonstrated with the optical links operating at400 M b / s , and further work toward 1 G b / s is in progress.

Components for optical links in computers have been developed and demonstrated at I B M . 1 2 A four-chan­nel GaAs laser driver was connected to a four-element graded- index sep­arate-confinement heterostructure single quantum well laser array with hybrid integration techniques. Four optical fibers were aligned in an array of V grooves and polished to reflect the light onto four interdigitated Schottky photodiodes. The pho­todiodes were monolithically inte­grated with a M E S F E T receiver. The high level of integration (2000 devices in the receiver) will be important to the further development of many parallel channels.

Presently, electrical interconnec-

FIGURE 2. Illustration of free-space optical inter­connections between boards, showing signal transmission by laser beams in direct paths.

24 OPTICS & PHOTONICS NEWS • OCTOBER 1990

tions are arranged in two-dimensional planes, in chips, on boards, and in backplanes. For board-to-board com­munication, a signal from a chip must be sent to the edge of the board, through a (often slow) backplane, and then from the edge of the next board to the desired position on the board. With an optical interconnection, the signal can be transmitted over a light beam from one board to the next, as shown in Fig. 2, without physical contact between boards. Intercon­nections can be located anywhere on the board, not just at one or two edges, which could add a third dimen­sion to the layout of logic systems.

Currently, a 1-Gb/s fiber-optic telecommunication transmitter mod­ule and its corresponding receiver are expensive, each costing several thou­sand dollars. However, the intercon­nection problem is different from the long-distance telephone problem. The short distance and the small attenuation, combined with very effi­cient lasers and detectors, can mini­mize the driver and receiver circuitry required for the interconnection and thereby reduce its cost.

High-speed optical fiber communi­cation systems designed for long-dis­tance transmission typically have an overall electrical differential current efficiency, differential current out of the detector divided by differential current into the laser when both are biased at the operating point, of 1 to 3% for short links. Attenuation in a long fiber will reduce the efficiency even further. Consequently, when components designed for long-dist­ance fiber systems are employed in a short-distance optical interconnection, only 1-3% of the current into the laser

is delivered by the detector. Since all interconnections must deliver the same standardized voltage level for compatibility with conventional digi­tal logic, the detected signal requires amplification to recover the logic level.

There are three fundamental contri­butions to the overall efficiency. For a laser biased at or above threshold, the differential current efficiency of a link is proportional to the product of the laser differential quantum efficiency, the detector quantum efficiency, and

the optical system efficiency, where the optical system efficiency is defined as the fraction of the photons emitted by the laser that are incident on the detector. Because the electrical power out of the detector is proportional to the square of the detector current, the electrical output power varies as the square of the quantum efficiencies. Thus, efficient devices are very impor­tant. The need to maintain high optical system efficiency dictates the use of lenses or holograms for free-space interconnections to ensure that a large fraction of the emitted photons arrives on the desired detectors. With an efficient optical interconnection, much of the interface circuitry re­quired by inefficient systems (laser driver, preamplifier, automatic gain control, timing recovery, etc.) can be eliminated. Costs, power consump­tion, and the need for high-gain amplifiers—best avoided in a digital environment with nearby logic gates switching on and off—are reduced.

FIGURE 3. Optical interconnection with high-efficiency optoelectronic components. (a) Diagram showing intercon­nections made directly to digital circuits. (b) Upper trace shows the output of the D-type flip-flop as well as the zero level when the optical beam is blocked, and lower trace shows the desired digital pattern, both with the system operating at a clock rate of 1 Gb/s.

OPTICS & PHOTONICS NEWS • OCTOBER 1990 25

Our effort 1 3 has been aimed at implementing an optical interconnec­tion between digital gates without the use of a complicated fiber-optic type transmitter and receiver. Shown in Fig, 3(a) is an optical interconnection that was made by connecting a diode laser directly to one digital circuit on one board, and a detector directly to another circuit on another board. The voltage levels of the GaAs integrated circuits were compatible with emitter-coupled logic (ECL) . A diode laser was chosen over a light-emitting di­ode because of its inherent higher efficiency, its narrower emission cone, and its higher speed. A 1.3-mm Ga­InAsP laser 1 4 was connected to a GaAs code generator that produces a fixed word sequence. The laser threshold is 5 mA and the external differential quantum efficiency is 35% per facet. The low threshold allows the inter­connection to be driven directly by logic gates, while the high differential efficiency allows the detector to gen­erate signal levels sufficiently high to be connected directly to another dig­ital gate without the need for pre­amplification.

The laser light is collected and collimated by an inexpensive aspheric compact-disc lens of 0.55 numerical aperture and 5 mm diameter. The light travels 24 cm, many times the distance required for typical board-to-board applications, to an identical lens that focuses the light onto a commer­cially available 100-µm-diameter Ga­InAs p-i-n photodiode with a nominal 70% quantum efficiency. The output of the photodiode is connected di­rectly to the input of another GaAs digital circuit—a D-type flip-flop-which, as shown in Fig. 3(a), has an

internal amplifier or comparator. In another experiment with a laser with a differential efficiency of 30% per facet, an overall differential electrical current efficiency of 12.5% was mea­sured. This is much better than the 1-3% efficiency common in high­

speed fiber-optic systems. Even higher differential current efficiencies of 19% were measured with a laser having a differential efficiency of 61% per facet.

The D-type flip-flop regenerates the digital output at rates from 100 M b / s up to 1 G b / s [Fig. 3(b)], the limit of the GaAs code generator. N o external laser drivers or preamplifiers are required. This demonstrates that with efficient opto-electronic compo­nents, digital circuits can be intercon­nected simply and directly. The inter­connection in Fig. 3(a) has the poten­tial for a fanout of two, since the light from the back facet could also be collected and sent to another gate. The power consumption of the final stage in the GaAs digital circuit and the diode laser is estimated to be 2 0 -25 mW for a bit pattern with an equal probability of ones and zeros. The resistor network interface between the photodiode and the gate input serves to supply any input bias current to the gate as well as to provide an interface to the photodiode, but is probably not optimal and dissipates about 20 mW with 2 mA of photocurrent.

In our present setup, the transmit­ter and receiver boards are positioned with micropositioner stages, and both the lenses and stages are mounted on an optical table. This is not practical for electronic systems. Eventually, we want to incorporate the lens within a module and eliminate the optical translation stages. Then, the question will be whether the dependence of the performance on the mechanical align­ment tolerances will prove practical.

Prospective performance The evolution of optical intercon-

FIGURE 4. Diode laser with high-reflectivity coating. (a) Improvement of single-ended external differential quantum efficiency after coating. (b) -3 dB frequency as a function of drive level. The actual drive current is indicated at each point.

26 OPTICS & PHOTONICS NEWS • OCTOBER 1990

nections will be shaped by future advances in many component techno­logies. These interconnection-related technologies include: monolithic opto-electronic integrated circuits, hybrid integration, uniform laser ar­rays, surface-emitting lasers, micro-optics, fiber coupling techniques, ho­lograms, laser amplifiers, modulators, spatial light modulators, and heteroepitaxy of GaAs on Si. We will try to consider some of the prospects 1 5

for optical interconnections. As already stated, efficiency is a key

issue in an optical link. Calculations for 1.3 µm buried-heterostructure lasers show that a high-reflectivity coating on one facet can result in laser differential efficiencies of over 60% with no increase in the laser threshold. Quantum-well structures and other material systems can have even higher efficiencies. Recent experimental re­sults with high-reflectivity coatings on a 1.3 µm laser [Fig. 4(a)] have dem­onstrated the predicted efficiency. The threshold actually drops by a factor of two relative to the threshold for the uncoated laser to 3.8 mA. The output power is 10 mW at only 25 mA of laser current. With the use of such efficient lasers combined with anti-reflection coatings on the optics and detector, overall differential current efficiencies of 56% or better may be possible in the future .

Most high-speed diode lasers re­quire current levels of 100 mA or higher to achieve speeds of the order of 10 G H z . A t these current levels, the efficiency and the reliability of the lasers suffer. In addition, this drive current cannot be obtained with the current available from a standard dig­ital circuit (e.g., commercial GaAs

circuits which supply 60 mA of drive current). For the laser in F ig . 4(a), the -3 dB frequency is almost 10 G H z at 20 mA current levels, as shown in Fig. 4(b).

The power dissipation of electrical interconnections in C M O S circuits has been considered by several groups and compared to that of optical inter­connections. For clock rates of 1 G H z or faster and 3 µm C M O S design rules, the somewhat optimistic calcu­lations by Feldman et al.5 predict that diode laser-based optical interconnec­tions can dissipate less power than electrical wires at distances of 1 mm or longer based on consideration of the energy required to charge the capacitance of an integrated-circuit trace (interconnection line) and gate. A similar calculation for external mod­ulators indicates that optical intercon­nections will dissipate less power at distances of the order of 1 mm or longer for risetimes of 1 ns or less for a link efficiency of 45%.

Fanout is one of the areas where optical interconnections present intri­guing possibilities. The fanout is in­dependent of the length of the inter­connection for distances of interest. Fanout can be achieved either electri­cally or optically, with each method having its merits. With external mo­dulators and adequate gate output, fanout can be accomplished by the parallel electrical connection of mod­ulators to a single digital gate. O n the other hand, the fanout can also be accomplished through optical split­ters such that one light beam is split into N beams. Very large fanout ratios are possible. However, even for ideal spli t t ing, the optical power drops as 1 / N , the electrical power out of the

detector drops as 1 / N 2 , and the signal level must be restored through amplification.

Electrical splitting can also be ac­complished for diode-laser-based op­tical interconnections by operating a number of diode lasers in series. In fact, two columns of series-connected diode lasers can be operated in parallel to get fanouts of up to 20 with a modest driver i f the lasers have low threshold. For many lasers of the type represented in Fig. 4, speeds of 10 G H z should be possible with only 20-30 mA of drive capability.

The work on both fiber16 and diode laser 1 7 optical amplifiers is very prom­ising for lossless optical splitting. Laser amplifiers with a gain of at least 20 dB could make splitting ratios of 100:1 possible with no loss. Diode laser amplifiers have been demon­strated at 0.8, 1.3, arid 1.5 µm. Most of the work on fiber amplifiers has been at 1.5 µm, and net gains of 25-40 dB have been demonstrated with Er-doped fibers that can be pumped with diode lasers.

In fiber and waveguide systems, the guided-wave structure must be carefully aligned to the lasers, modu­lators, or detectors. These systems may also require many splices or couplers. Alignment in free-space sys­tems may be solved by several ap­proaches, which include machining fixtures to several micrometer toler­ances, and employing active alignment with feedback for certain applications. A n o t h e r approach is to des ign the optical system for reduced sensi­tivity to misalignment, and our method for accomplishing this is briefly explained here. Also, presently under consideration at Delft Univer­

OPTICS & PHOTONICS NEWS • OCTOBER 1990 27

sity 1 8 are alignment invariant designs, in which software determines the pixels that are interconnected for any given alignment.

The use of optical modules contain­ing a source or detector for each interconnection together with pre­aligned optics can greatly simplify the alignment problems. 1 9 The optics of the transmitter module allow the laser beam to expand to a size consistent with circuit board tolerances, and the optics in the receiver module serve to focus the expanded beam onto a detector. A l l critical alignments be­tween the lens and the source or detector are performed during mod­ule assembly. During circuit assembly, the modules need only be aligned to the much less critical (approximately millimeter) tolerances of the expand­ed-beam input/output optics. For low insertion loss in the optical link, the receiver lens must be located in the near field of the transmitter lens, so that most of the signal can be collected. The effects of misalignment have been considered one at a time19

using simplified thin-lens approxima­tions of the optics and a uniform distribution of light. For a board-to-board interconnection with a trans­mitter graded-index (GRIN) lens of 1 mm diameter and a receiver G R I N lens of 2 mm diameter, 0,7-mm lateral misalignment or 2-degree mis­alignment of either the transmitter or the receiver is consistent with an optical system efficiency of 80% and rise times of 50 ps for a board separa­tion of 3 cm.

Very high densities of interconnec­tions are possible with optical inter­connections. Two-dimensional planes of interconnections can be easily im­

aged with optics. Over a million vertical cavity surface-emitting lasers per square centimeter have been fab­ricated by Jewell et al.20 although operation at this density has not been demonstrated. Improvements in ther­mal dissipation and in the series resis­tance of such devices can lead to a wide range of applications, including inter­connections.

The role of optics The areas where optical and electri­

cal interconnections should be impor­tant are shown in Fig. 5. A t low speeds and short distances (lower left region of the figure), simple wires will pre­dominate because they are cheap and effective. Possible reasons to consider optics in this region are to overcome the difficulties associated with physi­cal layouts, crosstalk, or when imple­menting high-density or high-con­nectivity interconnections. As fre­quencies and distances increase, cross­talk is more of a problem and the reactance (capacitive or inductive) of a wire begins to cause signal distor­tion. The line corresponding to an inductive impedance of 5 Ω for a 25 mil diameter wire is a rough indication of where other interconnection tech­niques may be considered. This in­ductance is that of a 50 µm-diameter wire. Depending on the impedances and current-handling capabilities of the digital circuits and the geometry, size, and resistance of the wire (e.g., C M O S versus E C L circuits, flat versus round wires, etc.), wires may have limitations to the lower left of the 5-W line, or remain useful to the upper right of the line. The physical limit on the use of a wire, even a superconduct­ing wire, occurs when the length of

the interconnection approaches the wavelength of the signal being trans­mitted. A n approximate limit is shown in Fig. 5 by the solid line that indicates a λ/10-long interconnection.

When unacceptable distortion or crosstalk results from the use of simple wires, alternative interconnection technologies, electro-magnetic trans­mission lines, and optics, must be employed. Transmission lines include twisted pair, microstrip, stripline, and coax. Transmission lines become dis­persive and suffer losses from the skin effect as the speed increases. The approximate limit for the use of trans­mission-line techniques is shown in Fig. 5 by the line for a 3-mm-diameter 77-coax copper airline at room tem­perature. The sloped portion of the line corresponds to a 3 dB propaga­tion loss while the horizontal portion

FIGURE 5. Regions of inter­est for electrical and optical interconnections based on both speed and distance.

28 OPTICS & PHOTONICS NEWS • OCTOBER 1990

is at the frequency where higher-order modes become important. Supercon­ductors may extend these limits.

The region where optical intercon­nection techniques begin to promise advantages in terms of power, isola­tion, and signal distortion begins roughly at the boundary where wires begin to have difficulties. There exists a large and currently important range of speed and interconnection distance over which optical and transmission-line techniques are of interest, and the unique features of each may be used depending on the application. A t high speeds and long distances, optical interconnections offer the best solu­tion because of the bandwidth and low loss of both optical fibers and free space. The high bandwidth of present optical channels allows signal trans­mission above 1 0 1 1 H z , although appropriate sources and receivers have yet to be developed.

In summary, various types of optical interconnections have been demon­strated in the laboratory and in com­mercial computers. Optical inter­connections have prospects for even better performance with improved concepts, devices, and systems and should facilitate the development of advanced computers. The challenges ahead are to develop the potential of optical interconnections in the labo­ratory, to transfer the technology into the computer environment, and to make the technology cost effective.

Acknowledgments The author would like to thank R.

C . Williamson and D. L. Spears for helpful discussions and M . J . Corco­ran for technical assistance. This work was sponsored by the Defense A d ­

vanced Research Projects Agency.

R E F E R E N C E S 1. W. T. Cathey and B.J . Smith, High concurrency

data bus using arrays of optical emitters and detectors, Appl. Opt. 18, 1979. p.1687.

2. J. W. Goodman, F. J. Leonberger, S.-Y. Kung, and R. A. Athale, Optical interconnections for VLSI systems, Proc. IEEE 72, 1984. p. 850.

3. R. K. Kostuk, J. W. Goodman, and L. Hesselink, Optical imaging applied to microelectronic chip-to-chip interconnection, Appl. Opt. 24, 1985. p. 2851.

4. W. H. Wu, L. A. Bergman, A. R. Johnston, C. C. Guest, S. C. Esener, P. K. L. Yu, M. R. Feldman, and S. H. Lee, Implementation of optical inter­connections for VLSI, IEEE Trans. Electron Devices ED-34, 1987. p. 706.

5. M. R. Feldman, S. C. Esener, C. C. Guest, and S. H. Lee, Comparison between optical and electrical interconnects based on power and speed considerations, Appl. Opt. 27, 1988. p. 1742.

6. This point has also been made by Alan Huang and Ron Reedy.

7. A. L. Lentine, H. S. Hinton, D. A. B. Miller, J. E. Henry, J. E. Cunningham, L. M. F. Chirovsky, Symmetric self-electro-optic effect device: optical set-reset latch, Appl. Phys. Lett. 52, 1988. p. 1419.

8. C. T. Sullivan, Optical waveguide circuits for printed wire-board interconnections, Proc. SPIE 994, 1988. p. 92.

9. D. H . Hartman, G. R. Lalk, J. W. Howse, and R. R. Krchnavek, Radiant cured polymer optical waveguides on printed circuit boards for photonic interconnection use, Appl. Opt. 28, 1989. p. 40.

10. See for example P. Yeh, A. E. T. Chiou, J. Hong, Optical interconnection using photorefractive dynamic holograms, Appl. Opt. 27, 1988. p. 2093.

11. T. A. Lane, J. A. Quam, B. O. Kahle, and E. C. Parish, Gigabit optical interconnects for the Connection Machine, Proc. SPIE 1178, 1989. p. 24.

12. J. D. Crow, Fiber-optic modules for high speed computer networks, 1989 Proc. 39th Electronic Components Conference, IEEE, New York, 1989. p. 355.

13. D. Z. Tsang, One-gigabit per second free-space optical interconnection, Appl. Opt. 29, 1990. p. 2034.

14. D. Z. Tsang and Z. L. Liau, Sinusoidal and digital high speed modulation of p-type substrate mass-transported diode lasers, J. Lightwave Technol. LT-5, 1987. p. 300.

15. The interested reader is referred to Applied Optics 29 (1990) for recent feature issues on optical interconnections (pp. 1067-1161) and optical computing (pp. 1999-2187).

16. P. Urquhart, Review of rare earth doped fibre

lasers and amplifiers, IEEE Proc. 125, Pt. I,1988. p. 385.

17. T. Saitoh and T. Mukai, Recent progress in semiconductor laser amplifiers, J. Lightwave Technol. 6, 1988. p. 1656.

18. E. E. Freitman, private communication. 19. D. Z. Tsang, Alignment and performance trade­

offs for free-space optical interconnections, 1989 Technical Digest Series, vol. 9 (Optical Society of America, Washington, D.C., 1989), p. 146.

20. J. L. Jewell, Y. H. Lee, A. Scherer, S. L. McCall, N. A. Olsson, J. P. Harbison, and L. T. Florez, Surface-emitting microlasers for photonic switch -ing and interchip connections, Opt. Eng. 29, 1990. p. 210.

D E A N T S A N G is a staff member in the Applied Physics Group at M I T ' s L in ­coln Laboratory.

OPTICS & PHOTONICS NEWS • OCTOBER 1990 29