Reverse engineering the 59-pound printer onboard the Space Shuttle

The Space Shuttle contained a bulky printer so the astronauts could receive procedures, mission plans, weather reports, crew activity plans, and other documents. Needed for the first Shuttle launch in 1981, this printer was designed in just 7 months, built around an Army communications terminal. Unlike modern printers, the Shuttle's printer contains a spinning metal drum with raised characters, allowing it to rapidly print a line at a time.

The Space Shuttle's Interim Teleprinter. The horizontal rails allowed it to be mounted in a Space Shuttle stowage locker.

The Space Shuttle's Interim Teleprinter. The horizontal rails allowed it to be mounted in a Space Shuttle stowage locker. Click this image (or any other) for a larger version.

This printer is known as the Space Shuttle Interim Teleprinter System.1 As the name "Interim" suggests, this printer was intended as a stop-gap measure, operating for a few flights until a better printer was operational. However, the teleprinter proved to be more reliable than its replacement, so it remained in use as a backup for over 50 flights, often printing thousands of lines per flight. This didn't come cheap: with a Shuttle flight costing $27,000 per pound, putting the 59-pound teleprinter in space cost over $1.5 million per flight.

Pilot Overmyer reading a printout from the teleprinter, STS-5, November 16, 1982. From National Archives. The description says that this output is from the Text and Graphics System, but the yellow paper and the date show that this is the Interim Teleprinter.

Pilot Overmyer reading a printout from the teleprinter, STS-5, November 16, 1982. From National Archives. The description says that this output is from the Text and Graphics System, but the yellow paper and the date show that this is the Interim Teleprinter.

We obtained access to a Shuttle teleprinter (probably a development system that remained on the ground) and wanted to put it into operation. I had to reverse engineer three of the boards inside the printer to determine the data format the printer accepted: serial data encoded into audio. But after analyzing the printer and performing a lot of maintenance, we succeeded in getting the printer to print. In this article, I'll describe the Shuttle's Interim Teleprinter, explain its circuitry and drum-based printing mechanism, and show it in operation.

History of the Shuttle's Interim Teleprinter

The motivation for the teleprinter goes back to the Apollo program. During Apollo missions, the only way to send information to the astronauts was by talking to them over the radio and having the astronauts write down the data. NASA decided that the Space Shuttle should include a mechanism to send text and images to the astronauts, a 78-pound, high-tech fax machine called the Uplink Text & Graphics System (TAGS). A high-resolution grayscale image was sent to the Shuttle as a digital data stream. Onboard the Shuttle, a squat CRT displayed the image one line at a time and a fiber-optic faceplate transferred each line to light-sensitive silver emulsion paper. The paper was developed by passing it over a hot roller at 260ºF for 25 seconds, creating a permanent image.

The one flaw in this plan was that sending the digital image to the Shuttle required the Tracking and Data Relay Satellite System (TDRS), which due to delays wouldn't be ready until the sixth Shuttle flight. (The TDRS was a space-based replacement for the worldwide network of ground stations that was used during Apollo.) As a result, NASA decided just seven months before the first Shuttle launch that they needed an interim system "for transmission of real-time, flight-plan changes and other operational data to the crew."2

The Shuttle teleprinter is the result of this rushed effort to create a printer that could work over the existing audio channel rather than the digital TDRS satellite. Due to the time pressure, the Shuttle teleprinter needed to be based on an off-the-shelf printer. Thermal and electrostatic printers were rejected due to toxicity and flammability problems. (The Shuttle teleprinter used a roll of yellowish paper, which required a NASA waiver due to its flammability, a concern ever since the Apollo-1 disaster).

The AN/UGC-74 military communications terminal. This terminal was developed by the Army but also used by the Navy and Air Force. Image from the Operator's Manual, TM 11-5815-602-10.

The AN/UGC-74 military communications terminal. This terminal was developed by the Army but also used by the Navy and Air Force. Image from the Operator's Manual, TM 11-5815-602-10.

The decision was made to use a military communications terminal, the the AN/UGC-743 "Tactical Teletype". The terminal's interfacing was very flexible, supporting serial data in either ASCII or Baudot format, with multiple configurations and baud rates (up to 1200 baud), using either a current-loop or voltage signals. The military terminal supported two-way communication, so it had a keyboard. Remarkably, the terminal also implemented a word processor, controlled by a Motorola 6800 microprocessor (ancestor of the famous MOS 6502). The word processor allowed messages to be composed offline, minimizing the radio transmission time, which was important in a hostile environment. As will be seen, this 100-pound military system required many large changes to be usable on the Space Shuttle, most visibly removing the keyboard.

The printing mechanism

The teleprinter uses a spinning drum with raised characters, shown below.4 To print a character, the printer fires a hammer, forcing the inked ribbon and paper against the raised character on the drum. The drum is 80 characters wide, matching the line length, and there are 80 corresponding hammers, one for each print position. The drum has 64 printable characters, wrapped around each position of the drum.

The printer's drum rotating drum has 64 raised characters in each column. The characters spiral around the drum and are in reverse order, minimizing the chance that a line will fire all the hammers near-simultaneously.

The printer's drum rotating drum has 64 raised characters in each column. The characters spiral around the drum and are in reverse order, minimizing the chance that a line will fire all the hammers near-simultaneously.

The printer prints a line at a time, not instantaneously, but during each revolution of the drum. When the drum makes one complete revolution, each of the 64 characters passes by each print position once. Printing requires precise timing of the hammers to strike the right character on the drum as it whizzes by. The printer control circuitry triggers each hammer at the proper time, when the desired character on the drum is lined up with the hammer, producing the desired text.5

The character set is slightly different between the military printer and the Shuttle printer. The military drum had 64 ASCII characters (upper-case letters only, numbers, and special characters). The drum doesn't contain an explicit space character, since nothing is printed for a space. In its place, the drum has a diamond "◊", used as a special character to indicate a parity error or other error. The drum for the Shuttle teleprinter replaces 10 ASCII special characters with symbols that are more useful to the Shuttle, such as Greek letters for angles. Specifically, the characters ;@[\]^!"#$ are replaced by θ✓‾↑↓~αβΔϕ.

With the teleprinter disassembled, the 20 hammer cards are visible at the front. Two hammer driver cards are to the right of the hammer cards.

With the teleprinter disassembled, the 20 hammer cards are visible at the front. Two hammer driver cards are to the right of the hammer cards.

The video below shows a closeup of the hammers as they strike the paper to print text. The text is the teleprinter's built-in test message: "THE LAZY YELLOW DOG WAS CAUGHT BY THE SLOW RED FOX AS HE LAY SLEEPING IN THE SUN". This test message is based on the traditional quick brown fox..., which is a pangram, containing all 26 letters, but the teleprinter's test sentence is missing J, K, M, Q, and V. However, the test message is exactly 80 characters long and replaces spaces with the diamond "◊", so it is effective for verifying that all 80 columns work.

The electronics

The photo below shows the circuitry inside the teleprinter, looking down from above. At the left are the three interface boards, custom boards that demodulate the incoming audio signal. In front of the interface boards are large inductors to filter the incoming power. Hidden beneath them, a solid-state relay controls the power to the rest of the printer, implementing the low-power standby mode. In the middle, the blue board is the surprisingly complex switching power supply, mounted on a thick metal plate for cooling. Normally, the large roll of paper is mounted above the power supply board. At the right, four large circuit boards implement the main logic of the printer: a printer driver board, a communications board, a memory board, and the processor board. The rotating drum is protected by the perforated black metal grill at the front.

Inside the Shuttle teleprinter, showing the electronics.

Inside the Shuttle teleprinter, showing the electronics.

The demodulator boards

The original military teleprinter received data as a serial bitstream. However, on the Space Shuttle, data was encoded as frequencies on the audio link. Three custom boards were constructed to demodulate the audio data so the rest of the printer could handle it. These boards also performed Shuttle-specific tasks such as powering up the printer when a message comes in, and then returning the printer to standby mode. I reverse-engineered these boards to determine how they work and to determine the data encoding. (Schematics are in the footnotes.7) In this section, I'll discuss these three boards, which are on the left side of the printer.

To summarize, the serial bitstream is encoded with Frequency Shift Keying, with a 1 represented by 3600 Hz and a 0 represented by 7200 Hz.6 The serial data is transmitted at 600 baud, even parity, one stop bit. The demodulation process first converts the input audio to a digital signal by thresholding it. (That is, the input sine wave is converted to a square wave.) The digital signal is autocorrelated to distinguish the 3600 Hz and 7200 Hz signals, recovering the underlying serial data. This signal is passed to the printer's logic boards (part of the original military teleprinter), which convert the serial signal to ASCII bytes and prints them.

Signal processing starts with the "FSK input" board, shown below. First, it amplifies the input audio signal. (The two large resistors provide a 600 Ω load for the audio input.) Next, a 900 Hz high-pass filter eliminates low-frequency noise. (The filter is implemented by a two-stage Sallen-Key topology.)

The input board.

The input board.

The signal bounces from board to board, going to the "output FSK demod" board next. This board has a carrier-detect circuit that turns on the rest of the printer if it detects an input signal. This allows the printer to sit idle until it receives a signal from Earth. This board also applies the threshold to the signal to turn it into a digital waveform, which goes to the "control" board.

The output board.

The output board.

The output board also holds the 5-volt and 12-volt linear regulators that power the three boards; these are the metal-can ICs at the bottom of the board. To reduce the load on the regulators, two large resistors drop the input voltage (28 volts) to a lower level before it is regulated.

The control board holds the FSK decoder, an interesting circuit that converts the two FSK frequencies to binary by implementing a digital auto-correlator. It uses a 64-bit shift register to delay the digital input by 139 µs. The input and the delayed input are XOR'd together, generating a result that depends on the frequency. A 7200 Hz signal repeats every 139 µs, so the input and the delayed input match, yielding 0 from the XOR. However, a 3600 Hz square wave switches state every 139 µs, so the two XOR inputs will always differ, resulting in a 1 output. Thus, the circuit cleanly distinguishes between a 3600 Hz input and a 7200 Hz input.

The control board.

The control board.

The digital demodulator avoids some of the problems of an analog FSK demodulator. It is not sensitive to signal levels, since the signal is converted to digital. The digital demodulator is also not sensitive to harmonics, which can cause problems with analog demodulators. Finally, it doesn't require the carefully-tuned filters of an analog circuit.

The demodulated signal passes from the control board back to the output board. This board applies a 400 Hz low-pass filter and then a threshold to convert the signal back to binary. If the input frequencies are not exact, the demodulator will produce the correct 0 or 1 value over most of the waveform, but there will be glitches at the edges. The low-pass filter removes these glitches. (You might be concerned that a 600-baud signal would be wiped out by a 400 Hz low-pass filter. However, the worst case signal (alternating 0's and 1's) would be 300 Hz because it takes two bits to make one cycle, so the filter has plenty of margin.) Next, the board blocks the signal unless a carrier is detected. This ensures that random noise isn't demodulated and printed. Finally, the serial binary signal leaves the custom Shuttle boards and goes to the teleprinter's communication board, part of the standard teleprinter.

I noticed two unusual things about these boards. First, they have some modifications: "bodge" wires and added components. Second, the boards are not conformal coated, which is unusual for aerospace boards. (The four logic cards, in comparison, are protected with conformal coating.) My hypothesis is that these boards were development boards, early in the design process of the Shuttle teleprinter, so they were modified as the design changed. The teleprinter is also marked "Not for flight", which supports this theory.

Mission Specialist Thagard getting output from the teleprinter. Flight STS-7, June 24, 1983. From NARA. Although the description says this is the Text & Graphics System, it is clearly the Interim Teleprinter.

Mission Specialist Thagard getting output from the teleprinter. Flight STS-7, June 24, 1983. From NARA. Although the description says this is the Text & Graphics System, it is clearly the Interim Teleprinter.

The logic cards

The military teleprinter contained four logic circuit cards: a CPU card, a memory card, a communications card, and a print control card, mounted at the right rear of the teleprinter. These cards are used unchanged in the Shuttle teleprinter.

The circuitry is more complex than you might expect, with four large cards full of ICs. There are several reasons for this. First, the cards use 1970s microprocessor technology, so it takes a lot of circuitry to do anything. In particular, many simple 7400-series logic chips perform "glue" functions: decoding addresses, buffering data, latching signals, and so forth. Moreover, a drum printer is inherently complicated, since 80 hammers must be driven at the right time based on the desired characters. Third, the teleprinter is very flexible, supporting multiple signal levels and two character formats (ASCII and Baudot). Most surprisingly, the teleprinter implements a word processor, allowing messages to be composed and edited offline. Of course, since the Shuttle's teleprinter is only used to receive data, and doesn't even have a keyboard, the word processor feature is entirely useless.

The CPU card

The CPU card holds the microprocessor that controls the teleprinter. Its most important function is to convert a line of ASCII characters into print drum codes. These codes are stored in memory for use by the print control card. The CPU also implements configuration and self-test functions.

The diagram below shows some of the main components. The CPU card contains a Motorola 6800 CPU, 4 kilobytes of memory, and a ROM that holds its program code.8 Inconveniently, all the IC part numbers are military numbers so it takes some investigation to determine what a part really is. The MC6822 is a Peripheral Interface Adapter, a Motorola chip that provides two parallel I/O ports. This chip is used on three of the cards to support a variety of I/O tasks. On the CPU card, the I/O ports drive eight status lamps (most of which were removed for the Shuttle teleprinter) as well as internal status signals such as "paper low" or "keyboard present" and the baud rate setting input.

The CPU card is centered around a Motorola 6800 microprocessor.

The CPU card is centered around a Motorola 6800 microprocessor.

The print control card

In a sense, the print control card is the heart of the printer, since it causes characters to be printed by firing hammers against the rotating drum. As the drum goes through one revolution, all 64 characters will spin past each of the 80 print positions. By firing hammers at the exact time, the card prints a line of text.9 In more detail, for each row on the drum, the printer card scans through the 80-character memory buffer using Direct Memory Access (DMA). If the value in memory matches the current drum row number, the hammer is fired. Note that the hammers don't fire simultaneously, but in sequence as memory is scanned.

This diagram shows how the print control board interacts with the rest of the system. From the Maintenance manual, TM 11-5815-602-24.

This diagram shows how the print control board interacts with the rest of the system. From the Maintenance manual, TM 11-5815-602-24.

The diagram above shows the interaction between the drum, the print control card, and the 80 hammers. The hammers are implemented on 20 print hammer cards, each with 4 hammers. Electrically, the hammers are arranged in a matrix. One wire out of 20 (S1-S20) selects the hammer board, the group of four. Another wire selects one of four hammers (Col 1-4). This approach simplifies the electronics, since 20 + 4 driver circuits and wires are used, rather than 80 (one for each column). The print control card is synchronized to the drum by two photo-transistor sensors that detect the drum's position. One sensor is triggered on each row, while the other sensor triggers once per revolution.

The print control card is shown below, with the main functional blocks labeled. The large purple-and-gold chip is the PIA, the same I/O chip that appeared on the CPU card. It handles a variety of signals such as the self-test request, paper out, and the drum stop signal. The mode control logic generates timing signals depending on the printer's mode. The data compare logic increments the row counter on each drum pulse, and compares the row counter to the value read from memory.10 The hammer driver circuitry on the left selects one of the 20 hammer cards, while the hammer driver circuitry on the right selects one of four hammers. The ribbon circuitry raises and lowers the ribbon so the ribbon doesn't block the text when the printer is idle. The line feed circuitry advances the paper for a line feed operation.

The print control card prints data by driving the hammers.

The print control card prints data by driving the hammers.

The photo below shows one of the hammer cards, with four hammers. Each hammer has an electromagnet that pulls a lever, rotating the hammer wheel, and causing the hammer to strike the paper. (The hammers themselves are in the upper right of the photo.) A screw adjustment controls the distance between each hammer and the paper, allowing precise adjustment of the timing. (Marc had to carefully adjust all the hammers to make the print quality readable.)

One of the 20 Hammer driver cards. Photo courtesy of Marcel.

One of the 20 Hammer driver cards. Photo courtesy of Marcel.

The communication card

The communication card handles the teleprinter's serial data input. The key chip is the 8251A, a USART (Universal Synchronous/Asynchronous Receiver/Transmitter). This complex chip performs the conversion between the serial data stream and the bytes that the processor uses. (Note that the military teleprinter both sent and received serial data, while the Shuttle teleprinter only receives data.) The chip has a few support chips, labeled "UART" in the diagram below. The board has another Peripheral Interface Adapter chip, providing two I/O ports. These ports have functions such as reading the serial line settings (ASCII vs. Baudot, odd or even parity, number of stop bits, and current loop levels).

The communication card converts the serial input to parallel byte data.

The communication card converts the serial input to parallel byte data.

The board also has circuitry to generate the clock pulses for the selected baud rate. The mode circuitry handles various phases of transmit/receive. The filter/demod circuitry handles different input types, digitally filtering and demodulating as necessary.11

The memory card

The memory card supports the word-processing feature. It provides additional RAM to hold the text buffer as well as the ROM holding the software for editing. The 16 DRAM chips on the left (MK4027) provide 8 KB of RAM while the two ROM chips on the right provide 8K of ROM. The chips in the middle to the right of the resistors split the 12 address bits into row and column addresses as required by the RAM chips. The address signals go through the numerous 24 Ω resistors in the middle; I don't know why. According to the manual, the printer operates fine without this card, except without the word processor. Since the word processor was irrelevant to the Shuttle, I wonder why this card wasn't removed to reduce weight.

The memory card has additional RAM and ROM to support the word processing feature.

The memory card has additional RAM and ROM to support the word processing feature.

The power supply

The power supply board (shown earlier) implements separate power supplies for different parts of the printer.12 The supplies are implemented as switching power supplies, which were not as common at the time as now. The microprocessor supply provides +5V, +12V, and -5V, voltages required by memory chips in the 1970s. A separate switching power supply provides +5V, -8.6V, and +8.6V for the keyboard, dustcover, and interface module, components that were removed for the Shuttle teleprinter. Another supply powers the printer's status lamps.

The drum motor supply is important because its voltage is regulated to control the rotational speed of the drum. A sensor on the drum provides a feedback pulse for each row on the drum. (I think the drum speed is 868 RPM.) These pulses control the drum motor's switching supply. If the drum spins too slowly, the voltage is increased, and similarly if it spins too fast.

The hammers have an unusual constant-current power supply. When the printer is active, this power supply generates +18 V. However, the power supply is designed to use a constant current of 600 mA regardless of the hammer activity. A capacitor provides a reservoir of power that is filled by the constant current. If the hammers are using less current, the excess current is bled off through a resistor. The purpose of this is "to mask printing intelligence during periods of message traffic." In other words, if you used a teleprinter in the embassy in Moscow, for instance, spies could monitor power transients to see when hammers are firing, and perhaps figure out what is being printed. By keeping the current constant, this source of intelligence is blocked. Of course, this feature is useless on the Space Shuttle and only wastes power.

The military teleprinter accepted multiple input voltages: 22-30 VDC, 115 VAC, or 230 VAC, along with a 12 VDC battery backup. The transformers and diodes to support these voltages were part of the interface module that was removed for the Shuttle teleprinter. Instead, the Shuttle teleprinter is powered by 28 VDC.

Mechanical changes

The military teleprinter underwent significant mechanical changes to make it suitable for the Shuttle. These changes reduced its weight from 100 pounds to 59 pounds. The most visible change to the printer is the removal of the keyboard. The entire front section of the printer was replaced, removing the controls that were not needed in the Shuttle.13 The rugged frame of the original printer was replaced with a lighter-weight (but still substantial) frame. Horizontal rails were added to the frame to support the printer in the Shuttle locker.

The photo below shows the front of the Shuttle teleprinter. While the military teleprinter had numerous lights and switches on the front, the Shuttle teleprinter has just two lights and four switches.

Front view of the Shuttle teleprinter. The bar across the middle holds a paper cutter for removing the output.

Front view of the Shuttle teleprinter. The bar across the middle holds a paper cutter for removing the output.

NASA was concerned that the temperature of the teleprinter could become hazardous to the astronauts. To mitigate this danger, the teleprinter had a large heat-sensitive warning sticker. The yellow sticker on the left of the teleprinter changes color and displays an image if it heats up: it shows a bandaged hand and the word "HOT". Above it is an "Omegalabel" temperature monitoring sticker that shows the highest temperature the device reached. There are more of these stickers inside the teleprinter on various motors.

The Interim Teleprinter inside the Space Shuttle

The teleprinter was too large to be mounted on the flight deck, so it was mounted in a storage locker on the middeck, one level lower. The photo below shows the location of the locker that held the teleprinter (although the teleprinter was not present in this photo), looking backward (aft) toward the airlock. The locker is denoted MA9F, indicating Mid-deck Aft, position 9F (details), in the back on the right side of the Shuttle.

This photo shows the locker that held the teleprinter. Photo by DMolybdenum, panorama viewed on renderstuff.

This photo shows the locker that held the teleprinter. Photo by DMolybdenum, panorama viewed on renderstuff.

The teleprinter was noisy because of its impact printing; even with it in a locker, the sound outside was 69.5 dB. The solution was to soundproof the locker with acoustic insulation. Various insulating materials were tested until one was found that passed the toxicity requirements. Another flammability waiver was required for the insulation.

Putting the teleprinter in an insulated locker without cooling caused another problem: overheating. The military teleprinter used 34 watts even while idle, which would cause the printer to become dangerously hot after just 6 orbits. The printer was redesigned to support a standby mode that used just 1 watt. When a signal from Earth was detected, the printer would power up while in use, and then return to standby mode. A circuit was added to send a tone back to Earth when the printer was activated, reassuring Mission Control that the printer had switched out of standby mode. These circuits were on the three custom Shuttle boards described earlier.

Putting the teleprinter in a locker made cabling difficult. The solution was a panel on the locker door with connectors for power and audio. The panel has a power switch and light as well as a light to indicate that a message has been received.

The panel on the outside of the locker, used for connection to the teleprinter. From distantsuns, NASA Space Flight forum.

The panel on the outside of the locker, used for connection to the teleprinter. From distantsuns, NASA Space Flight forum.

The photo below shows the teleprinter locker with the connection panel on the far left. Note the cables attached to the connectors. These cables went across the back of the Shuttle to the left side, where they went up to the flight deck; the cable routing was performed before launch.14 For this flight, the neighboring locker MA16F held 3300 honeybees for a student experiment.

The teleprinter in middeck locker MA9F on flight STS-41C.  The hands belong to mission specialist van Hoften.  From National Archives; the description says the photo is from 1995 and shows the Thermal Impulse Printer system, but both are wrong. (STS-41C was in April, 1984.)

The teleprinter in middeck locker MA9F on flight STS-41C. The hands belong to mission specialist van Hoften. From National Archives; the description says the photo is from 1995 and shows the Thermal Impulse Printer system, but both are wrong. (STS-41C was in April, 1984.)

The teleprinter cables connect to the shuttle at panel A15 on the aft bulkhead of the flight deck on the left side of the Shuttle. In other words, if you sat in the Shuttle Commander's seat in the cockpit and turned around, this is what you would see.

The connections for the teleprinter in the flight deck. This photo shows Atlantis in the Kennedy Space Center visitor complex. In use, the Shuttle was much more cluttered.

The connections for the teleprinter in the flight deck. This photo shows Atlantis in the Kennedy Space Center visitor complex. In use, the Shuttle was much more cluttered.

The audio cable from the teleprinter went to the Payload Specialist communication connection on panel A15, while the power cable went to the DC power connection right below. During launch, this audio connection was needed for crew communication, so the teleprinter was plugged in after launch and the audio settings were reconfigured on panel L9. A cue card was placed above panel L9 with instructions on the teleprinter.

The teleprinter's replacements

The Shuttle teleprinter was supposed to be used for a short time until the Uplink Text and Graphics System (TAGS) entered service, but things didn't work out that way. TAGS, described earlier, was the fax-like system that could receive grayscale images, but it depended on the TDRS satellites with their support for digital data. The first TDRS satellite was launched by the sixth shuttle flight, STS-6 (1983). This allowed the use of TAGS on STS-7, but the printer promptly jammed.15 TAGS had constant problems with jamming; on STS-35, the printer jammed and then the unjamming tool broke. Due to the unreliability of the TAGS, the Interim Teleprinter was kept in service as a backup device. TAGS was mounted on a dual cold plate in avionics bay 3 of the crew compartment middeck (details), on the other side of the airlock from the teleprinter.

The Uplink Text and Graphics System, serial number 2. Photo from Smithsonian National Air and Space Museum.

The Uplink Text and Graphics System, serial number 2. Photo from Smithsonian National Air and Space Museum.

After a decade, another printer, the Thermal Impulse Printer System (TIPS) was put into service, probably on flight STS-56 in 1993. Once TIPS proved its reliability, it replaced both the teleprinter and the Text and Graphics System (TAGS). The TIPS printer was installed in mid-deck locker MF28E; the F indicates the locker was on the forward wall, not the aft wall that held the Interim Teleprinter. As a backup for the TIPS, the Shuttle flew with a second TIPS.

The Thermal Impulse Printer System (TIPS) on flight STS-58. From National Archives. The description says that this device is the teleprinter but it is TIPS.

The Thermal Impulse Printer System (TIPS) on flight STS-58. From National Archives. The description says that this device is the teleprinter but it is TIPS.

One motivation behind the TIPS thermal printer was NASA's desire to use more commercial-off-the-shelf (COTS) equipment instead of expensive custom equipment. The TIPS printer is the Raytheon TDU-850 printer (below), a commercial product that sold for $4950. A custom communication interface board inside the printer provided the interface between the printer and the Shuttle's S-Band and Ku-Band communications systems. This interface also allowed astronauts to use the TIPS as a printer for an onboard personal computer.

The Raytheon TDU-850 printer (Thermal Display Unit). From EDN, Mar 17, 1988, p.251.

The Raytheon TDU-850 printer (Thermal Display Unit). From EDN, Mar 17, 1988, p.251.

The photo below shows the TIPS printer in use, printing a long stream of output that Eileen Collins is reading. Collins was the first woman to pilot the Space Shuttle; she flew on the Shuttle four times, twice as pilot and twice as commander.

Pilot Collins reading output from the TIPS printer, the gray box on the right. This is flight STS-84, Atlantis. Photo from National Archives.

Pilot Collins reading output from the TIPS printer, the gray box on the right. This is flight STS-84, Atlantis. Photo from National Archives.

The teleprinter, operational

We succeeded in making the Shuttle teleprinter operational. The printer had many mechanical problems, mainly because the rubber rollers had turned to liquid and gummed up the mechanism. Marc disassembled the printer, carefully cleaned the mechanism, and realigned everything. I won't discuss the restoration process here since there will be a video on CuriousMarc's channel. We were able to send FSK-modulated data to the printer and it was printed successfully, as shown below.

Conclusions

At first, I thought that the Shuttle's Interim Teleprinter was a terrible design. It's absurdly heavy and was in danger of overheating. Although the design started with an existing product, much of it required redesign: the front section, the new drum, the interface, and even the frame. The design inherited features it couldn't use, such as the built-in word processor. And the constant-current feature was pointless for the Shuttle and just wasted power.

When I learned that the design had to be completed in just seven months, my opinion of the teleprinter improved. Moreover, the design had many constraints, such as toxicity and flammability restrictions, that limited the potential approaches.

In the end, the teleprinter was used on over 50 flights, acting as a reliable backup to the somewhat flaky Text and Graphics System (TAGS).16 Despite its name, the Interim Teleprinter turned out to be a long-lasting solution, not interim at all. So I have to conclude that the teleprinter was a good design, working much better and much longer than intended.17

In any case, the Interim Teleprinter is an interesting piece of hardware and I hope you enjoyed this article. Follow me on Mastodon as @[email protected] or RSS. Thanks to Marcel for providing the printer. Restoration performed with CuriousMarc, Eric Schlapefer, and Mike Stewart.

Notes and references

  1. References for the teleprinter:
    The Interim Teleprinter and its development is described in detail in: M.D. Schuette, “Space Shuttle Interim Teleprinter System,” in Conference record: NTC ’82, Systems for the Eighties, IEEE. (I'll call this the "teleprinter paper" for short.)
    The Shuttle Crew Operations Manual has extensive information on the shuttle and some information on the teleprinter.
    The teleprinter is briefly discussed here.
    Some teleprinter information is in the "Crew Systems Equipment Workbook" via RR Auction.
    The layouts of the Shuttle panels are in Orbiter OV-102 Display and Control Panel Configuration.
    The lockers are described in Orbiter middeck/paylod standard interfaces control document.
    The manuals for the AN-UGC/74 are at RadioNerds.
    An enormous collection of Shuttle documents is at gandalfddi

  2. The teleprinter paper mentions that Shuttle had one other option for receiving hardcopy data: the Text Uplink to Mass Memory System (TUMMS). This allowed text to be displayed on a CRT and the crew could take a Polaroid photo. This was obviously an impractical solution. I couldn't find any other references to TUMMS, so TUMMS may be a proposal that wasn't implemented. 

  3. Specifically, the Shuttle teleprinter was based on the Honeywell Model AN/UGC-74A9(V)3 Communications Terminal. 

  4. The mechanism of a drum printer is similar to a chain printer such as the IBM 1403 line printer: each print position has a hammer that fires when the correct character is in that position. However, chain printers have better print quality than drum printers, due to the effect of timing errors. In a drum printer, a small timing error on a hammer will cause the character to be printed too high or too low. In a chain printer, however, a timing error will cause the character to be shifted to the left or right. Vertical mispositioning is obvious and looks terrible. Horizontal mispositioning is much less noticeable since character spacing is normally slightly variable. 

  5. To be precise, the hammer is fired 1.5 characters early due to its travel time. By the time the hammer hits the drum, the drum has rotated enough to put the desired character in place. Each hammer has a screw to adjust its distance to the drum, necessary to get the timing exact. It's amazing that this system works and doesn't produce a smudged mess. 

  6. After reverse-engineering the boards, I found a paper on the Shuttle teleprinter that specified the FSK frequencies as 1600 Hz for a 0 and 2057 Hz for a 1, different from what we used. Perhaps the frequencies were changed during development. 

  7. I created schematics of the three Shuttle-specific boards. Click an image for a larger (readable) version.

    Schematic of the input board.

    Schematic of the input board.

    Schematic of the control board.

    Schematic of the control board.

    Schematic of the output board.

    Schematic of the output board.

     

  8. The block diagram below shows the main functional blocks of the CPU card.

    CPU block diagram. From Maintenance Manual, TM 11-5815-602-24, p3-6

    CPU block diagram. From Maintenance Manual, TM 11-5815-602-24, p3-6

     

  9. I expected that a line would be printed during one drum revolution but looking at the print pattern, it appears to take multiple revolutions per line. Perhaps the printer is avoiding hammers firing too close together to minimize current spikes. Moreover, the published print speed of 60 characters per second is considerably slower than one revolution. Or perhaps the hammer pattern is randomized so spies can't listen in and determine what is being printed. I'm still investigating. 

  10. Looking at the circuitry, I think the memory buffer holds the drum row number for each position, and the print control card fires the hammer if the value matches the current row number. In contrast, the "obvious" approach would put the character values in the memory buffer and the print control card would match against the current drum character. The implemented solution puts less work on the print control card, which only needs to update the target comparison value once per line, rather than every character. However, it requires the CPU card to transform the input characters into row values. 

  11. The teleprinter accepts two types of inputs: NRZ and D10. NRZ (Non-Return to Zero) is the straightforward encoding of the serial signal as 0's or 1's. The manual doesn't define D10, but I think it is Manchester encoding, using a 01 sequence for a 0 and a 10 sequence for a 1 (or inverted). The D10 signal is self-clocking, since each bit contains a transition. The demodulation circuit converts the D10 signal into a straight bit sequence. An NRZ signal can either use an external clock or an internal clock from the baud rate generator. With the internal clock, the input is sampled four times and digitally filtered since the input may not exactly line up with the internal clock. 

  12. The power supply is explained in the Maintenance Manual. The fold-out power supply schematics in that manual were not scanned for some reason but can be found in the B&C Maintenance Manual

  13. The military teleprinter contained a large interface module at the back, providing the signal and power connections to the terminal. The serial-line signals could be a 20-milliamp current loop, a 60-milliamp current loop, or MIL-STD-188/144 (similar to RS-422). The interface module converts these signals to the TTL signals used internally. The interface module also contains a power supply for the interface circuitry. Since this interfacing was not required for the Shuttle, the interface module was discarded and replaced with the Shuttle's custom FSK interface cards. The AC power supply and filtering was also removed. 

  14. I was a bit surprised that the teleprinter cables would run for a long distance through the Shuttle. But the Shuttle is full of wires and cables running in all directions, as shown in the photo below. This photo is from the same angle as the earlier diagram showing where the teleprinter is connected. This flight was after the teleprinter was retired, but the teleprinter would have been plugged in behind the exercise equipment.

    The aft flight deck of Discovery during STS-116. From National Archives.

    The aft flight deck of Discovery during STS-116. From National Archives.

     

  15. One source says that the inaugural flight of TAGS was STS-29 (March 1989). Another source says that testing of the "new" TAGS system continued on STS-29. Contradicting this, TAGS was used on STS-7 (June 1983), jamming after the first page. TAGS was also used on STS-8 (August 1983) but failed after five pages. The TAGS unit was not flown on STS-41B (Feb 1984, the next Challenger flight after STS-8). (Note that STS-41B was the tenth flight, considerably before STS-29, the 28th flight. The Space Shuttle mission numbers are a mess.) It's hard to reconcile these statements. Probably, TAGS was still in the testing stage as late as STS-29 due to reliability problems. 

  16. The teleprinter had a few problems during use. On flight STS-6, the teleprinter got stuck in high power mode. On flight STS-30, messages were illegible (link). 

  17. The teleprinter shows the risk of building an interim solution that turns out to last much longer than expected. This also happened with the Interim Upper Stage (IUS), a launch system to boost Shuttle payloads to a higher orbit. The Interim Upper Stage was designed as a temporary solution until a space tug became available. Eventually, NASA realized that nothing was replacing the IUS, so it was renamed to "Inertial Upper Stage", preserving the acronym.

    I'll mention that this also happened with the 8086 processor. It was intended as an interim processor until the iAPX 432 "micro-mainframe" processor was ready. The iAPX 432 turned out to be a disaster, while the "stopgap" 8086 is still with us as the x86 architecture. 

Inside an IBM/Motorola mainframe controller chip from 1981

In this article, I look inside a chip in the IBM 3274 Control Unit.1 But before I discuss the chip, I need to give some background on mainframes. (I didn't completely analyze the chip, so don't expect a nice narrative or solid conclusions.)

Die photo of the Motorola/IBM SC81150 chip. Click this image (or any other) for a larger version.

Die photo of the Motorola/IBM SC81150 chip. Click this image (or any other) for a larger version.

IBM's vintage mainframes were extremely underpowered compared to modern computers; a System/370 mainframe ran well under 1 million instructions per second, while a modern laptop executes billions of instructions per second. But these mainframes could support rooms full of users, while my 2017 laptop can barely handle one person.2 Mainframes achieved their high capacity by offloading much of the data entry overhead so the mainframe could focus on the "important" work. The mainframe received data directly into memory in bulk over high-speed I/O channels, without needing to handle character-by-character editing. For instance, a typical data entry terminal (a "3270") let the user update fields on the screen without involving the computer. When the user had filled out the screen, pressing the "Enter" key sent the entire data record to the mainframe at once. Thus, the mainframe didn't need to process every keystroke; it only dealt with complete records. (This is also why many modern keyboards have an "Enter" key.)

A room with IBM 3179 Color Display Stations, 1984. Note that these are terminals, not PCs. From 3270 Information Display System Introduction.

A room with IBM 3179 Color Display Stations, 1984. Note that these are terminals, not PCs. From 3270 Information Display System Introduction.

But that was just the beginning of the hierarchy of offloaded processing in a mainframe system. Terminals weren't attached directly to the mainframe. You could wire 16 terminals to a terminal multiplexer (such as the 3299). This would in turn be connected to a 3274 Control Unit that merged the terminal data and handled the network protocols. The Control Unit was connected to the mainframe's channel processor which handled I/O by moving data between memory and peripherals without slowing down the CPU. All these layers allowed the mainframe to focus on the important data processing while the layers underneath dealt with the details.3

An overview of the IBM 3270 Information Display System attachment. The yellow highlights indicate the 3274 Control Unit. From 3270 Information Display System: Introduction.

An overview of the IBM 3270 Information Display System attachment. The yellow highlights indicate the 3274 Control Unit. From 3270 Information Display System: Introduction.

The 3274 Control Unit (highlighted above) is the source of the chip I examined. The purpose of the Control Unit "is to take care of all communication between the host system and your organization's display stations and printers". The diagram above shows how terminals were connected to a mainframe, with the 3274 Control Unit (indicated by arrows) in the middle. The 3274 was an all-purpose box, handling terminals, printers, modems, and encryption (if needed). It could communicate with the mainframe at up to 650,000 characters per second. The control unit below (above) is a boring beige box. The control panel is minimal since people normally didn't interact with the unit. On the back are coaxial connectors for the lines to the terminals, as well as connectors to interface with the computer and other peripherals.

An IBM 3274-41D Control Unit. From bitsavers.

An IBM 3274-41D Control Unit. From bitsavers.

The Keystone II board

In 1983, IBM announced new Control Unit models with twice the speed: these were the Model 41 and Model 61. These units were built around a board called Keystone II, shown below. The board is constructed with IBM's peculiar PCB style. The board is arranged as a grid of squares with the PCB traces too small to see unless you zoom in. Most of the decoupling capacitors are in IBM's thin, rectangular packages, although I see a few capacitors in more standard blue packages. IBM is almost a parallel universe with its unusual packaging for ICs and capacitors as well as the strange circuit board appearance.

The Keystone II board. The box is labeled Keystone II FCS [i.e. First Customer Shipment] July 23, 1982. Photo from bitsavers, originally from Bob Roberts.

The Keystone II board. The box is labeled Keystone II FCS [i.e. First Customer Shipment] July 23, 1982. Photo from bitsavers, originally from Bob Roberts.

Most of the chips on the board are IBM chips packaged in square aluminum cans, known as MST (Monolithic System Technology). The first line on each package is the IBM part number, which is usually undocumented. The empty socket can hold a ROS chip; ROS is Read-Only Store, known as ROM to people outside IBM. The Texas Instruments ICs in the upper right are easier to identify; the 74LS641 chips are octal bus transceivers, presumably connecting this board to the rest of the system. Similarly, the 561 5843 is a 74S240 octal bus driver while the 561 6647 chips are 74LS245 octal bus transceivers.

The memory chips on the left side of this board are interesting: each one consists of two "piggybacked" 16-kilobit DRAM chips. IBM's part number 8279251 corresponds to the Intel 4116 chip, originally made by Mostek. With 18 piggybacked chips, the board holds 64 kilobytes of parity-protected memory.

The photo below shows the Keystone II board mounted in the 3274 Control Unit. The board is in slot E towards the left and the purple Motorola IC is visible.

The Keystone II card in slot E of a 3274-41D Control Unit. Photo from bitsavers.

The Keystone II card in slot E of a 3274-41D Control Unit. Photo from bitsavers.

The Motorola/IBM chip

The board has a Motorola chip in a purple ceramic package; this is the chip that I examined. Popping off the golden lid reveals the silicon die underneath. The package has the part number "SC81150R", indicating a Motorola Special/Custom chip. This part number is also visible on the die, as shown below.

The corner of the die is marked with the SC81150 part number. Bond pads and bond wires are also visible.

The corner of the die is marked with the SC81150 part number. Bond pads and bond wires are also visible.

While the outside of the IC is labeled "Motorola", there are no signs of Motorola internally. Instead, the die is marked "IBM" with the eight-striped logo. My guess is that IBM designed the chip and Motorola manufactured it.

The IBM logo on the die.

The IBM logo on the die.

The diagram below shows the chip with some of the functional blocks identified. Around the outside are the bond pads and the bond wires that are connected to the chip's grid of pins. At the right is the 16×16 block of memory, along with its associated control, byte swap, and output circuitry. The yellowish-white lines are the metal layer on top of the chip that provides the chip's wiring. The thick metal lines distribute power and ground throughout the chip. Unlike modern chips, this chip only has a single metal layer, so power and ground distribution tends to get in the way of useful circuitry.

The die with some functional blocks identified.

The die with some functional blocks identified.

The chip is centered around a 16-bit bus (yellow line) that connects many part of the chip. To write to the bus, a circuit pulls bus lines low. The bus lines are kept high by default by 16 pull-up transistors. This approach was fairly common in the NMOS era. However, performance is limited by the relatively weak pull-up current, making bus lines slow to go high due to R-C delays. For higher performance, some chips would precharge the bus high during one clock cycle and then pull lines low during the next cycle.

The two groups of I/O pins at the bottom are connected to the input buffer on the left and the output buffer on the right. The input buffer includes XOR circuits to compute the parity of each byte. Curiously, only 6 bits of the inputs are connected to the main bus, although other circuits use all 8 bits. The buffer also has a circuit to test for a zero value, but only using 5 of the bits.

I've put red boxes around the numerous PLAs, which can be identified by their grids of transistors. This chip has an unusually large number of PLAs. Eric Schlaepfer hypothesizes that the chip was designed on a prototype circuit board using commercial PAL chips for flexibility, and then they transferred the prototype to silicon, preserving the PLA structure. I didn't see any obvious structure to the PLAs; they all seemed to have wires going all over.

The miscellaneous logic scattered around the chip includes many latches and bus drivers; the latch circuit is similar to the memory cells. I didn't fully reverse-engineer this circuitry but I didn't see anything that looked particularly interesting, such as an ALU or counter. The circuitry near the PLAs could be latches as part of state machines, but I didn't investigate further.

I was hoping to find a recognizable processor inside the package, maybe a Motorola 6809 or 68000 processor. Instead, I found a complicated chip that doesn't appear to be a processor. It has a 16×16 memory block along with about 20 PLAs (Programmable Logic Arrays), a curiously large number. PLAs are commonly used in processors for decoding instructions, since they can match bit patterns. I couldn't find a datapatch in the chip; I expected to see the ALU and registers organized in a large but regular 8-bit or 16-bit block of circuitry. The chip doesn't have any ROM4 so there's no microcode on the chip. For these reasons, I think the chip is not a processor or microcontroller, but a specialized data-handling chip, maybe using the PLAs to interpret bits of a protocol.

The chip is built with NMOS technology, the same as the 6502 and 8086 for instance, rather than CMOS technology that is used in modern chips. I measured the transistor features and the chip appears to be built with a 3.5 µm process (not nm!), which Motorola also used for the 68000 processor (1979).

The memory buffer

The chip has a 16×16 memory buffer, which could be a register file or a FIFO buffer. One interesting feature is that the buffer is triple-ported, so it can handle two reads and one write at the same time. The buffer is implemented as a grid of cells, each storing one bit. Each row corresponds to a 16-bit word, while each column corresponds to one bit in a word. Horizontal control lines (made of polysilicon) select which word gets written or read, while vertical bit lines of metal transmit each bit of the word as it is written or read.

The microscope photo below shows two memory cells. These cells are repeated to create the entire memory buffer. The white vertical lines are metal wiring. The short segments are connections within a cell. The thicker vertical lines are power and ground. The thinner lines are the read and write bit lines. The silicon die itself is underneath the metal. The pinkish regions are active silicon, doped to make it conductive. The speckled golden lines are regions are polysilicon wires between the silicon and the metal. It has two roles: most importantly, when polysilicon crosses active silicon, it forms the gate of a transistor. But polysilicon is also used as wiring, important since this chip only has one layer of metal. The large, dark circles are contacts, connections between the metal layer and the silicon. Smaller square regions are contacts between silicon and polysilicon.

Two memory cells, side by side, as they appear under the microscope.

Two memory cells, side by side, as they appear under the microscope.

It was too difficult to interpret the circuits when they were obscured by the metal layer so I dissolved the metal layer and oxide with hydrochloric acid and Armour Etch respectively. The photo below shows the die with the metal removed; the greenish areas are remnants in areas where the metal was thick, mostly power and ground supplies. The dark regions in this image are regions of doped silicon. These are the active areas of the chip, showing the blocks of circuitry. There are also some thin lines of polysilicon wiring. The memory buffer is the large block on the right, just below the center.

The chip with the metal layer removed. Click to zoom in on the image.

The chip with the metal layer removed. Click to zoom in on the image.

Like most implementations of static RAM, each storage cell of the buffer is implemented with cross-coupled inverters, with the output of one inverter feeding into the input of the other. To write a new value to the cell, the new value simply overpowers the inverter output, forcing the cell to the new state. To support this, one of the inverters is designed to be weak, generating a smaller signal than a regular inverter. Most circuits that I've examined create the inverter by using a weak transistor, one with a longer gate. This chip, however, uses a circuit that I haven't seen before: an additional transistor, configured to limit the current from the inverter.

The schematic below shows one cell. Each cell uses ten transistors, so it is a "10T" cell. To support multiple reads and writes, each row of cells has three horizontal control signals: one to write to the word, and two to read. Each bit position has one vertical bit line to provide the write data and two vertical bit lines for the data that is read. Pass transistors connect the bit lines to the selected cells to perform a read or a write, allowing the data to flow in or out of the cell. The symbol that looks like an op-amp is a two-transistor NMOS buffer to amplify the signal when reading the cell.

Schematic of one memory cell.

Schematic of one memory cell.

With the metal layer removed, it is easier to see the underlying silicon circuitry and reverse-engineer it. The diagram below shows the silicon and polysilicon for one storage cell, corresponding to the schematic above. (Imagine vertical metal lines for power, ground, and the three bitlines.)

One memory cell with the metal layer removed. I etched the die a few seconds too long so some of the polysilicon is very thin or missing.

One memory cell with the metal layer removed. I etched the die a few seconds too long so some of the polysilicon is very thin or missing.

The output from the memory unit contains a byte swapper. A 16-bit word is generated with the left half from the read 1 output and the second half from the read 2 output, but the bytes can be swapped. This was probably used to read an aligned 16-bit word if it was unaligned in memory.

Parity circuits

In the lower right part of the chip are two parity circuits, each computing the parity of an 8-bit input. The parity of an input is computed by XORing the bits together through a tree of 2-input XOR gates. First, four gates process pairs of input bits. Next, two XOR gates combine the outputs of the first gates. Finally, an XOR gate combines the two previous outputs to generate the final parity.

The arrangement of the 14 XOR gates to compute parity of the two 8-bit values A and B.

The arrangement of the 14 XOR gates to compute parity of the two 8-bit values A and B.

The schematic below shows how an XOR gate is built from a NOR gate and an AND-NOR gate. If both inputs are 0, the first NOR gate forces the output to 0. If both inputs are 1, the AND gate forces the output to 0. Thus, the circuit computes XOR. Each labeled block above implements the XOR circuit below.

Schematic of an XOR gate.

Schematic of an XOR gate.

Conclusion

My conclusion is that the processor for the Keystone II board is probably one of the other chips, one of the IBM metal-can MST packages, and this chip helps with data movement in some way. It would be possible to trace out the complete circuitry of the chip and determine exactly how it functions, but that is too time-consuming a project for this relatively obscure chip.

Follow me on Twitter @kenshirriff or RSS for more chip posts. I'm also on Mastodon occasionally as @[email protected]. Thanks to Al Kossow for providing the chip and Dag Spicer for providing photos. Thanks to Eric Schlaepfer for discussion.

Notes and references

  1. The 3274 Control Unit was replaced by the 3174 Establishment Controller, introduced in 1986. An "Establishment Controller" managed a cluster of peripherals or PCs connected to a host mainframe, essentially a box that provided a "kitchen-sink" of functionality including terminal support, local disk storage, Ethernet or token-ring networking, ASCII terminal support, encryption/decryption, and modem support. These units ranged from PC-sized boxes to mini-fridge-sized boxes, depending on how much functionality was required. 

  2. I'm serious that my laptop can barely handle one person; my 2017 MacBook Air starts dropping characters if it has even a moderate load, and I have to start one-finger typing. You would think that a 1.8 GHz dual-core i5 processor could handle more than 2 characters per second. I don't know if there's something wrong with it, or if modern software just has too much overhead. Don't worry, I upgraded and do most of my work on a faster, more recent laptop. 

  3. The IBM hardware model had the CPU focusing on the big picture, while the hierarchy of boxes underneath processed data, performed storage, handled printing, and so forth. In a sense, this paralleled the structure of offices in that era, where executives had assistants and secretaries to do the tedious work for them: typing, filing, and so forth. Nowadays, the computer hierarchy and the office hierarchy are both considerably flatter. Maybe there's a connection? 

  4. A ROM and a PLA are similar in many ways. The general distinction is that a ROM activates one word (row) at a time, while a PLA can activate multiple rows at a time and combine the values, giving more flexibility. A ROM generally has a binary decoder to select the row. This decoder can be recognized by its binary structure: transistors alternating by 1's, by 2's, by 4's, and so forth. 

Standard cells: Looking at individual gates in the Pentium processor

Intel released the powerful Pentium processor in 1993, a chip to "separate the really power-hungry folks from ordinary mortals." The original Pentium was followed by the Pentium Pro, the Pentium II, and others, spawning a long-running brand of high-performance processors, Intel's flagship line until the Core processors took over in 2006. The Pentium eventually became virtually synonymous with "PC" and even made it into pop culture.

Even though the Pentium is a complex chip with 3.3 million transistors, its transistors are visible under a microscope, unlike modern chips. By examining the chip, we can see the interesting circuits used for gates, flip-flops, and other circuits, including the use of an unusual technology called BiCMOS. In this article, I take a close look at the original Pentium chip1, showing how much of its circuitry was built out of structured rows of tiny transistors, a technique known as standard-cell design.

The die photo below shows the Pentium's fingernail-sized silicon die under a microscope. I removed the chip's four metal layers to show the underlying silicon, revealing the individual transistors, which are obscured in most die photos by the layers of metal. Standard-cell circuitry, indicated by red boxes, is recognizable because the circuitry is arranged in uniform columns of cells, giving it a characteristic striped appearance. In contrast, the chip's manually-optimized functional blocks are denser and more structured, giving them a darker appearance. Examples are the caches on the left, the datapaths in the middle, and the microcode ROMs on the right.

Die photo of the Intel Pentium processor with standard cells highlighted in red. The edges of the chip suffered some damage when I removed the metal layers. Click this image (or any other) for a larger version.

Die photo of the Intel Pentium processor with standard cells highlighted in red. The edges of the chip suffered some damage when I removed the metal layers. Click this image (or any other) for a larger version.

Standard-cell design

Early processors in the 1970s were usually designed by manually laying out every transistor individually, fitting transistors together like puzzle pieces to optimize their layout. While this was tedious, it resulted in a highly dense layout. Federico Faggin, designer of the popular Z80 processor, was almost done when he ran into a problem. The last few transistors wouldn't fit, so he had to erase three weeks of work and start over. The closeup of the resulting Z80 layout below shows that each transistor has a different, complex shape, optimized to pack the transistors as tightly as possible.2

A closeup of transistors in the Zilog Z80 processor (1976). This chip is NMOS, not CMOS, which provides more layout flexibility. The metal and polysilicon layers have been removed to expose the underlying silicon. The lighter stripes over active silicon indicate where the polysilicon gates were. I think this photo is from the Visual 6502 project but I'm not sure.

A closeup of transistors in the Zilog Z80 processor (1976). This chip is NMOS, not CMOS, which provides more layout flexibility. The metal and polysilicon layers have been removed to expose the underlying silicon. The lighter stripes over active silicon indicate where the polysilicon gates were. I think this photo is from the Visual 6502 project but I'm not sure.

Because manual layout is slow, difficult, and error-prone, people developed automated approaches such as standard-cell.3 The idea behind standard-cell is to create a standard library of blocks (cells) to implement each type of gate, flip-flop, and other low-level component. To use a particular circuit, instead of arranging each transistor, you use the standard design from the library. Each cell has a fixed height but the width varies as needed, so the standard cells can be arranged in rows. The Pentium die photo below shows seven cells in a row. (The rectangular blobs are doped silicon while the long, thin vertical lines are polysilicon.) Compare the orderly arrangement of these transistors with the Z80 transistors above.

Some standard cell circuitry in the Pentium.
I removed the metal to show the underlying silicon and polysilicon.

Some standard cell circuitry in the Pentium. I removed the metal to show the underlying silicon and polysilicon.

The photo below zooms out to show five rows of standard cells (the dark bands) and the wiring in between. Because CMOS circuitry uses two types of transistors (NMOS and PMOS), each standard-cell row appears as two closely-spaced bands: one of NMOS transistors and one of PMOS transistors. The space between rows is used as a "wiring channel" that holds the wiring between the cells. Power and ground for the circuitry run along the top and bottom of each row.

Some standard cells in the Pentium processor.

Some standard cells in the Pentium processor.

The fixed structure of standard cell design makes it suitable for automation, with the layout generated by "automatic place and route" software. The first step, placement, consists of determining an arrangement of cells that minimizes the distance between connected cells. Running long wires between cells wastes space on the die, since you end up with a lot of unnecessary metal wiring. But more importantly, long paths have higher capacitance, slowing down the signals. Once the cells are placed in their positions, the "routing" step generates the wiring to connect the calls. Placement and routing are both difficult optimization problems that are NP-complete.

Intel started using automated place and route techniques for the 386 processor, since it was much faster than manual layout and dramatically reduced the number of errors. Placement was done with a program called Timberwolf, developed by a Berkeley grad student. As one member of the 386 team said, "If management had known that we were using a tool by some grad student as a key part of the methodology, they would never have let us use it." Intel developed custom software for routing, using an iterative heuristic approach. Standard-cell design is still used in current processors, but the software is much more advanced.

A brief overview of CMOS

Before looking at the standard cell circuits in detail, I'll give a quick overview of how CMOS circuits are implemented. Modern processors are built from CMOS circuitry, which uses two types of transistors: NMOS and PMOS. The diagram below shows how an NMOS transistor is constructed. The transistor can be considered a switch between the source and drain, controlled by the gate. The source and drain regions (green) consist of silicon doped with impurities to change its semiconductor properties, forming N+ silicon. The gate consists of a layer of polysilicon (red), separated from the silicon by a very thin insulating oxide layer. Whenever polysilicon crosses active silicon, a transistor is formed. Diagram showing the structure of an NMOS transistor.

Diagram showing the structure of an NMOS transistor.

The NMOS and PMOS transistors are opposite in their construction and operation. A PMOS transistor swaps the N-type and P-type silicon, so it consists of P+ regions in a substrate of N silicon. In operation, an NMOS transistor turns on when the gate is high, while a PMOS transistor turns on when the gate is low.4 An NMOS transistor is best at pulling its output low, while a PMOS transistor is best at pulling its output high. In a CMOS circuit, the transistors work as a team, pulling the output high or low as needed; the "C" in CMOS indicates this "Complementary" approach. NMOS and PMOS transistors are not entirely symmetrical, however, due to the underlying semiconductor physics. Instead, PMOS transistors need to be larger than NMOS transistors, which helps to distinguish PMOS transistors from NMOS transistors on the die.

The layers of circuitry in the Pentium

The construction of the Pentium is more complicated than the diagram above, with four layers of metal wiring that connect the transistors.5 Starting at the surface of the silicon die, the Pentium's transistors are similar to the diagram, with regions of silicon doped to change their semiconductor properties. Polysilicon wiring is created on top of the silicon. The most important role of the polysilicon is that when it crosses doped silicon, a transistor is formed, with the polysilicon as the gate. However, polysilicon is also used as wiring over short distances.

Above the silicon, four layers of metal connect the components: multiple metal layers allow signals to crisscross the chip without running into each other. The metal layers are numbered M1 through M4, with M1 on the bottom. A few rules control the wiring: a metal layer can connect with the layer above or below through a tungsten plug called a "via". Only the bottom metal, M1, can connect to the silicon or polysilicon, through a "contact". The layers usually alternate between horizontal wiring and vertical wiring (at least locally). Thus, a signal from a transistor may travel through M1, bounce up to M2 and M3 to cross other signals, and then go back down to M1 to connect to another transistor. As you can see, automated place and route software has a complicated task, producing millions of complicated wiring paths as densely as possible.

The diagram below shows how the layers appear on the chip. (This photo shows one of the rare spots on the chip where all the layers are visible.) The M4 metal layer on top of the chip is the thickest, so it is mostly used for power, ground, and clock signals rather than data. An M4 ground wire covers the top of this photo. The next layer down is M3. In this part of the chip, M3 lines run vertically. (Due to optical effects, the vertical M3 lines may look like they are on top of M4, but they are below.) The horizontal M2 metal lines are lower and appear brown rather than golden, due to the oxide layers that cover them. The bottom metal layer is M1. The vertical M1 lines are thick in this part of the chip because they provide power to the circuitry.

The Pentium is constructed with four layers of metal. Because the chip has a three-dimensional structure, I used focus stacking to get a clearer image.

The Pentium is constructed with four layers of metal. Because the chip has a three-dimensional structure, I used focus stacking to get a clearer image.

The silicon and polysilicon are mostly obscured in the above photo. By removing all the metal layers, I obtained the image below. This image shows the same region as the image above, but it is hard to see the correlation because the metal layers almost completely obscure the silicon. The orderly columns of transistors reveal the standard-cell design. The irregular dark regions are doped silicon, which forms the chip's transistors. The dark or shiny horizontal bands are polysilicon. I will explain below how these regions form gates and other circuits.

A closeup of the silicon and polysilicon.

A closeup of the silicon and polysilicon.

Inverter

The fundamental CMOS gate is an inverter, shown in the schematic below. The inverter is built from one PMOS transistor (top) and one NMOS transistor (bottom). If the gate input is a "1", the bottom transistor turns on, pulling the output to ground (0). A "0" input turns on the top transistor, pulling the output high (1). Thus, this two-transistor circuit implements an inverter.10

Schematic diagram of a CMOS inverter.

Schematic diagram of a CMOS inverter.

The diagram below shows two views of how a standard-cell inverter appears on the Pentium die, with and without metal. The inverter consists of two transistors, just like the schematic above. The input is connected to the two polysilicon gates of the transistors. The metal output wire is connected to the two transistors (the left sides, specifically).

A standard-cell CMOS inverter in the Pentium.

A standard-cell CMOS inverter in the Pentium.

In more detail, the image on the left includes the bottom (M1) metal layer, but I removed the other metal layers. Two thick metal lines at the top and bottom provide power and ground to the standard cells. The multiple dark circles are contacts between the M1 metal layer and the metal layer on top (M2), providing a path for power and ground that eventually reaches the top (M4) metal layer and then the chip's pins. (The power and ground wires are thick to provide sufficient current to the circuitry while minimizing voltage drops and noise.) The small, lighter circles are vias that connect the M1 metal layer to the underlying silicon or polysilicon. The input to the gate is provided from the M2 metal, which connects to the M1 layer at the indicated contact. The smaller black dots at the top and bottom of this metal strip are vias, connections to the underlying silicon.

For the image on the right, I removed all four metal layers, revealing the polysilicon and doped silicon. Recall that a transistor is constructed from regions of doped silicon with a stripe of polysilicon between the regions, forming the transistor's gate. The diagram shows the two transistors that form the inverter. When combined with the metal wiring, they form the inverter schematic shown earlier. The final feature is the "well tap". The PMOS transistors are constructed in a "well" of N-doped silicon. The well must be kept at a positive voltage, so periodic "taps" connect the well to the +3.3V supply. As mentioned earlier, the PMOS transistor is larger than the NMOS transistor, which allowed me to figure out the transistor types in the photo.

By the way, the chip is built with a 600 nm process, so the width of the polysilicon lines is approximately 600 nm. For comparison, the wavelength of visible light is 400 to 700 nm, with 600 nm corresponding to orange light. This explains why the microscope photos are somewhat fuzzy; the features are the size of the wavelength of light.6

NAND gate

Another common gate in the Pentium is the NAND gate. The schematic below shows a NAND gate with two PMOS transistors above and two NMOS transistors below. If both inputs are high, the two NMOS transistors turn on, pulling the output low. If either input is low, a PMOS transistor turns on, pulling the output high. (Recall that NMOS and PMOS are opposites: a high voltage turns an NMOS transistor on while a low voltage turns a PMOS transistor on.) Thus, the CMOS circuit below produces the desired output for the NAND function.

Schematic of a CMOS NAND gate.

Schematic of a CMOS NAND gate.

The implementation of the gate as a standard cell, below, follows the schematic. The left photo shows the circuit with one layer of metal (M1). A thick metal line provides 3.3 volts to the gate; it has two contacts that provide power to the two PMOS transistors. The metal line for ground is similar, except only one NMOS transistor is grounded. The thinner metal in the middle has two contacts to get the transistor outputs and a via to connect the output to the M2 metal layer on top. Finally, two tiny bits of M1 metal connect the inputs from the M2 layer to the underlying polysilicon.

Implementation of a CMOS NAND gate as a standard cell.

Implementation of a CMOS NAND gate as a standard cell.

The right photo shows the circuit with all metal removed, showing the polysilicon and silicon. Since a transistor is formed where a polysilicon line crosses doped silicon, the two polysilicon lines create four transistors. Polysilicon functions both as local wiring and as the transistor gates. In particular, the inputs can be connected at the top or bottom of the circuit (or both), depending on what works best for wiring the circuitry. Note that the transistors are squashed together so the silicon in the middle is part of two transistors. An important asymmetry is that the output is taken from the middle of the PMOS transistors, wiring them in parallel, while the output is taken from the right side of the NMOS transistors, wiring them in series.

Zooming out a bit, the photo below shows three NAND gates. Although the underlying standard cell is the same for each one, there are differences between the gates. At the top, horizontal wiring links the inputs to M2 through vias. The length of each polysilicon line depends on the position of the metal. Moreover, in the middle of each gate, the metal connection to the output is positioned differently. Finally, note that the power wiring shifts upward in the upper right corner; this is to make room for a larger cell to the right. The point is that the standard cells aren't simply copies of each other, but are adjusted in each case to put the inputs, outputs, and power in the right location. Also note that these standard cells are not isolated, but are squeezed together so the PMOS transistors are touching. This optimization slightly increases the density.

Three NAND gates in the Pentium.

Three NAND gates in the Pentium.

OR-NAND gate

The standard cell library includes some complex gates. For instance, the gate below is a 5-input OR-NAND gate, computing ~((A+B+C+D)⋅E). In the NMOS circuit, transistors A through D are paralleled while E is in series. The PMOS circuit is the opposite, with A through D in series and E in parallel. To provide sufficient current, the PMOS circuit has two sets of transistors for A through D, so the PMOS block is much larger than the NMOS block.

The OR-NAND gate as it appears on the die. The left image shows the M1 metal layer while the right image shows the silicon
and polysilicon.

The OR-NAND gate as it appears on the die. The left image shows the M1 metal layer while the right image shows the silicon and polysilicon.

Latch

One of the key building blocks of the Pentium's circuitry is the latch. The idea of the latch is to hold one bit, controlled by the clock signal. A latch is "transparent": the latch's input immediately appears on the output while the clock is high. But when the clock is low, the latch holds its previous value. The latch is implemented with a feedback loop that passes the latch's output back into the latch. The heart of this latch circuit is the multiplexer (mux), which selects either the previous output (when the clock is low) or the new input (when the clock is high). The inverters amplify the feedback signal so it doesn't decay in the loop. An inverter also amplifies the output so it can drive other circuitry.

The circuit for a latch.

The circuit for a latch.

The circuit for a multiplexer is interesting since it uses "pass transistors". That is, the transistors simply pass their input through to the output, rather than pulling a signal to power or ground as in a typical logic gate. The schematic shows how this works. First, suppose that the select line is low. This will turn on the two transistors connected to the first input, allowing its level to flow to the output. Meanwhile, both transistors connected to the second input will be turned off, blocking that signal. But if the select line is high, everything switches. Now, the two transistors connected to the second input turn on, passing its level to the output. Thus, the multiplexer selects the first input if the control signal is low, and the second input if the control signal is high.

A multiplexer and its implementation in CMOS.

A multiplexer and its implementation in CMOS.

The diagram below shows a multiplexer, part of a latch. On the left, an inverter feeds into one input of the multiplexer.7 On the right is the other input to the multiplexer. The output is taken from the middle, between the pairs of the transistors.

A multiplexer as it appears on the Pentium die.

A multiplexer as it appears on the Pentium die.

Note that the multiplexer's circuit is opposite, in a way, to a logic gate. In a logic gate, you want either the NMOS transistor on or the PMOS transistor on, so the output is pulled low or high respectively. This is accomplished by giving the signals on the transistor gates the same polarity, so the same polysilicon line runs through both transistors. In a multiplexer, however, you want the corresponding PMOS and NMOS transistors to turn on at the same time, so they can pass the signal. This requires the signals on the transistor gates to have opposite polarity. One polysilicon line runs through the right PMOS transistor and the left NMOS transistor. The other polysilicon line runs through the left PMOS transistor and the right NMOS transistor, connected by metal wiring (not shown). The multiplexer includes an inverter to provide the necessary signal, but I cropped it out of the diagram below.

The flip-flop

The Pentium makes extensive use of flip-flops. A flip-flop is similar to a latch, except its clock input is edge-sensitive instead of level-sensitive. That is, the flip-flop "remembers" its input at the moment the clock goes from low to high, and provides that value as its output. This difference may seem unimportant, but it turns out to make the flip-flop more useful in counters, state machines, and other clocked circuits.

In the Pentium, a flip-flop is constructed from two latches: a primary latch and a secondary latch. The primary latch passes its value through while the clock is low and holds its value when the clock is high. The output of the primary latch is fed into the secondary latch, which has the opposite clock behavior. The result is that when the clock switches from low to high, the primary latch stops updating its output at the same time that the secondary starts passing this value through, providing the desired flip-flop behavior.

A standard-cell flip-flop.

A standard-cell flip-flop.

The photo above shows a standard-cell flop-flop, with an intricate pattern of metal wiring connecting the various sub-components. There are a few variants; with minor logic changes, the flip-flop can have "set" or "reset" inputs, bypassing the clock to force the output to the desired state. (Set and reset functions are useful for initializing flip-flops to a desired value, for example when the processor starts up.)

The BiCMOS buffer

Although I've been discussing CMOS circuits so far, the Pentium was built with BiCMOS, a process that allows circuits to use bipolar transistors in addition to CMOS. By adding a few extra processing steps to the regular CMOS manufacturing process, bipolar (NPN and PNP) transistors can be created. The Pentium made extensive use of BiCMOS circuits since they reduced signal delays by up to 35%. Intel also used BiCMOS for the Pentium Pro, Pentium II, Pentium III, and Xeon processors (but not the Pentium MMX). However, as chip voltages dropped, the benefit from bipolar transistors dropped too and BiCMOS was eventually abandoned.

The schematic below shows a standard-cell BiCMOS buffer in the Pentium chip.8 This circuit is more complex than a CMOS buffer: it uses two inverters, an NPN pull-up transistor, an NMOS pull-down transistor, and a PMOS pull-up transistor.9

Reverse-engineered schematic of the BiCMOS buffer.

Reverse-engineered schematic of the BiCMOS buffer.

In the die images below, note the circular structure of the NPN transistor, very different from the linear structure of the NMOS and PMOS transistors and considerably larger. A sign of the buffer's high-current drive capacity is the output's thick metal wiring, much thicker than the typical signal wiring.

A BiCMOS buffer in the Pentium.

A BiCMOS buffer in the Pentium.

Conclusions

Standard-cell layout is extensively used in modern chips. Modern processors, with their nanometer-scale transistors, are much too small to study under a microscope. The Pentium, on the other hand, has features large enough that its circuits can be observed and reverse engineered. Of course, with 3.3 million transistors, the Pentium is too much for me to reverse engineer in depth, but I still find it interesting to study small-scale circuits and see how they were implemented. This post presented a small sample of the standard cells in the Pentium. The full standard-cell library is much larger, with dozens, if not hundreds, of different cells: many types of logic gates in a variety of sizes and drive strengths. But the fundamental design and layout principles are the same as the cells described here.

One unusual feature of the Pentium is its use of BiCMOS circuitry, which had a peak of popularity in the 1990s, right around the era of the Pentium. Although changing tradeoffs made BiCMOS impractical for digital circuitry, BiCMOS still has an important role in analog ICs, especially high-frequency applications. The Pentium in a sense is a time capsule with its use of BiCMOS.

I hope that you have enjoyed this look at some of the Pentium's circuits. I find it reassuring to see that even complex processors are made up of simple transistor circuits and you can observe and understand these circuits if you look closely.

For more on standard-cell circuits, I wrote about standard cells in an IBM chip and standard cells in the 386 (the 386 article has a lot of overlap with this one). Follow me on Twitter @kenshirriff or RSS for updates. I'm also on Mastodon occasionally as @[email protected].

Notes and references

  1. In this blog post, I'm focusing on the "P54C" version of the original Pentium processor. Intel produced many different versions of the Pentium, and it can be hard to keep them straight. Part of the problem is that "Pentium" is a brand name, with multiple microarchitectures, lines, and products. At the high level, the Pentium (1993) was followed by the Pentium Pro (1995) Pentium II (1997), Pentium III (1999), Pentium 4 (2000), and so on. The original Pentium used the P5 microarchitecture, a superscalar microarchitecture that was advanced but still executed instruction in order like traditional microprocessors. The Pentium Pro was a major jump, implementing a microarchitecture called P6 that broke instructions into micro-operations and executed them out of order using dataflow techniques. The next microarchitecture version was NetBurst, first used with the Pentium 4. NetBurst provided a deep pipeline and introduced hyper-threading, but it was disappointingly slow and was replaced by the Core microarchitecture. The Core microarchitecture is based on the P6 and is Intel's current microarchitecture.

    I'll focus now on the original Pentium, which went through several substantial revisions. The first Pentium product was the 80501 (codenamed P5), running at 60 or 66 MHz and using 5 volts. These chips were built with an 800 nm process and contained 3.1 million transistors.

    The power consumption of these chips was disappointing, so Intel improved the chip, producing the 80502. These chips, codenamed P54C, used 3.3 volts and ran at 75-120 MHz. The chip's architecture remained essentially the same but support was added for multiprocessing, boosting the transistor count to 3.3 million. The P54C had a much more advanced clock circuit, allowing the external bus speed to stay low (50-66 MHz) while the internal clock speed—and thus performance—climbed to 100 MHz. The chips were built with a smaller 600 nm process with four layers of metal, compared to the previous three. Visually, the die of the P54C is almost the same as the P5, with the additional multiprocessing logic at the bottom and the clock circuitry at the top. For this article, I examined the P54C, but the standard cells should be similar in other versions.

    Next, Intel moved to the 350 nm process, producing a smaller, faster Pentium chip, codenamed the P54CS; the die looks almost identical to the P54C (but smaller), with subtle changes to the bond pads. Another variant was designed for mobile use: the Pentium processor with "Voltage Reduction Technology" reduced power consumption by using a 2.9- or 3.1-volt supply for the core and a 3.3-volt supply to drive the I/O pins. These were built first with the 600 nm process (75-100 MHz) and then the 350 nm process (100-150 MHz).

    The biggest change to the original Pentium was the Pentium MMX, with part number 80503 and codename P55C. This chip extended the x86 instruction set with 57 new instructions for vector processing. It was built on a 350 nm process before moving to 280 nm, and had 4.5 million transistors. More obscure variants of the original Pentium include the P54CQS, P54CS, P54LM, P24T, and Tillamook, but I won't get into them. 

  2. Circuits that had a high degree of regularity, such as the arithmetic/logic unit (ALU) or register storage were typically constructed by manually laying out a block to implement the circuitry for one bit and then repeating the block as needed. Because a circuit was repeated 32 times for the 32-bit processor, the additional effort was worthwhile. 

  3. An alternative layout technique is the gate array, which doesn't provide as much flexibility as a standard cell approach. In a gate array (sometimes called a master slice), the chip had a fixed array of transistors (and often resistors). The chip could be customized for a particular application by designing the metal layer to connect the transistors as needed. The density of the chip was usually poor, but gate arrays were much faster to design, so they were advantageous for applications that didn't need high density or produced a relatively small volume of chips. Moreover, manufacturing was much faster because the silicon wafers could be constructed in advance with the transistor array and warehoused. Putting the metal layer on top for a particular application could then be quick. Similar gate arrays used a fixed arrangement of logic gates or flip-flops, rather than transistors. Gate arrays date back to 1967

  4. The behavior of MOS transistors is complicated, so the description above is simplified, just enough to understand digital circuits. In particular, MOS transistors don't simply switch between "on" and "off" but have states in between. This allows MOS transistors to be used in a wide variety of analog circuits. 

  5. The earliest Pentiums had three layers of metal wiring, but Intel moved to a four-layer process with the P54C die, the version that I'm examining. 

  6. To get this level of magnification with my microscope, I had to use an oil immersion lens. Instead of looking at the chip in air, as with a normal lens, I had to put a drop of special microscope oil on the chip. I carefully lower the lens until it dips into the oil (making sure I don't crash the lens into the chip). The purpose of the oil is that its index of refraction is almost the same as glass, much higher than air. This gives the lens a higher "numerical aperture", allowing the lens to resolve smaller details. 

  7. For completeness, I'll mention that the inverter feeding the multiplexer inverter isn't exactly an inverter. Specifically, the inverter's two transistors are not tied together to produce an output. Instead, the inverter's NMOS transistor provides an input to the multiplexer's NMOS transistor and likewise, the PMOS transistor provides an input to the PMOS transistor. The omission of this connection does not affect the circuit's behavior, but it makes calling the circuit an inverter and a multiplexer a bit of an abstraction. 

  8. Intel called this gate "BiNMOS" rather than "BiCMOS" because it uses a bipolar transistor and an NMOS transistor to drive the output, rather than two bipolar transistors. The Pentium's BiCMOS circuitry is described in a conference paper, showing a second NPN transistor to protect the first one. I don't see the second transistor on the die so the two transistors may be implemented in one silicon structure. Reference: R. F. Krick et al., “A 150 MHz 0.6 µm BiCMOS superscalar microprocessor,” IEEE Journal of Solid-State Circuits, vol. 29, no. 12, Dec. 1994, doi:10.1109/4.340418

  9. The Pentium contains multiple types of BiCMOS standard cells, which I'll show in this footnote. The cell below is an inverter. It is similar to the BiCMOS buffer described earlier, except it lacks the first inverter in the circuit. To make room for the NPN transistor on the left, the PMOS transistors are shifted to the right. As a result, they don't line up with the PMOS transistors in other cells. This is a break from the traditional orderliness of standard cells.

    A BiCMOS inverter with PMOS on the left and NMOS on the right. The input is at the bottom and the output is in the middle.

    A BiCMOS inverter with PMOS on the left and NMOS on the right. The input is at the bottom and the output is in the middle.

    The BiCMOS inverter below is similar, except it uses two NPN transistors, providing more output drive. I removed the M1 metal layer to provide a better view of the transistors.

    A BiCMOS inverter with two NPN transistors. The PMOS transistors are in the lower left and the NMOS transistors are in the lower right.

    A BiCMOS inverter with two NPN transistors. The PMOS transistors are in the lower left and the NMOS transistors are in the lower right.

    Another interesting BiCMOS circuit is the D flip-flop with enable and BiCMOS output, shown below. This is similar to the earlier flip-flop except it has an enable input, allowing it to either load a new value triggered by the clock, or to hold its earlier value. This allows the flip-flop to remember a value for more than one clock cycle. The additional functionality is implemented by another multiplexer, selecting either the old value or the new value. (This multiplexer is, in a way, one level higher than the multiplexer in each latch.) The transistor for the BiCMOS output is in the upper right, poking out from under the metal. (This circuit might be implemented as two independent cells, one for the flip-flop and one for the driver; I'm not sure.)

    A D flip-flop in the Pentium.

    A D flip-flop in the Pentium.

     

  10. One puzzling inverter variant is used in a gate I'll call the "slow buffer". This buffer consists of two inverters, so it passes its input through to the output, buffered. The strange part is that the first inverter uses transistors with wide gates, which makes these transistors much weaker than regular transistors. As a result, the first inverter will be slow to switch states. My guess is that this circuit is used to delay signals, for example, to keep a signal aligned with another signal that is delayed by multiple logic gates.

    The buffer consists of two inverters. The first inverter uses wide, weak transistors.

    The buffer consists of two inverters. The first inverter uses wide, weak transistors.

    You might expect that larger transistors would be stronger, not weaker. The problem is that these transistors are larger in the wrong dimension. If you make the gate wider, the effect is similar to multiple transistors in parallel, providing more current. But if you make the gate longer (as in this case), the effect is similar to multiple transistors in series, so the resistances add and the total current is reduced. In most cases, transistors are constructed with the smallest gate length possible, which is determined by the manufacturing process, so the transistors here are unusual. This chip was manufactured with an 800 nm process, so the smallest gate length is approximately 800 nm. The gate width (the normal direction for variation) varies dramatically depending on the circuit, optimized to provide maximum performance.