Desinging Analog Chips
Desinging Analog Chips
Desinging Analog Chips
February 2005
The author is indebted to the following for comments, suggestions and corrections:
Bob Pease, Jim Feit, Ted Bee, Jon Fischer, Tim Camenzind, Jules Jelinek, Brian
Attwood, Ray Futrell, Beat Seeholzer, David Skurnik, Barry Schwartz, Dale Rebgetz,
Tim Herklots, Jerry Gray, Paul Chic, Mark Leonard, Yut Chow, Gregory Weselak, Lars
Jespersen and Wolfgang Horn.
Camenzind: Designing Analog Chips Table of Contents
Table of Contents
Analog World
1 Devices 1-1
Semiconductors 1-1
The Diode 1-5
The Bipolar Transistor 1-6
The Integrated Circuit 1-13
Integrated NPN Transistors 1-14
The Case of the Lateral PNP Transistor 1-22
CMOS Transistors 1-23
The Substrate PNP Transistor 1-27
Diodes 1-27
Zener Diodes 1-28
Resistors 1-29
Capacitors 1-32
Other Processes 1-33
CMOS vs. Bipolar 1-34
2 Simulation 2-1
What You Can Simulate 2-2
DC Analysis 2-2
AC Analysis 2-3
Transient Analysis 2-4
The Big Question of Variations 2-6
Models 2-8
The Diode Model 2-8
The Bipolar Transistor Model 2-10
The Model for the Lateral PNP Transistor 2-13
MOS Transistor Models 2-14
Resistor Models 2-16
Models for Capacitors 2-17
Pads and Pins 2-17
Just How Accurate is a Model? 2-18
3 Current Mirrors 3-1
4 The Royal Differential Pair 4-1
5 Current Sources 5-1
Bipolar 5-1
CMOS 5-7
The Ideal Current Source 5-7
6 Time Out: Analog Measures 6-1
dB 6-1
RMS 6-2
Noise 6-4
Fourier Analysis, Distortion 6-6
Frequency Compensation 6-9
7 Bandgap References 7-1
Low-Voltage Bandgap References 7-11
Analog World
* * *
You will find that almost all analog ICs contain a number of
recognizable circuit elements, functional blocks with just a few transistors.
These elements have proven useful and thus re-appear in design after
design. Thus it makes sense to first look at such things as current mirrors,
compound transistors, differential stages, cascodes, active loads, Darlington
connections or current sources in some detail and then examine how they
are best put together to form whole functions.
* * *
1 Devices
Semiconductors
In 1874 Ferdinand Braun was a 24-year old teacher in Leipzig,
Germany. He published a paper which was nothing short of revolutionary:
he had found that some materials violated Ohm's law. Using naturally
formed crystals of Galena (lead sulfite, the chief ore mineral of lead) and
other sulfites, he pressed a spring-loaded metal tip against their surfaces and
observed that the current through this arrangement was dependent on the
polarity of the applied voltage. Even more puzzling was the fact that, in the
direction which had better conduction, the resistance decreased as the
current was increased.
What Braun (who later would give us the CRT) had discovered, we
now know as the diode, or rectifier. It was not a very good one, there was
only a 30% difference between forward and reverse current. And there
were no practical applications. Braun could not explain the effect, nor
could anybody else.
In 1879 Edwin Hall of Johns Hopkins University discovered what
was later named the Hall Effect: when you pass a magnetic field through a
piece of metal it deflects the current running through the metal. In all the
metals he tried the deflection was to one side; he was greatly relieved to see
that this confirmed the negative charge on the electron.
But then the surprise came. In some materials the deflection went
the other way. Where there perhaps positive electrons?
Nothing much happened until about 1904. Radio appeared on the
scene and needed a "detector". The signal was amplitude modulated and to
make the music or speech audible the radio frequency needed to be rectified
(i.e. averaged). Thus, 30 years after Braun's discovery, the "odd behavior"
of a wire touching Galena (and now many other materials, such as silicon
carbide, tellurium and silicon) found a practical application. The device
was called the "Cat's whisker", but it actually didn't work very well; one
had to try several spots on the crystal until one was found which produced a
loud enough signal.
And it was replaced almost immediately by the vacuum tube, which
could not only rectify but amplify as well. Thus the semiconductor rectifier
(or diode) went out of fashion.
It was not until 1927 that another practical application appeared:
large-area rectifiers. These were messy, bulky contraptions using copper-
oxide (and later selenium) to produce DC from line voltage, chiefly to
charge car batteries. But there was still no understanding of how these
devices worked.
ability to move from atom to atom is also the basis of electrical conduction:
in conductors the electrons roam widely and are easily enticed to move in
an electrical field, whereas in an insulator they stay close to home.
Electrically, pure silicon is a terribly uninteresting material. It is an
insulator, but not a very good one. The fun begins when we add the right
impurities, or dopants.
Just to the right of silicon in the periodic table is phosphorus,
element number 15. Like Silicon, it has two electrons in the first orbit,
eight in the second but there are five in the third. Now let's say we were
able to pluck out an atom in a block of silicon and replace it with a
phosphorus atom. Four of the valence electrons of this new atom will
circulate with the silicon electrons, but the fifth one won't fit in. This
excess electron creates a negative charge and the silicon becomes what we
now call n-type.
This introduction of excess electrons is unlike static charge. When
you brush your hair so that it stands upright, you have simply moved some
electrons temporarily. When you "dope" silicon, the charge is permanent,
fixed in the crystal lattice (and does not become a battery).
Similarly, to the left of Silicon and one space up in the periodic table
is boron, element number 5. It has two electrons in a first level and three in
a second, a valence of three. If we replace a silicon atom with a boron one,
there is an electron missing and we create a positive charge, or p-type
material. As with the excess electron in n-type silicon, we can apply an
electric field and cause a current to flow, but the net-effect is the flow of
holes, not electrons. This is what makes the Hall effect go the wrong way.
It is important to understand this mechanism of moving holes and
electrons in doped semiconductors. In n-type material an excess
phosphorus electron wanders into the path of a neighboring silicon electron
and displaces it. The displaced electron then takes the orbit of another one
and so on until the last electron ends up at the starting point, the phosphorus
atom.
This endless game of musical chairs - proceeding at near the speed
of light - depends greatly on the temperature. At absolute zero there is no
movement. At about -60oC the movement is sufficient for semiconductor
effect to start in silicon. At about 200oC there is so much movement that
silicon practically becomes a conductor. It is only within a relatively
narrow range, about -55oC to 150oC, that silicon is a useful semiconductor.
In p-type material the movement starts with an electron in the
neighborhood of the boron atom. It fills the vacancy and then is itself
replaced by another electron and so on until the first electron moves away
from the boron atom again. The moving is done by electrons, but the net
effect is a moving hole.
When an electric field is present the movement takes on a direction:
electrons flow toward the positive electrode and are replaced by other
electrons flowing out of the negative electrode.
It is amazing how few dopants it takes to make n-type or p-type
material. Silicon has 5x1022 atoms per cubic centimeter. A doping level
can easily be as low as 5x1015 boron or phosphorus atoms per cubic
centimeter, i.e. one dopant atom for every 10 million silicon atoms. No
wonder it took so long to discover the true nature of the semiconductor
effects; in nature, the number of miscellaneous impurities is far larger than
one in 10 million.
The Diode
Even with a dopant present silicon is uninteresting. It is not a good
conductor and as a resistor it is inferior to metal film or even carbon. But if
we have both n-type and p-type atoms in the same silicon crystal, things
suddenly happen.
Opposite charges attract each other, so the excess electrons near the
border of the n-type section move into the p-type material and stay there.
An electron fills a hole and the electric charges cancel each other.
This only happens over a short distance, as far as an electron (or
hole) can roam. The resulting region is called the space-charge layer or
depletion region.
Now suppose you
connect a voltage to the
two terminals. If the p-
region is connected to the
negative terminal of the
supply and the n-region to
the positive one, you
simply push the charges
away from each other,
enlarging the depletion
region.
Fig. 1-1: A depletion region forms between p-
doped and n-doped semiconductor areas.
If, however, the p-
region is positive and the
n-region negative, you push the charges closer together as the voltage
increases. The closer proximity forces more and electrons and holes to
cross the depletion region. The effect is exponential: at 0.3 Volts (at room
temperature) very little current flows; at 0.6 Volts the current is substantial
and at 0.9 Volts very large.
The expression for the diode voltage is:
Vd⋅q
I1 = Is e kT − 1
kT I1
Vd = ln or
q Is
where Vd = voltage across the diode
k = Boltzman constant (1.38E-23 Joules/Kelvin)
T = the absolute temperature in Kelvin
q = the electron charge (1.6E-19 Coulombs)
I1 = the actual current through the diode
and Is = diffusion current
Note that 1.38E-23 is a more convenient notation for 1.38x10-23.
The diffusion current Is depends on the doping level of n-type and
p-type impurities, the area of the diode and (to a very high degree) on
temperature. A reasonable starting point for a small-geometry IC diode is
Is=1E-16.
The equations neglect a few things. There is a limit in the voltage
that can be applied in the reverse direction. Similar to an arc-over in any
insulator, there comes a point when the electric field becomes too large and
the opposing charges crash into each other. This breakdown voltage
depends on the concentration of dopants: the higher the concentration, the
lower the breakdown voltage.
There is a price to be paid for high breakdown voltage. As the
dopant concentration is lowered, the depletion layer becomes larger and the
higher voltage pushes it deeper yet. This distance must be accommodated
in the design.
The opposing charges in a semiconductor junction are no different
from those on the plates of a capacitor. So every junction has a capacitance;
but since the distance between the electrons and holes changes with applied
voltage, the capacitance becomes voltage dependent. The lower the
voltage, the higher the capacitance, increasing right into the forward
direction.
Lastly, there is resistance in the semiconductor material not taken up
by the depletion region. For our "typical" concentration of 5E15 (atoms per
cubic centimeter, giving a practical breakdown voltage in an IC of about 25
Volts), the resistivity is about 1 Ohm-cm for phosphorus (n-type) and 3
Ohm-cm for boron (p-type). For comparison, aluminum has a resistivity of
2.8 microOhm-cm, copper 1.7 microOhm-cm. Resistivity (ρ or rho) is
R∗ w∗ h Ohm∗ cm∗ cm
ρ= = = Ohm∗ cm (or Ohm-cm)
l cm
calculations there should have been a large change. But the effect - if there
was any - was at least 1500 times smaller than theoretically predicted.
It was at this time, that John Bardeen, 37, joined Shockley's group.
He looked at Shockley's failed experiment and mulled it over in his head for
a few months. In March 1946 he came up with an explanation: it was the
surface of the silicon which killed the effect. Where the silicon stops, the
four valence electrons are no longer neatly tied up by the neighboring
atoms. Bardeen correctly perceived that some of them were left dangling
and thus produced a surface charge (or voltage), which blocked any voltage
applied to an external control electrode.
With this theoretical breakthrough the group now decided to change
directions; instead of attempting to make a device, they investigated the
fundamentals of semiconductor surfaces. It was a long, painstaking
investigation; it took more than a year. On November 17, 1947 Robert B.
Gibney, another member of the group and a physical chemist, suggested
using an electrolyte to counteract the surface charge. On November 20 he
and Brattain wrote a patent disclosure for an amplifying device as tried by
Shockley but using electrolyte on the surface. Then they went to the lab and
made one. The electrolyte was extracted from an electrolytic capacitor with
a hammer and nail. The device worked, the electrolyte did precisely the job
that Gibney thought it would.
But, although this "field effect" device amplified, it was very slow,
amplifying nothing faster than about 8Hz. Brattain and Bardeen suspected
that it was the electrolyte that slowed down the device so, on December 16,
1947, they tried a different approach: a gold spot with a small hole in the
center was evaporated onto germanium, on top of the insulating oxide. The
idea was to place a sharp point-contact in the center without touching the
gold ring, so that the point would make contact with the germanium, while
the insulated gold ring would shield the surface. And now, for the first
time, they got amplification.
There was only one thing wrong with this device: it didn't work as
expected. A positive voltage at the control terminal increased the current
through the device when, according to their theory, it should have decreased
it. Bardeen and Brattain investigated and found they had inadvertently
washed off the oxide before evaporating the gold, so that the gold was in
contact with the germanium. What they were observing was an entirely
different effect, an injection of carriers by the point contact. They realized
that, to make such a device efficient, the distance between the two contacts
at the surface needed to be very small. They evaporated a new gold spot,
split it in half with a razorblade and placed two point contacts on top. Now
the device worked even better and they demonstrated it to the Bell
exponentially, both in the base and the collector. It is not a linear voltage
amplifier; only the currents have a (more or less) linear relationship.
Also notice that the emitter current is always larger than that of the
collector, since it contains both the collector and base current.
We have shown here an NPN transistor. If we reverse all the doping
and the voltages we create a PNP transistor. It works the same way in every
respect except that it is a bit handicapped: it is slower and has a lower gain;
holes, now the minority carriers in the base, just don't move as well as
electrons.
The point-contact transistor was a nightmare to manufacture and had
very poor reliability. Also, these devices were made from germanium,
which has a rather limited useful temperature range. The junction
transistors were made by alloying dopant materials on either side of a flat
piece of germanium or silicon. It was difficult to make the base uniformly
thin and the process created considerable leakage current.
The next big step was again invented at Bell Labs: diffusion. At
room temperature gases mix even if they are held perfectly still. This
happens because each atom or molecule moves around randomly due to the
energy it receives by temperature. The higher the temperature, the more
pronounced is this movement and thus the mixing or diffusion. If the
temperature is high enough (e.g. over 1000oC) such gases can even diffuse
into solid material, though their diffusion speed decreases enormously.
Thus, for example, silicon exposed in a high-temperature furnace to n-type
impurity (gas) atoms develops an n-layer at its surface with a depth as far as
the impurities penetrate. This may require a temperature close to the
melting point of silicon and take several hours for a penetration of just a
few micro-meters, but it is far more controlled than alloying.
Moreover, you can dope repeatedly. Suppose you have a piece of
silicon which has been doped n-type. If you diffuse p-type impurities into
the surface, you convert a layer from n-type to p-type if there are more p-
type impurities than n-type. The junction is located at the depth at which
the two impurities are equal in concentration. A second diffusion of a yet
higher concentration can then convert the material back to n-type again.
However, you have to pay attention to the fact that subsequent exposure to
high temperature causes any previous layer to diffuse further.
There are a few more dopants available too: p-type gallium (rarely
used) and n-type arsenic and antimony. The latter two have the advantage
that they diffuse more slowly than phosphorus or boron. For this reason
they are primarily used early in the process and are thus less affected by
subsequent diffusions.
When, in 1956, the three inventors of the transistor were awarded
the Nobel Prize for physics, only Walter Brattain was still at Bell
Laboratories. John Bardeen had left in 1951 to become a professor at the
University of Illinois and, for his research there in superconductivity, he
received a second Nobel Prize in 1972.
Bill Shockley left Bell Labs in 1954. Banking on his reputation,
which had risen proportionally to the acceptance of the transistor, he
managed to strike a deal with the Beckman Instruments Company. A
subsidiary, called the Shockley Semiconductor Laboratories, was set up in
Palo Alto, California. Shockley's fame had risen to such a height that he
could pick some of the best people. Within a year he had some 20 people -
predominantly Ph.D.s - working for him, among them Robert Noyce, 28,
Gordon Moore, 27, and Jean Hoerni, 32.
For all of these people there was a brief period of fascination after
they joined. But then the true Bill Shockley appeared from behind the glitter
of fame and they discovered that Shockley was, in fact, a rather erratic and
unpleasant man. He would fire his employees for minor mistakes, throw
tantrums over trivial problems and change directions for no apparent
reasons. He incessantly tried innovative management techniques, such as
posting everybody's salaries on the bulletin board.
Noyce and Moore were pushing Shockley to make silicon transistors
using the diffusion approach. Shockley wasn't interested; his hope was for
his laboratory to come up with an entirely new device, a device which
would represent as large a step over the transistor as the transistor had been
over the vacuum tube.
Now totally dissatisfied, the crew talked to Arnold Beckman, the
president of the parent company, and informed him of the impossible
situation. Beckman promised to hire a business-minded individual who
could act as buffer between Shockley and his staff. But the solution didn't
work, Shockley refused to let go of the day-to-day decision-making. Out of
patience, eight staff members reached a deal with the Fairchild Camera and
Instrument Company and, in October 1957, the group departed.
The new company, called Fairchild Semiconductor, was at first an
independent operation, with Fairchild Camera and Instrument holding an
option for a buy-out. The product they began to develop was the one they
had proposed to Shockley. The detailed structure of this device, called the
Mesa transistor, had been tried in germanium before, but not in silicon. It
required two diffusions, both into the same side of a silicon wafer. The first
diffusion was p-type, the second n-type, and the difference in depth between
the two layers created the base region which, for the first time, could be
made with a high degree of accuracy. The top surface of the transistor was
then masked with wax and the exposed silicon etched away, giving the
the metal particles from the welding of the case short it out. Secondly,
photographic methods could be used to delineate not just one but hundreds
of transistors simultaneously. Thus individual, delicate masking of each
transistor was no longer required, giving the planar transistor a huge
potential for reduced cost. Noyce, who was by now the general manager,
saw the advantage of the planar process and quietly moved it into
production.
There was another advantage to the planar transistor: once the
dopant enters the silicon it diffuses in all directions, including sideways.
The P-N junction, therefore, ends up underneath the oxide, never exposed to
either human handling or the contamination of air. For this reason the planar
junction is the cleanest (and most stable) junction ever produced. Fairchild's
customers who, in early 1959, didn't know that their transistors were now
being manufactured by an entirely new process, were surprised to find
leakage currents one thousand times smaller than those of previous
shipments.
While Fairchild flourished, Shockley Transistor went downhill. It
was sold twice, then closed in 1969. Shockley became interested in
sociology and announced a theory called "dysgenics", which proposed that
poor people were doomed to have low IQs. By the time he died in 1989 his
reputation was ruined.
motivation was primarily cost, not size. He realized that it didn't make
sense to fabricate precisely arranged transistors on a wafer, cut them apart,
place them in a housing and arrange them again in on a circuit board; if the
additional components on the circuit board could be placed on the wafer, a
considerable number of manufacturing steps could be saved. Noyce had no
problem visualizing capacitors and resistors made in silicon, he was
constantly dealing with these (unwanted) effects. What was needed,
though, was an inexpensive way to connect all these components on the
wafer. The idea of using wires had no chance in Noyce's mind, it would
have simply been too expensive. But he saw that, in the planar process, this
problem was already solved: the aluminum layer used to connect the
transistors and the wires could also be used between the components.
In1959 Noyce entered his idea into his notebook and filed for a
patent application. Kilby's and Noyce's patent applications were clearly in
interference and a bitter battle between the two companies started in the
courts. Texas Instruments won because Kilby application mentioned a thin
film of gold, thus seemingly anticipating Noyce. Fairchild appealed.
While the two patents were fought over in the courts, neither TI nor
Fairchild could collect any royalties for integrated circuit, which were
already showing explosive growth. So the two companies came to an
agreement, declaring Kilby and Noyce co-inventors of the integrated circuit.
Shortly after this the appeals court handed down its decision: Noyce, not
Kilby, was declared the inventor of the IC.
It could not have been otherwise. Even today every single IC is
made exactly as Noyce described it, while Kilby's approach has long been
abandoned. But the most important contributor to the invention of the IC
was clearly Jean Hoerni with his planar process, for which he has never
been adequately recognized. The planar process rates as one of the great
inventions of the 20th century.
Robert Noyce died in 1990 at age 62. In 2000 Jack Kilby won the
Nobel Prize for the invention of the integrated circuit
Light is then projected through the mask onto the wafer. The higher
the frequency of the light, the
greater the detail, so ultra-violet
light or even x-rays are used.
The photoresist is then
developed and the portions not
exposed to light are washed off.
Fig. 1-4: The first step: A light-sensitive and etch (There are both positive and
resistant layer (photoresist) is spread on the wafer
and exposed to light through a mask.
negative photoresists; you have
the choice of removing the areas
Fig. 1-4: The first step: A light-sensitive and which are either exposed or not
etch-resistant layer (photoresist) is spread on
the wafer and exposed to light through the
exposed to light).
mask. Next the entire wafer is
immersed in an acid which
removes the oxide in the areas
where it is not protected by the
photoresist. In more modern
processes a plasma is used; acid
etches not only downward but
also slightly sideways
underneath the photoresist, while
Fig. 1-5: The photoresist is developed like a plasma etches downward only.
photograph and the wafer is ready for etching. The wafer is then placed
into a furnace (a quartz tube
heated to greater than 1000oC).
A gas carrying the desired
dopant (in this case boron,
arsenic or antimony) swirls
around the wafer and slowly
Fig. 1-6: The oxide is etched away and the diffuses into the surface.
photoresist is removed.
Note two important facts
here: 1. There is a crowding of
dopants near the surface of the
silicon. With time they will
diffuse deeper into the silicon,
but there will always be more
dopants near the surface. Thus
any diffused region has a marked
gradient. 2. Dopants not only
Fig. 1-7: A gas containing N-type dopants
(boron, arsenic or antimony) diffuses slowly into diffuse downward, but also side-
the surface of the wafer at high temperature. ways. (Since supply is more
limited at the very edge, the side-ways diffusion extends to only about half
the distance of the downward one). This places the junction (where n = p)
underneath the oxide and is thus never exposed to the (dirty) environment.
After diffusion the exposed silicon surface is covered again by an
oxide layer so that the wafer is ready for the next masking step, which could
be another diffusion or the
etching of contact holes.
There is an important
feature here, which should not
go unnoticed. SiO2 is glass,
Fig. 1-8: After the diffusion the oxide is re- which is transparent to light.
grown, ready for the next masking step.
The light is reflected at the
bottom of the oxide by the
silicon and interference patterns are created, i.e. the sum of direct and the
reflected light eliminates some frequencies. Thus the color of the oxide
layer depends on its thickness. This not only makes for beautiful
photographs but, more importantly, it allows subsequent masks to be
precisely aligned with previous ones.
Here then is one form of an NPN transistor made with the planar
process. The substrate (the starting wafer) is doped p-type as the silicon is
grown. There are three diffusions in succession, the first being rather deep.
After the diffusions, contact holes are made (with the same basic photoresist
process), aluminum is deposited over the entire wafer, patterned (another
photoresist step) and etched away where it is not wanted.
Alas, this transistor
has a rather significant
shortcoming: high collector
resistance. The current has
to flow through the region
between the base and the
substrate. That is the far end
of the collector diffusion, the
end which has the fewest
Fig. 1-9: A simple planar NPN transistor. dopant atoms and therefore
the highest resistance.
Since the invention of the planar process a few more ways of
fabricating have been added:
Epitaxy. If you strip a silicon wafer of its oxide and put it into a
furnace which is filled with gas containing not only a dopant but also
silicon, you can grow a doped single-crystal layer. As the atoms carried by
the gas deposit themselves on the surface of the wafer, they will align
Gain versus Current. For any bipolar transistor the current gain falls off
both at low and high current.
First, the low end. There is always a leakage current across any
junction; for a perfectly clean surface this is the diffusion current. In the
base-emitter junction this leakage current takes away a portion of the
supplied base current. In our graph here the current shunted by leakage at
the low end (10nA Ie, or about 50pA Ib) amounts to 33% of Ib, i.e. the gain
has dropped by one third.
If you extend this plot to much lower current, you will see the gain
rise to almost infinity. This is nothing more than the effect of the collector-
base leakage current.
At the high end two effects take place simultaneously: 1. The
number of electrons present in the base simply becomes so large that they
are no longer the minority carriers and the whole effect comes to a halt. 2.
The base current must flow
350
from the contact to the flat
area between the emitter and 300
place base contacts on both sides of the emitter and lengthen the emitter.
Shown here on the left is the top view of a minimum-geometry transistor
and on the right a version for higher current.
To make the life of a designer easier, the isolation pattern is usually
drawn as a rectangle and then inverted when making the mask, i.e. the
isolation diffusion is actually between devices, not in the device area.
Many processes require that all contacts be the same size, in which
case the contact rectangles must be broken up into small, identical (and
properly spaced) squares.
Be aware, that transistors of different sizes (as drawn here) do not
match well. At low current a large emitter area produces a higher gain than
a small one, because the minority carriers have a higher chance to be
captured by the collector. If you want to produce a precise ratio, use only
one emitter size and identical base contacts. The emitters can be in a
common base area and the collector size is of no consequence except for
collector resistance (or saturation voltage).
Substrate Current. There is only leakage current across the
collector-substrate junction, unless the transistor saturates.
Assume the collector is connected through a resistor to the positive
supply voltage and the base is driven so hard that the collector voltage drops
to near the potential of the emitter (termed
Collector
saturation).
There are now two diodes in parallel and
the base current has two paths; the new one forms a Base NPN
PNP transistor with the NPN base becoming the
emitter, the NPN collector the base and the
PNP
substrate the collector. Since the NPN collector is
much larger than its emitter, some (or all) of the
base current flows to the substrate.
Emitter Substrate
There is little danger in this, except when
you drive the base very hard, trying to get the Fig. 1-14: When an NPN
lowest possible collector voltage, or if you have transistor saturates a
stray PNP device leaks
many saturating NPN transistors. The path in the current to the substrate.
substrate from a transistor to the -V connection has
some resistance. If the substrate current is so large
that the voltage drop across this resistance can forward-bias some substrate-
collector junction on the way, you may get some really bad effects,
including latch-up.
Maximum Voltage. To get a high operating voltage requires high
resistivity - low doping concentration. But there is a price to be paid: the
depletion regions become wide.
Despite of all of this, with good surface control you can get a gain
in excess of 100. But the current range is limited, rarely exceeding 100uA
for a minimum geometry device.
And there is somewhat of a problem with substrate current. There is
a competing PNP transistor, using the same emitter and base, but with the
substrate (and the isolation diffusion) as the collector. In normal operation
a current about half the magnitude of the base current flows from the
emitter to the substrate terminal. When the lateral PNP transistor saturates,
the substrate current becomes almost equal to the collector current. If you
don't have a buried layer, it gets quite a bit
worse.
One advantage of the lateral PNP
transistor: the collector can be split into two
(or more) sections. The emitter current,
flowing radially outward is collected by the
segments according to their length at the
inside. There is a small loss in gain because
of the gaps, but the matching between the two
collector currents is excellent.
In a CMOS process emitter and
collector are usually formed by the p-type
diffusion of a p-channel MOS transistor. The
intervening space (the base-width) is the same
as a p-channel gate, with poly-silicon on top.
Connect the poly region to the PNP emitter; it Fig. 1-17: A split-collector
will act as a static shield and have a (slight) lateral PNP transistor.
beneficial effect.
CMOS Transistors
It took almost 20 years after the invention of the bipolar transistor
for MOS to make its appearance. Shockley (and many others) had thought
of this device first, it was (or should have been) much more simple: put a
plate close to the surface of silicon, connect it to a voltage and move the
carriers inside the silicon electro-statically.
The problem was the surface of silicon. Here the silicon atoms are
no longer neatly tied up with each other by sharing the outermost electrons.
They face an entirely different material, SiO2 (or worse, some covering
with unknown impurities mixed in). This material doesn't even have a
crystal structure, it is amorphous.
40
there is direct conduction between
35
the two. This is done with a positive
30
voltage at the gate which pushes
holes away from the surface and the
Drain Current / µA
25
20
device is called an enhancement-
15
mode transistor (there are also
10
depletion-mode devices in which a
5
channel is implanted or diffused and
00
then cut off with a negative gate
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
voltage).
Drain Voltage/V 200mV/div
This is true only at zero or
Fig. 1-20: Drain current vs. drain voltage very low drain voltage. As the drain
with the gate voltage held constant. voltage is increased, a depletion
region forms around it. Since there is now a voltage drop along the
channel, with the drain side at a higher voltage than the source, the
depletion region along the channel gradually increases toward the drain,
cutting more and more into the channel. Thus the resistance of the channel
increases.
The initial slope of the
drain voltage / drain current 200
Drain Current / µA
120
highest drain voltage represents
100
its resistance with the depletion
80
layer almost pinching off the 60
channel. It is an unfortunate 40
fact that this region is called 20
the "saturation region", which 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
clashes badly with the earlier Gate Voltage/V 200mV/div
definition for the bipolar
transistor. Fig. 1-21: Drain current vs. gate voltage
with the drain voltage held constant.
Above a certain gate
potential, which has to be exceeded
to attract any carriers to the surface (the threshold voltage) an MOS
transistor is basically a square-law device: doubling the gate voltage results
in four times the drain current. The measure of gain is the
transconductance, drain current divided by gate voltage. So again, like the
bipolar transistor, this is a non-linear device:
Id = k
W
L
(Vgs − VT )
2
The region below the channel also influences the gain. It forms a
back-gate. In an n-well n-channel transistor this is the substrate, common
to all devices. You have no choice but to connect it to the lowest negative
voltage. But there is a choice for the p-channel transistor. If you place all
the p-channel transistors in a common n-well, you get the smallest total area
and therefore the lowest cost. But if the source of such a transistor is
operated below the positive supply, the back-gate (the n-well) pinches off
the channel further and you get a reduced gain (by perhaps 30%). You can
avoid this by placing this transistor in its own n-well.
Diodes
There are several p-n junctions in an integrated circuit, each and
every one a diode. But few of them can actually be used by themselves
without unpleasant side-effects.
Take a simple bipolar process. There are three types of junctions:
emitter/base, base/collector and collector/substrate
P (all referring to the NPN transistor). The last one is
hardly ever useful because the substrate is
permanently connected to the most negative supply
voltage. The base/collector diode is, as we have
seen, part of a substrate PNP transistor with a gain; a
N current perhaps ten times the magnitude of the diode
Fig. 1-22: Properly current will flow to the substrate.
connected diodes in The emitter/base junction makes a good
a bipolar process.
diode, but it has a low breakdown voltage (about 6
Volts) and the base has a fairly high resistance. You
could connect the surrounding collector to the most negative supply voltage
and thus keep it always reverse-biased. But a much better diode results if
you short collector and base together, creating a diode-connected
transistor. The transistor is active, it has gain. Only a small fraction of the
current flows through the base, which divides the base resistance by the
current gain. This connection in fact gives you an almost ideal diode over
about five decades of current.
If the emitter/base breakdown voltage is too low, consider a diode-
connected lateral PNP transistor. This devices has the full operating voltage
of the process, but it is limited in current (see above).
In a CMOS process the restrictions are even more severe. The only
free-floating junction is between the p-channel source/drain and the n-well.
But, as we have seen, these are also part of the substrate PNP transistor.
Were you to run a current through this junction, a current of about ten times
its magnitude would flow to the substrate.
The term "diode-connected" is often used for an MOS transistor
with its gate and drain connected together. Don't be misled by this term:
there is no diode as in "junction" diode.
Zener Diodes
In a bipolar process the base-emitter diode almost always has a low
breakdown voltage (perhaps 6 Volts) with a fairly low temperature
coefficient, which makes it useful as a reference voltage.
But exercise care with this device: the same junction is also used as
a fusible device.
At low current (e.g. less than 100uA for a minimum-geometry
device) the Zener diode behaves well. As you increase the current the
region between the emitter contact and the edge of the emitter diffusion
lights up faintly (a plasma, which you can observe under a microscope, with
all lights turned off). At some high current level a thin aluminum strip is
formed abruptly underneath the oxide, which converts the Zener diode into
a short-circuit. This effect is used for trimming and carries the earthy name
Zener-zapping.
Such a Zener diode is also somewhat noisy. For lower noise (and
better accuracy) use a bandgap reference.
Moving an n-channel and p-channel source/drain diffusion in a
CMOS process close together so that they intersect can also result in a
useful low breakdown voltage, but data for such a device are rarely
available from the wafer-fab.
Resistors
Every free-floating layer in an integrated circuit can, when properly
patterned, become a resistor. But for all of them this is only a secondary
duty; their intended application is in a transistor, which is the hardest device
to make. It shouldn't come as a surprise then that their values have a higher
variation and greater temperature coefficient and their range is more
restricted than that of even the least expensive discrete resistor.
Discrete resistors can be tested and adjusted during manufacturing.
In ICs the manufacturing is done while the silicon is red-hot, at which
temperature it is no longer a semiconductor; you have to wait until it cools
down to measure any parameter.
What saves the integrated resistor is its natural ability to match well.
Whatever error may have occurred in making one applies to any other on
the same wafer. They may both be as much as 25% high in value, but both
will be high by (almost) exactly the same amount.
The resistance of any material is given by
rho ⋅ l
R=
A
where rho (ρ) = resistivity in Ohm-cm
l = length
A = area (cross-section)
If we make a square, i.e. w = l, then we get a measure of resistance
which is independent of size, the sheet resistance, in Ohms per square (or
Ohms/o).
Note the term is sheet resistance, not sheet resistivity. A square in a
layer with a sheet resistance of 100 Ohms per square always measures 100
Ohms from one side to the other no matter how large the square.
In a bipolar process the layer most often used for resistors is the
(NPN) base (about 200 Ohms/¨). The emitter layer is more heavily doped
and thus has a lower sheet resistance (as low as 5 Ohms/¨).
In a CMOS process you have a wider choice: the n+ and p+
diffusions (implants) for the drains and sources, the n-well and usually two
different poly layers. Of these the p+ diffusion (about 150 Ohms/¨) and
one of the poly layers (around 50 Ohms/¨) are generally best suited.
Sheet resistances depend greatly on the process; you should use the
values given here only as a starting point and get the actual data (including
temperature coefficients and tolerances) from the wafer fab.
Diffused resistors must be placed in an island of opposite doping
and this island must be connected to a bias voltage so that the junction is
reverse-biased. For example, a (p-type) base resistor must be in an n-type
(epi) island. This island (sometimes called the "tub") can contain just one
resistor or all of them, but its voltage must be at a level equal to or greater
than the largest voltage on any resistor. In this case the easiest and safest
connection is to +V.
Diffused resistors (and to a lesser degree, poly resistors) have a
voltage coefficient. The biased surrounding layer pushes a depletion region
into the resistor, reducing its cross-section. As the difference in voltage
between the resistor and the surrounding layer becomes larger, the depletion
region widens, the cross-section becomes smaller and the resistance
increases. This effect is especially pronounced in lightly doped layers: the
n-well in CMOS and implanted resistors in a bipolar process. (The latter
uses an additional implant to create a high sheet resistance).
This voltage dependence is especially critical if you have two (or
more) resistors which need to match but are at different DC levels. You can
place each resistor in a separate island, biased at the positive end of its
resistor. Or you can simply accept the change caused by the depletion layer
and adjust the ratio. For this, however, you need a model for the resistor
which includes its voltage dependence. (in a 200 Ohms/¨ base layer, for
example, the change in resistance is about 1% for a 5V bias difference).
There is also a (distributed) capacitance associated with an
integrated resistor, low for poly, higher (and voltage dependent) for diffused
ones. If you make a high-value (i.e. very long) resistor, this stray
capacitance can seriously cut frequency response. Also, if there is noise on
the supply which biases the surrounding region for diffused resistors, it will
be capacitively coupled into the resistor. Again, a
good model is required to show these effects in a
simulation.
Two correction factors have to be used when
designing a resistor. The first concerns the width of
the resistor. In diffused (or implanted) resistors there
is always a sideways diffusion, which makes the
actual resistor wider than drawn. The effect of the
side-ways diffusion is dependent on the width of the
resistor.
The second correction factor recognizes the Fig. 1-23:
end-effect. If the resistor has minimum width, you Resistor contacts.
will need to enlarge both ends to place a contact inside. You will then need
to estimate the resistance of this additional area and of the contact itself
(totaling perhaps 0.4 squares from the end of the narrow part).
If you draw a wide resistor, the contacts can be fitted inside the
resistor, but they will not cover the entire width, even if converted to one
long contact. There is, therefore a small additional resistance (about 0.2
squares from the inside edge of the contacts).
The matching of resistors depends entirely on the width. Sub-
micron processes are not developed to get good matching, just maximum
speed. You will find that minimum-dimension devices (all devices, not just
resistors) match very poorly. When greatly magnified under a microscope
all edges appear somewhat ragged. The width of a resistor, for example,
fluctuates considerably. It is only when you make a relatively large device
that these fluctuations become insignificant and thus devices match well.
Figure on using something like ten times the minimum width for matching
of 0.5% or better.
Because of the end-effect you cannot expect resistors of different
lengths to match well. For optimum matching use only identical resistors.
It also helps divide resistors into identical sections and intermingle them
with other resistors (in the same identical sections) which are intended to
match.
One more thing about IC resistors: the Seebeck effect. Discovered
in 1821 by Thomas Seebeck (and used by Ohm four years later for his
measurements of resistance), it is the thermocouple effect: metallic
interfaces at the ends of a wire produce a voltage if the ends are at different
temperatures. For the contacts of a diffused or poly resistors this voltage is
between 0.2mV/oC and 1.4mV/oC, depending on the doping level and
composition of the metal. This is a danger if thermal gradients are present,
e.g. with a power transistor on the chip. To avoid it, lay the resistor out so
that beginning and end are close together.
Pinch resistors (or pinched
resistors) are sometimes used in bipolar
processes to get a high resistance without
wasting a lot of area. The base-pinch
resistor is simply a base resistor with the
emitter diffusion placed over part of it.
This reduces the effective cross-section
(only the deepest part of the base diffusion
is left, which has also the highest
Fig. 1-24: Top view and cross- resistance). The device needs to be in its
section of base-pinch resistor. own epi island, with the epi (and emitter
Capacitors
The oxide insulating the metal interconnection from the silicon (or
between metal layers) is dimensioned to give minimum stray capacitance.
Even a small capacitor (say 5pF) would take up an enormous amount of
space. Enormous at least in microelectronic dimensions.
Thus fabricators often provide an additional mask step to outline an
area where the oxide (or nitride) is thinned down considerably, producing a
higher capacitance (about 2fF/um2 - that's femto-Farads, or 10-15F/um2).
With this figure (which of course varies from process to process) a
50x50um area gives you all of 5pF, easily be the most expensive component
in your chip. If you specify anything greater than 100pF, your colleagues
may think you have a degree in macroeconomics.
One plate of the capacitor is always either metal or poly. For the
second plate you could use a diffusion, but that creates a slight voltage
dependence (there is always a depletion layer in silicon which widens as the
voltage increases, adding to the distance between the plates). Poly or metal
for the second plate are better choices.
The oxide underneath an MOS gate is already thinned down to
achieve a reasonable transconductance, so it too has a higher capacitance
per unit area than the ordinary (field) oxide. But be careful here. At zero
(DC) voltage there is no channel (source and drain form the lower plate, the
gate the upper one), so the only capacitance is the one from the gate to the
overlapping parts of drain and source. When the voltage exceeds the
threshold, the channel comes into existence and the capacitance increases
markedly. Figure 1-26 shown here depicts the behavior of a large
6
(10x20um) 3V n-channel
device.
5
MOS Gate Capacitance There is also junction
vs Gate Voltage
capacitance, which you should
Capacitance / µuF
Other Processes
What we have considered so far are two simple, basic processes,
requiring as few as 8 masks. There are many variations, all based on these
two:
- "Mixed Mode" CMOS, with devices for (somewhat) higher
operating voltages and additional poly (and metal) layers;
- BICMOS processes which add full-fledged bipolar transistors to
CMOS;
- Bipolar processes with vertical (high-speed) PNP transistors;
- CMOS processes with some high-voltage devices (500V).
All of these variations have one factor in common: they increase the
number of masks (and processing steps) required and are thus more
expensive. However, they tend to make the design of high-performance
analog circuits easier, especially when both CMOS and bipolar transistors
are available.
2 Simulation
to judge the quality of the models available to you. Read this part lightly
and then use it for later reference.
DC Analysis
2.5
2
mode range).
1.5
You could enhance this
analysis by repeating it at various
1
temperatures, i.e. by automatically
0.5
"stepping" the temperature, either
0 1 2 3 4
at regular intervals or at three or
Input Voltage/V 1V/div
four points. While you do all this
you can measure the input current
Fig. 2-2: A DC analysis, showing the
common-mode range. (either at the base of Q1 or at
either terminal of Vin), the current
consumption (at one of the terminals of Vcc), the substrate current (out of
the symbol SUB) and even the power dissipation of the entire circuit or any
component.
Place a current source from Out to ground and you can determine
how well the circuit handles a load, i.e. determine the output impedance.
There are two sub-categories in a DC Analysis. The Transfer
Function gives you the relationship between two nodes (not used very
often) and the Sensitivity Analysis tells you which parameters (including
transistor parameters) are most responsible for a change in a particular
voltage or current at any node.
AC Analysis
The one thing you never want to forget about a Spice AC analysis is
this: The signal is treated as if it were insignificantly small. You may
specify a 1-Volt input signal (most people do, it represents zero dB and is
thus very convenient), but the analysis program will process it without
disturbing any of the bias levels. If you have a high gain, say 60dB, the
output plot will show a voltage of 1000 Volts without even blushing. What
it is intended to show you is the gain relative to the input; the actual values
taken out of context are often absurd.
Our plot shows the
0
output response of our buffer in
-1
the most simple AC analysis: a
-2
1-Volt ac signal at the input on
-3 top of a DC Voltage of 2V, with
Gain / dB
Transient Analysis
3.4
2.9
3.2
2.8
3
2.7
Output Voltage / V
2.8
Output Voltage / V
2.6
2.6
2.5
2.4
2.4
2.2
2.3
2
2.2
1.8
2.1
1.6
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
0 20 40 60 80 100 120 140 160 180 200
Time/mSecs 200µSecs/div
Time/nSecs 20nSecs/div
Fig. 2-5: Transient analysis 1: a 1-Volt Fig. 2-6: Transient analysis 2: a 1Vpp
pulse at the input. sine-wave at the input.
Change the input to a sine-wave and you will learn a great deal more
about the circuit. It will be immediately obvious if the circuit can reproduce
the waveform without clipping it at either the high or low excursions. But
that is only a rough impression of fidelity. What you need to know is the
amount of distortion in the waveform.
In some programs you simply display the sine-wave, click on
distortion and get the result. But if you want to have the entire information,
nothing beats a Fourier Analysis.
The Fast Fourier Transform (FFT) is a routine which extracts the
frequency components from a waveform. It is rather tricky to use and
sometimes produces errors. Shown here is the result of a continuous
Fourier analysis (Simetrix),
which is both more detailed 1
values. If you allow the deviation to be three times as large (±3s ) you
enclose 99.73% of all measurements.
This sounds like you are discarding only 0.27% of all values, but the
figure is deceiving. So far we have considered one parameter only, but
there are many in an integrated circuit. Suppose your design is influenced
by 50 of them. The total parameter "yield" then is .997350, or 87.4%. In
other words, you would discard 12.6% of all chips on a wafer.
Unless you simply don't care about cost, you need to design an
analog IC so no chips are lost because of parameter variations, i.e. the
design can withstand a variation of each and every device parameter to at
least 3σ; 4σ would be better.
But how do you find out how much parameter variation your design
can take? The answer is Monte Carlo analysis, and only Monte Carlo
analysis.
There is in use what is called a "four-corner analysis". Device
parameters are bundled together in four groups, representing extremes, or
worst cases. The plain fact is this: it doesn't work for analog circuits. The
four-corner models are just barely able to predict the fastest or slowest
speed of digital ICs, but the grouping simply doesn't apply to analog ones.
In fact, no grouping is possible; a parameter's influence differs from design
to design. Analog designers who are satisfied with a four-corner analysis
simply fool themselves into believing that they have a handle on variations,
when in truth the result is quite meaningless.
A true Monte Carlo analysis varies the device parameters in a
random fashion, so that every combination of variations is covered. This is
also what you get in production.
You don't need to vary every parameter of a device, only the major
ones. For example, varying IS, BF and the capacitances in a bipolar
transistor model is sufficient; the same is true for the threshold voltage, the
transconductance and the capacitances in MOS transistors. If matching is
expected, there must be two additional entries, one for the absolute
variation, and one for the variation between devices on the same chip.
These "tolerances" are either inserted into the model file directly or
contained in a separate file, depending on the analysis program used. The
Monte Carlo program then simply runs the chosen analysis repeatedly, each
time with a different set of variations, randomly chosen. Our example
shows the variation over temperature for a bandgap reference (untrimmed).
How many runs need to be specified for a Monte Carlo analysis?
There is an easy way to find out. Start with 20. Then increase this number
until the extremes no longer change. For this analysis 50 runs were used,
which is more than needed. But with today's fast computers you can afford
to go overboard: the analysis took all of 8 seconds!
This one picture gives
you the variation of a 1.27
Vref / V
1.23
Temperature/Centigrade 20Centigrade/div
know how much variation to
expect until several wafers Fig. 2-9: The result of a Monte Carlo analysis.
Each curve represents the behavior of one circuit
from several different lots (among 50) in production.
have been tested (a single
prototype wafer cannot tell you what the variations in production will be).
Models
The Diode Model
D says the device is a diode. The remaining entries are in Amperes, Ohms
and Farads.
This is about as simple a diode model as you can possibly make it;
just three parameters are specified. IS, together with the series resistance
RS, determine the DC characteristics, CJO the junction capacitance. Let's
first look at the DC behavior.
As we have seen in chapter 1, the current/voltage relationship of an
ideal junction is given by:
Vd⋅q
I1 = Is e k⋅T
where Is = the diffusion current
Vd = the voltage across the current source (i.e. not including RS)
q = the electron charge
k = the Boltzmann constant
T = the temperature in Kelvin
Current Gain: The main parameter here is BF (the forward beta, or hFE)
and its temperature coefficient XTB. Without any additional parameter the
current gain would be the same at any collector voltage or current. The
Early effect is represented by VAF (the Early voltage, see chapter 1). The
drop-off at high current is produced by IKF (the current at which hFE starts
to drop) and NK (the steepness of the drop). ISE and NE of the base-
emitter diode are responsible for the drop in hFE at the low-current end;
simply shunting a small amount of base current to the emitter.
Reverse Current Gain: You may be convinced that you will never operate
a transistor with the collector and emitter interchanged, but just in case
provided the parameters BR, NR, VAR, IKR and TR .
The Spice model for an integrated bipolar transistor can have either
three or four terminals. The fourth terminal (of an NPN transistor) is the
substrate and between it and the collector there is a diode, represented by
the five parameters ISS, NS, CJS, VJS and MJS. This is a major flaw in
Spice, for a mere diode here is inadequate. When the transistor saturates, a
substantial portion of the total current flows to the substrate, which this
model simply ignores.
Fortunately Spice also contains a solution to this problem. To
represent an NPN transistor correctly, you need to add a second transistor;
Spice lets you combine the two (or any number of devices) in a subcircuit.
For a lateral PNP transistor the Spice bipolar transistor model alone
is woefully inadequate. This type of transistor 3 (Emitter)
Once upon a time there was a company which brought out its own
variation of the Berkeley Spice program: HSPICE. It specialized in making
refined models for MOS transistors, many of them. The models were called
levels and many companies bought their own levels, like boxes at the opera.
AMD had three of them, Siemens acquired two. Motorola, National
Semiconductor, Sharp, Cypress, Siliconix and a few others only got one.
By 1995 there were 39 such levels and us poor ordinary folks couldn't
access most of them: they could only be used by the company who had
sponsored them.
At last good old Berkeley came to the rescue. A team of researchers
developed the BSIM model (Berkeley Short-channel IGFET Model). The
team stayed with it, through BSIM1, BSIM2, BSIM3 and even BSIM4.
These models divided the MOS transistors in ever finer structures, tracking
the trend toward geometries far below 1um. As of this writing BSIM3.3 is
the dominant model in the industry, leaving the many HSPICE levels in the
dust.
Naturally HSPICE took the BSIM models and made its own version,
adding more levels.
The increasing BSIM refinements have its toll: the number of
parameters has become very large, so large that it takes an entire book to
explain them. For digital ICs, which require utmost speed, this simply has
to be accepted. For analog designs, which invariably use larger dimensions
to obtain adequate performance (especially for matching), it is a burden
only grudgingly tolerated. MOS model-making has become an art
dominated by the digital realm, of limited use to the analog designer.
In a modern BSIM model you are confronted by a mass of data
which almost always is presented in an arbitrary way, lacking an
organization which would make it more understandable. To help in a minor
way, they are grouped here; the bold-faced parameters are absolute values;
all others are modifiers. Parameters in square brackets are temperature
coefficients.
Threshold Voltage: VTHO, K1, [KT1, KT1L], K2, [KT2], K3, K3B,
DVT0, DVT0W, DVT1, DVT1W, DVT2, DVT2W, VBM, VOFF, KETA,
PSCBE1, PSCBE2.
Saturation: VSAT, [AT], A0, AGS, A1, A2, B0, B1, DELTA, EM, PCLM,
PDIBLC1, PDIBLC2, PDIBLCB, DROUT, PVAG, AGS, ALPHA0,
BETA0.
Geometry: W0, DWB, DWG, LL, LLN, LW, LWL, LWN, WL, WLN,
WW, WWL, WWN.
Process Parameters: TOX, XJ, XT, NCH (PCH), NGATE, NLX, NSUB,
GAMMA1, GAMMA2, JS, [XTI], NJ, JSSW.
BSIM models also allow "binning": several models are written for
different geometries of the same device, and then selected to fit into a range
of gate width and length with the parameters LMIN, LMAX, WMIN and
WMAX. While this is not really necessary for the parameters listed above
(some foundries, notably AMS, manage to create equally accurate model
without binning), the Monte Carlo variations should be tied to channel
width and length (i.e. area). Note that the multiplier M is used for
transistors with a channel width beyond WMAX.
To get into more detail on the many parameters, you will need to
consult the original Berkeley documentation (see references). Be
forewarned: this is a lengthy document.
Resistor Models
0.6
V (1) + V ( 2 )
I1 = I (V 1) ⋅ 0.0033 ⋅ V ( 3) −
2
where 0.0033 and 0.6 determine the amount and shape of voltage
dependence. Note that the bias voltage is applied from terminal 3 to the
dB
-25
section lumped model is
remarkably accurate. -30
3 Sections
Frequency / Hertz
There are only two cases where a simple capacitor model (i.e. an
ideal capacitor) is inadequate:
1. There is a requirement for unusual precision. If one plate of an
oxide capacitor is a diffused layer (or a poly layer with a high sheet
resistance) the capacitance will decrease slightly as the potential across the
plates is increased. A competent model will reflect this non-linearity.
2. The capacitor is used at the high-frequency end. Here it is not of
great importance for the model to show the non-linearity, but to reflect any
series resistance and stray capacitances from both the lower and the upper
plate to neighboring regions.
3 Current Mirrors
Bob Widlar was a truly great designer of analog ICs. He was wild
and totally unmanageable and had an odd sense of humor. The press loved
him and he had a flair for self-promotion. He shunned computer analysis,
preferring to breadboard his circuits, but time and time again he came up
with nuggets of design details and products which were thought to be
impossible. Burned out by the frenzy of Silicon Valley he moved to
Mexico, where he died in 1991 at age 53.
One of Widlar's early contribution is the +V
current mirror, a design detail (or design element)
I2
which you will now find in just about any analog IC. I1
Start with the primary current, I1, which flows 50u
into the diode-connected transistor Q1. This produces
a voltage drop across Q1, namely that of its base-
Q1 Q2
emitter diode; this voltage drop is called a VBE.
Now connect the base and emitter of a second,
SUB
identical transistor, Q2, to the same nodes as those of
Q1. Since the base-emitter voltage of Q2 is the same
Fig. 3-1: The Widlar
as that of Q1, it follows that its collector current current mirror.
should be the same as that of Q1 and, therefore I2=I1.
Well, not so fast. There are errors, two of them. The first one
concerns the base currents. I1 splits into
54
three paths: the collector current of Q1
and the two base currents. Assuming a
52
minimum current gain of 100, each base
current amounts to 1% of the collector
I2 / uA
50
current, for a total of 2%. So the collector
48
currents of Q1 and Q2 are 2% smaller
than I1, worst-case.
46
The two transistors may be
0 1 2
identical, but they are not necessarily
3 4 5
All current mirrors start with a current source, from which one or
more currents are derived. For ICs, a current mirror is a more basic
element than a current source, which is the reason they are discussed
first.
However, be aware that there is a significant difference between a
theoretical current source (as in a simulation) and a practical one. In a
simulation a current source will do anything to keep its programmed
current level, including building up thousands of volts. In an actual
circuit the supply voltage limits the excursion.
Also, little distinction is usually made between a current source
and a current sink (e.g. I1 in figure 3-3) . For convenience all of them
are usually termed current sources.
we have seen in chapter 1, the gain is affected by the collector voltage (the
Early effect), increasing as the collector voltage is increased. Thus I2 is not
exactly steady. For this particular transistor (made in a process capable of
20 Volts) the change amounts to 8% from 0.3V (the saturation voltage of
Q2) to 5 Volts. (I1 is 50uA for all examples in this chapter).
This is the most simple current mirror and, as we shall see in figures
3-7 and 3-9, we can improve its performance considerably with additional
devices. There is also a lateral-PNP equivalent. Using a
+V
split collector (see figure 1-17), this current mirror needs
only a single device. Each collector being smaller, the
Q1
maximum current is more limited (depending on the
process, about 100uA).
I1
I2 The voltage 60
50u dependence of a PNP 58
current mirror is 56
54
SUB generally a bit worse 52
than that of an NPN
I2 / µA
50
Fig. 3-3: Current
mirror with
design (here about 48
44
voltage of the second 42
PNP collector can move to within 0 1 2 3 4 5
about 0.3 Volts of +V. If you let it go Voltage Collector 2 of Q1/V 1V/div
from 1 to 3 Volts 50
W=5u W=5u
L=2u L=2u (0.35u process), 49
I2 / µA
but only because 48
emitter resistors. 49
There 48.8
is a penalty: The voltage at the 48.6
collector of Q2 cannot go any 48.4
SUB
does this job and 49.5
coincidence the
I2 / µA
50.6
I1 collector voltage of Q1 to
50.4
the same level as that of
I(Q3-C) / µA
50.2
Q2. With this I2 is now 50
Q4 Q3 within 0.6% of I1 and 49.8
changes by less than 49.6
Q1 Q2
0.08% with voltage. 49.4
Fig. 3-14: Although the output current now between input and output current.
changes little with voltage, there is still If the critical transistor on the
considerable variation due to mismatch, as
a Monte Carlo analysis will show.
output side is increased in size, its
collector current is increased too.
+V +V In a bipolar transistor the current ratio is
I2=3xI1 I2=I1/3 determined by the size of the emitter
I1 I1 (more precisely, the active emitter length;
see chapter 1) but a accurate ratio is in
practice only achieved if you work with a
Q1 Q2 Q1 Q2 number of identical emitters. In figure 3-
15 Q1 has one emitter while Q2 has three
SUB SUB (they can all be in the same base),
Fig. 3-15: 1:3 resulting in a current which is three times
Fig. 3-16: 3:1
current ratio. current ratio.
that of I1. In figure 3-16 Q1 has three
emitters and Q1 one, which causes I2 to have one-third the value of I1. Any
ratio is possible (such as 3:2 or 5:3). In a CMOS design the ratio can be
SUB
number of separate transistors. In Fig. 3-18 emitter
resistors are used to get less of a change in I2 with a
Fig. 3-18: an varying output voltage (0.7% from 0.7V to 5V); if you
additional have 10 separate transistors, they all get 6kOhm in the
transistor supplies
the base current. emitter; if the current is simply multiplied by 10, R2 has
one-tenth the value of R1. Remember that best matching
is achieved if the resistors consist of identical sections, i.e. you create a
basic 600 Ohm device and use one for Q2 and 10 in series for Q1.
If you are thinking of turning I1 on and off rapidly, be aware that
this circuit is very slow to turn off; there is no discharge path for the bases
of Q1 and Q2. A resistor (or another current sink) from these bases to
ground helps to speed up the turn-off +V
time.
The base current problem does I1
49.9
50.4
49.85 50.2
49.8 50
49.8
49.75
49.6
I2 / µA
I2 / µA
49.7
49.4
49.65 49.2
49.6 49
48.8
49.55
48.6
0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5
Fig. 3-21: Performance of cascode Fig. 3-22: Figure 9-21 repeated with Monte
MOS current mirror. Carlo variations.
performance, but requires at least 0.7 Volts at the output. Lowering the
voltage drop across R1 lowers this minimum output voltage, but increases
the voltage dependence, which you can reduce again by using even larger
devices.
Again, don't get carried away by the impressive performance shown
with a single sweep, which assumes perfect matching. A Monte Carlo run
will show you the true behavior.
For CMOS Current mirrors there are three more sophisticated
schemes. Figure 3-23 is the one you frequently see in articles. M1 is a thin
device, producing a bias voltage about 100 to 200mV higher than the gate
voltages of M3 and M5. Since the gates of M2 and M4 are connected to
this point, these two devices act as cascodes, i.e. they shield the lower two
devices from voltage changes.
M2 M4
important for high 50.03
3-23
49.99
Performance. are for room
temperature only. 49.98
Make sure you simulate your 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8
Open any analog IC and you will find a differential pair. Or, more
likely, a half dozen. It has great advantages, even if amplifying a
"difference" is not even a goal.
The reasoning is simple: Individual integrated components have
large variations, but two (or more) of the same match very well. If you can
take advantage of the matching, you get better performance.
It isn't always true, of course. Noise, for example can be smaller in
a single-transistor stage and some of the most ingenious designs are
remarkably free of the common differential stage. But let's look at this
wondrous tool.
Two transistors - here bipolar - share a
common emitter current. If the voltages at
I2 I3 their bases are equal and the two transistors
Vcc
match perfectly, I1 is split into two equal
Q1 Q2 5 parts at the collectors, I2 and I3.
If we increase V1 (relative to V2), Q1
V1 V2
gets more of the current than Q2. If we
I1
100u 2
SUB decrease V1, the opposite is true.
But there are limitations and errors.
First of all, the current division (or input
Fig. 4-1: In a differential pair voltage to output current relationship) is not
a current is divided by two linear. We are dealing with two base-emitter
transistors.
diodes here, fundamentally exponential
devices. Not counting stray effects the emitter resistance is:
k∗T
re =
q∗ Ie
1
gm =
re + Re
µA
50
10
get a rather non-linear behavior.
1.8 1.85 1.9 1.95 2 2.05 2.1 2.15 2.2
Only a small portion of the curve V1/V 50mV/div
in the middle, when the two current
Fig. 4-2: The conversion from input
are equal or nearly equal, could be voltage to current (the transconductance)
called linear, though in truth it too is non-linear.....
is not a straight line.
The other variable in the
equation is temperature. The emitter
90
80
resistance is proportional to absolute
70
-50C 125C
temperature, so at high temperature
60 you get less gain or transconductance.
I2 / µA
10
from the bases, not the collectors.
1.8 1.85 1.9 1.95 2 2.05 2.1 2.15 2.2
With a minimum hFE of 100 this
Input Voltage/V 50mV/div
makes the sum of the collector current
smaller than the emitter current by
Fig. 4-3: ...and also temperature-
dependent. 1%, 2. Transistors only match well if
the are treated identically. In this case
that specifically means the collector
voltages have to be the same. If one is higher than the other, its transistor
will have a higher gain because of the Early effect; 3. Devices never match
perfectly; there will be some differences in both VBE and hFE and thus
some uncertainty in the voltage at which I2 and I3 are equal, showing up as
an offset voltage.
18 18
-50C
16 16
I3 I2
14 14
125C
I(M1-drain) / µA
12 12
µA
10 10
8 8
6 6
4 4
2 2
Fig. 4-5: A CMOS differential pair is Fig. 4-6: ...and the transconductance
also non-linear ..... also has a temperature coefficient.
take up more space than all the other Fig. 4-11: Output signal with a
components together. Second, if you 1mVp sine-wave input.
look at the output waveform closely, you
notice that there is a DC current flowing through R1. At the right end we
connect it to a 3-Volt bias point, but at the left end the center of the sine-
wave is not 3 but 3.25 Volts, i.e. we have a built-in offset. There are two
reasons for this: 1. the two collectors of Q3 are not at the same potential and
2. the collector current of Q1 has to supply the base current for Q3.
We need something like R1 to fix the DC potential at the output.
The two opposing collectors are current sources/sinks. The smallest
difference between the two would cause the output potential to move up so
much that Q3 would saturate or down so much that Q2 would saturate.
In figure 4-12 all of these problems are fixed at once by adding a
second stage. Q4 is the same size as Q3 and is operated at the same current.
Now the collector voltages of both Q1/Q2 and Q3 are identical. Moreover,
the collector current of Q2 has to supply the same amount of base current
(for Q4) as the collector of Q1 does. In other words, with no input signal
the circuit is perfectly balanced; there is no built-in offset.
But you need to be careful here.
Vcc
Q3 The gain of this circuit is no longer fixed
Q4
5 by a resistor ratio, it is dependent on
transistor parameters. If these two stages
Out
Q1 Q2 are made part of an operational amplifier,
the feedback will take care of this. Or, if
Vin1 I1 Vin2 I2
100u 100u
the circuit is used merely as a comparator,
2 SUB
gain is of lesser importance than offset.
5 Current Sources
Ever since the dawn of analog IC design (all the way back in 1962)
a succession of very clever people have been trying to conjure up
something that would produce an accurate current. The results have been
uniformly dismal.
There happens to be a capable voltage source in ICs, the bandgap
reference (which we shall get into next). So, to get a current, one would
think, all one needs is an accurate resistor; after all I = V/R. But, unless you
want to add a costly thin-film layer and laser trimming, there are no
accurate resistors. What we get are resistors made from diffused or
deposited silicon layers which vary in resistance from wafer to wafer and
have a considerable temperature coefficient.
So, don't expect any precision here. At best, an integrated current
source can provide a small current without the use of large-value resistors
and make this current more or less independent of the applied voltages.
16
the voltage at the SUB
14 collector of Q1
12
(the destination of Fig. 5-1: Current
I1 / µA
10 source based on
8
the current) drop VBE (Q3)
6 below about 1
4
Volt, Q1 saturates
2
00
and draws from the primary current.
1 2 3 4 5
I1 / µA
10
6
dependent on the supply voltage (a 4
+2% increase as the voltage moves 2
18 16
16 14
14 12
12
10
I1 / µA
I1 / µA
10
8
8
6
6
4
4
2
2
00 1 2 3 4 0 1 2 3 4 5
Fig. 5-5: I1 vs. output voltage of figure 5-4. Fig. 5-6: I1 vs. supply voltage of figure 5-4
12 12
I1 / µA
I1 / µA
10 10
8 8
6 6
4 4
2 2
00 1 2 3 4
0 1 2 3 4 5
Fig. 5-8: I1 vs. output voltage of figure 5-7. Fig. 5-9: I1 vs. supply voltage of figure 5-7.
8
simulation. Just try various values
6
for R1 and R3 until you get the right 4
current with minimal change. But , 2
We are about to take a rather daring step. The primary current in the
previous circuits is a nuisance; it wastes power and takes up considerable
resistance. Why not replace it with a current source derived from the
current the circuit generates?
There is one flaw in this argument: the current must exist first.
There are two possible modes, one in which the current levels are as
intended and one where there are no currents at all. In other words, there
must be a current in Q2, which can be mirrored and fed back to Q1 and the
base of Q2 so Q2 can have a current, etc.
12
bases of either Q1 or Q2 to 10
ground which can shunt 8
impedance is 50MOhm. 8
roughly that of R2, here assumed 00 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Voltage/V 200mV/div
to be ± 25%.
Fig. 5-15: Performance of the Erdi current
source.
CMOS Current Sources
None of the bipolar schemes work well for CMOS devices. They
are based on VBE or delta-VBE , for which there are no equivalents.
Trying to use circuits such as Figures 5-10 or 5-14 with CMOS width-ratios
I1 leads to inferior circuits. Also, due to the square-law
behavior of the gate voltage, the variations are roughly
M1
double those of bipolar designs.
Fortunately the CMOS transistor is a current
Vref
W=4.5u source. If we simply apply a constant voltage to the
L=20u
1.2 gate (such as a reference voltage) we can tailor the
width and length of the device to give us a certain
current.
Fig. 5-16: MOS Using a rather exaggerated length, we can
transistor as a minimize the channel-shortening effect. In the example
current source.
I1 / µA
threshold voltage) is large: ± 39%. 4
1
variation in the reference voltage (± 00 0.5 1 1.5 2 2.5 3 3.5
3% causes a change of 10% in I1). Output Voltage/V 500mV/div
dB
Analog scales tend to be very large. As an example, the hearing
threshold of a young adult is 20 micro-Pascal; the maximum level without
damaging the ear can be more than 20 Pascal, a ratio of one to 1 million.
Particularly because of the widely varying sound levels a need for a
logarithmic measure appeared early on in electronics. There are two of
them: The Neper is based on the natural logarithm and named after John
Napier, a 16th century Scottish mathematician who came up with the
logarithmic table (and whose name was most likely spelled Neper in his
time).
In the 1920's a measure based on logarithm with base 10 began to be
used at Bell Laboratories. At first it was called the "transmission unit", then
re-christened the Bel, after Alexander Graham Bell. The idea is simple, a
Bel is the logarithm of two power levels:
P1
Bel = log
P2
But the Bel turned out to be a bit coarse; one-tenth of that suited the
Bell Labs people better, hence the decibel, or dB:
P1
dB = 10∗ log
P2
This is the ratio of two power levels. Since power is related to the
square of the voltage (or current), we get:
V1
dB = 20∗ log
V2
difference between the ratios of power and those of voltages or currents (or
pressure).
Even though a logarithmic ratio is very convenient, it is helpful to
picture the (voltage) actual ratios:
-60dB 1/1000
-40dB 1/100
-20dB 1/10
-6dB 0.5 (exactly: 0.5012)
-3dB 0.707
0dB 1
20dB 10
40dB 100
60dB 1000
dBm Power ratio where 0dB = 1mW. Originally this was based
on an impedance of 600 Ohms (that of a telephone line),
but now is used with any impedance (which is fine as long
as we calculate power ratios, not voltage ratios).
RMS
RMS calculation was introduced by Charles Steinmetz more than
100 years ago. Steinmetz grew up in Breslau, Germany (now Poland), but
shortly before he got his Ph.D. in mathematics and physics he had to flee to
Switzerland because of his socialist activities. From there he emigrated to
the U.S. and found a job as an assistant draftsman in Yonkers. The company
fabricated hat-making machinery, but soon expanded into electrical motors.
The year was 1889.
Steinmetz was a small man with a hunchback and one leg shorter
than the other, a deformity he inherited from his father. Though he made the
proverbial bad first impression, the people around him soon were in awe of
his razor-sharp mind. No surprise then that the overqualified draftsman was
Vpeak
Vrms = = 0.707∗Vpeak
2
Voltage / V
finally at zero again.
2
52 + 0 2 + 2 2 + 0 2 1
Vrms = = 2.69V
4 00 50 100 150 200 250 300 350
Time/µSecs 50µSecs/div
But the RMS calculation has Fig. 6-1: Arbitrary waveform for RMS
some limitations: it doesn't work with calculation.
non-linear elements. In Steinmetz's
time there were no transistors, not even vacuum tubes. There were only
linear elements (save perhaps for the occasionally saturating transformer),
so he didn't consider what would happen if the impedance changes while
you are measuring RMS voltage and current.
Take the case of a transistor stage, either linear or switching. You
want to determine its power dissipation, so you measure the current through
it and the voltage across it. But the impedance of the transistor constantly
changes and Ohm's law doesn't hold. The product of RMS voltage and RMS
current gives an absurdly wrong result for power. The only way you can
determine the power is to integrate the instantaneous values of voltage times
current. In a simulation, Spice does this very well.
But measurements aren't nearly so easy: "True RMS" instruments do
indeed have a circuit element which measures RMS rather than average. But
the inputs to almost all of them are capacitively coupled. If the waveform
you are measuring has a DC component, it is ignored and the result reflect
on the AC portion of it.
Noise
Imagine a current flowing through a wire connected between a
negative terminal on the left and a positive terminal on the right. Through
the wire are flowing millions of electrons from left to right.
Each electron carries a charge of 1.6e-19 Coulombs. Let's say we
observe a current of 1uA, thus 7.8e12 electrons pass every second. If the
interval between electron were the same, we would then see a ripple at
7800Ghz, like the teeth of a saw-blade moving at high speed.
Inoise( rms ) = 2 ⋅ q ⋅ I ⋅ B
There are two things you should notice here. First: Noise increases
as the square-root of bandwidth. Second: When the current is decreased,
noise becomes a larger fraction of it.
Let's illustrate the second part. With a bandwidth of 10kHz, 1mA dc
produces 1.8nA(rms) of noise. That amounts to 0.00017% or -115dB.
With 1uA of current and the same bandwidth the noise is 56pA(rms), i.e.
0.0056% or -85dB. At 1nA we get 1.8pA of noise, which amounts to
0.18% or -54dB. All of which shows that it is harder to design a low-noise
circuit at low current levels.
There is also noise when no current flows at all. By the energy
imparted by temperature, some electrons will suddenly leave an orbit and
jump to another. The higher the temperature, the larger this irregularity
becomes. Thus a resistor, doing nothing but lying on a bench actually has a
noise voltage at its terminals:
Vnoise( rms ) = 4 ⋅ k ⋅ T ⋅ R ⋅ B
0
which can be extracted in a Fourier series.
-200 There is an algorithm called "Fast
-400
Fourier Transform", or FFT, which does
this. FFT uses an algorithm which allows
0 20 40
fewer computations compared to the
60 80 100 120 140 160 180
Amplitude / V
etc., with gradually decreasing 1m
Frequency/kHertz 5kHertz/div
disregard harmonics after the
fourth or fifth, since their Fig. 6-3: Fast Fourier transform with low
resolution.
amplitudes become very small:
Fundamental = 550mV
42.4
Distortion = = 0.077 = 7.7%
550
The peak at zero frequency shows the DC level, i.e. the asymmetry
caused by the clipping.
Before you run a fast Fourier transform in Spice you need to choose
two settings: how many samples should be taken over one period of the
waveform and how many periods should be analyzed. In the example of
Figure 6-3 there are in fact too few samples and periods, resulting in broad
peaks.
As a general rule start with 25 samples and 50 periods. The first is
determined by "maximum time-step" and "maximum print step" (set at
8usec in figure 6-4 for a 5kHz driving frequency) and the second by the
total time in the transient analysis (1msec for 50 periods at 5kHz). You get
the best results if both the number of samples per period and the number of
periods are integers.
The fast Fourier transform has some flaws and limitations. For
example, figure 6-4 shows peaks in between the harmonics, which, in
1 1
100m
100m
10m
10m
1m
Amplitude / V
Amplitude / V
100µ
1m
10µ
100µ
1µ
10µ 100n
10n
1µ0 5 10 15 20 25
1n 0 5 10 15 20 25
Frequency/kHertz 5kHertz/div
Frequency/kHertz 5kHertz/div
Fig. 6-4: Fast Fourier transform showing Fig. 6-5: Continuous Fourier transform
false peaks because of still insufficient with high resolution.
resolution.
reality are not there. A superior method is the Continuous Fourier
Transform, available in some analysis programs and shown in figure 6-5.
When you have a waveform which is symmetrical but not a sine-
1 1
0.8
400m
0.6
200m
0.4
100m
Amplitude / V
Amplitude / V
0.2
40m
-0
20m
-0.2
10m
-0.4
-0.6 4m
-0.8 2m
-10 1m 1 2 3 4 5 6 7 8 9 10
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Frequency Compensation
Feedback is a wonderful thing. We take the inverted output signal,
subtract the input signal from it and the amplifier will automatically correct
and difference between them. If we only feed back a fraction of the output
signal, the amplifier will automatically adjust its gain to one over that
fraction.
Using a single frequency (any frequency), inverting a signal
(negative feedback) is the same as a 180 degree phase-shift. And here
comes the problem: Each device in the amplifier has a little bit of delay. At
low frequency this has little effect, but as we go higher and higher in
frequency the delay becomes more and more noticeable. At some high
frequency the delay amounts to half a period of the signal and thus causes a
phase-shift of 180 degrees. What started out as negative feedback now
becomes positive feedback and the whole thing oscillates.
Frequency compensation is a design method which avoids this. The
principle is very simple: deliberately slow down one device so that it is
much slower than all others, i.e. it dominates the frequency response so that
the delay in all other devices is no longer important.
Feedback This is illustrated with a very
E1 simple simulation. E1 is a "voltage-
R1
In
100k
Out controlled voltage source" and acts like
C1
an ideal op-amp with a gain of 1 million
1Meg
Vin
10n (120dB), has no delay and the input and
output terminals are free-floating (but
are referenced here to ground). R1 and
Fig. 6-8: Abstract circuit to illustrate
phase-shift in a feedback amplifier.
Y2 Y1
Gain / dB
100 100
decreasing at about 100Hz. At Pole
80 80
this point the phase of the signal 60 60
at the output is considerably less 40 40
than 180 degrees, but as we go 20 20
oscillation.
The point at which the phase has turned by 45 degrees is called a
pole. At frequencies somewhat higher than the pole the amplitude drops by
6dB per octave (doubling of frequency) or 20dB per decade.
Now let's look at the same
simulation with another RC Y2 Y1
160 160
pole at a much higher frequency (C 140
Pole 1
140
Phase / deg
100 100
Gain / dB
second pole (at about 10kHz) by 80 80
60 60
the change in the steepness of the 40 40
point of interest is the frequency at -20 -20 1 10 100 1k 10k 100k 1M 10M
dB (i.e. a gain of 1). If the gain is Fig. 6-10: Two poles in a feedback path
less than 1 an oscillation cannot approach zero degrees phase-shift.
sustain itself. While the phase at
this point only approaches zero degrees, the margin is far too close for
comfort.
Y2 Y1 With three poles we are
180 180
160 160
Phase clearly out of luck. The phase
140 140 now reaches zero degrees a
120 120
Gain
decade before the gain drops
Phase / deg
100 100
below 0 dB. An amplifier which
Gain / dB
Oscillation
80 80
60 60
has these three poles will
40 40 oscillate, in fact we can tell with
20 20 certainty that it will oscillate at
0 0
200kHz.
-20 -20 1 10 100 1k 10k 100k 1M 10M
There are now three
Frequency / Hertz
remedies: 1. we can lower the
Fig. 6-11: Three poles in a feedback path. gain until it drops below 0dB
Phase-shift goes through zero degrees and before the phase reaches zero
oscillation takes place.
degrees; 2. we can insert a new
pole at a frequency so low that it
dominates the others and 3. you can introduce a zero.
To illustrate the effect of a zero, we use another artificial circuit.
R1/C1, R2/C2 and R3/C3 provide the three poles, delaying the phase of the
signal, each by the same amount as in figure 6-11. R4, together with C2
provides the zero, it advances the phase rather than retarding it. The
10n
1k
C2
100p 1k
C3
frequency where it is most
V1 Zero R4
1p
effective.
20k
At about 30kHz R4/C2
Fig. 6-12: Three poles and a zero. start turning back the phase, so
that at the critical frequency
(5MHz) the gain drops
below 0dB but the phase is Y2 Y1
100 100
Gain / dB
behavior will not oscillate, Zero
80 80
though the phase margin is 60 60
rather low. Since gain and 40 40
to variation in an IC, it 0 0
1Meg
gain, and an output stage which
1
has no (voltage) gain but provides
V1
a reasonably high output current.
Since high-current PNP transistors
Fig. 6-14: Measuring gain and phase in a
are often not available in an IC,
feedback loop. the lower portion of the output
stage uses a compound transistor;
from the second stage it looks like a PNP transistor, from the output like an
NPN one (but the combined device is achingly slow).
Q5 and Q7 are diode-connected transistors to bias Q6 and Q8.
The amplifier is investigated as a buffer, i.e. with a gain of one,
produced by connecting the output directly to the inverting input. Here,
though, there is an inductor in the path, which blocks AC but lets DC
through so that the circuit is properly biased. C2, a very large capacitor,
couples an AC signal to the negative input. In this way the feedback loop is
opened up and we can measure loop gain and phase. This can be done at
any convenient point in the loop, but the output to input connection is
clearly the most convenient. Note that L1 and C2 have impractically large
values. This is of no great consequence since these components are not
going to be part on the design; we want to make sure they don't influence
the AC behavior of the circuit.
We feed the AC signal into the loop after the inductor and then
measure the loop response before the inductor (at "Out").
First let's look at the loop without C1. The loop gain is about 92dB
and the phase drops rather sharply, reaching zero degrees long before the
gain reaches 0dB. (Gain
and phase have identical Y2 Y1
100 100
Therefore this circuit is 80 80
Gain
C1, the 40 40
compensation capacitor, 20 20
The result is self-evident. A new pole is created, about 100 times lower in
frequency than the next higher one. This pole now dominates up to at least
10MHz and the phase is still 65 degrees away from zero when the gain
drops below one. A stable circuit with an adequate safety margin.
Of course there is a
Y2 Y1
price to be paid for this
stability: the gain of the op-
160 160
Phase
amp may be more than
140 140 90dB at 10Hz, but it drops
Phase Margin
120 120
65 degrees steadily as the operating
100 100 frequency is increased. If
Phase / deg
Gain / dB
80 80
Gain we use this op-amp at
60 60 10kHz, we only have about
40 40
58dB of gain.
20 20
This analysis has
0 0
assumed that the op-amp is
-20 10 100 1k 10k 100k 1M 10M 100M going to be used with a
Frequency / Hertz gain of one. But if you are
Fig. 6-16: With C1 the circuit of figure 6-14 has a
creating a design with a
phase margin of 65 degrees, i.e. the gain drops fixed gain, say 40dB, there
through 0dB safely before the phase reaches 0. is no reason why it should
have to be stable at a gain
of one. Which makes frequency compensation much less demanding. Just
look at figure 6-15. Subtract 40dB from the gain curve (only the excess
gain counts) and the amplifier is almost stable, i.e. a much smaller
compensation capacitor is required.
The gain/phase analysis, as elegant and informative as it is, has a
serious flaw: it shows performance only at one particular operating point (it
is, after all, an AC analysis which does not disturb DC operating voltage
and currents). A real-life signal will change the DC operating point and the
loop gain and phase can change substantially.
Some simulators let you perform this AC analysis at different DC
operating points, but there is an easier way , one that is a surefire test for
stability. Get rid of the inductor and C2, close the feedback loop as
intended in the application and apply a square-wave at the input. The
square-wave should have fast edges (the default values in the simulator are
adequate).
Then observe the output and watch for overshoot. For this circuit,
with C1 at 20pF, there is a slight overshoot, one peak only. This circuit is
very stable. (You can also see that the large compensation capacitor affects
the slew-rate rather badly).
Output Voltage / mV
200
safe.
-600
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
7 Bandgap References
This is not strictly a straight line; it is slightly convex below about 150oC
and concave above (it asymptotically approaches zero volts). The bandgap
voltage at zero K, by the way, is strictly a theoretical concept; at that
temperature there are no semiconductors, in fact electrons don't move at all.
Widlar found that an equal but opposite temperature coefficient can
be created by running transistors at different current densities:
k∗T A1∗ I 2
deltaVBE = ∗ ln
q A2∗ I1
where A is the area (effective emitter area) of each transistor and I the
current running through it. Here you have the choice of either using
different emitter sizes, different current levels or both at the same time.
Delta-VBE is a true straight line, pointing to zero at zero K. But it is
relatively small. kT/q amounts to about 26mV at room temperature, so an
area (or current) ratio of 10 gives you a delta-VBE of about 60mV. As you
can see from the diagram in figure 7-2 you need about 600mV at room
temperature so it counteracts VBE.
But Widlar came up with a simple Vcc
Q5
10
14 transistors, so it wasn't so simple after all).
R1
1.5k The Brokaw cell needs a start-up
Q6
circuit, which has been added here (Q7 lifts
R2
6.9k Vref to one VBE, which is sufficient for Q1
SUB and Q2 to start drawing current).
Q2 has 10 times as many emitters as
Fig. 7-4: The Brokaw cell.
Q1 (an arbitrary choice, more would be better),
so there is a delta-VBE of 60mV (at room temperature) across R1. Q3 is a
current mirror, forcing Q1 and Q2 to run at the same current. Q4 completes
the feedback loop from the collector of Q1 back to the input bias of the
differential pair Q1/Q2 and supplies a moderate amount of output current.
1.2338
1.2336
to about 0.18%.
1.2334
1.2332
This curve was obtained
1.233 using models for a simple 5-Volt
1.2328
1.2326
0 20 40
bipolar process. The results are
60 80
Temperature/Centigrade
going to be different for other
20Centigrade/div
1.236
-30
1.234
-35
1.232
1.23 -40
dbV @ Vref / dB
1.228
-45
1.226
0 20 40 60 80 100 120 140 160 180
-50
Time/µSecs 20µSecs/div
-55
Q3
R1 Q4
24k
It's Widlar's turn again. Four years C1
at room temperature. Since there is only one current flowing through all
three resistors (save the base currents of Q1 and Q2) the voltage drop across
all of them is 36mV(53k/3k) or 636mV.
1.2
Add the 600mV VBE of the diode-
1.1
connected Q6 to this and you get a
0C temperature compensated reference
1
voltage of 1.236V. (Again, this value
Vref / V
0.9
100C and the required values for R1 or R3
0.8
may be somewhat different for other
0.7
processes).
0.6
The multiplying resistor as been
0 10 split into two parts (R1 and R3) to
20 30 40 50
I1/µA
provide enough headroom for the
10µA/div
R1
VBEs in series, so the output voltage is twice the
10p
40.5k bandgap potential, 2.45V. With R3 = 25kOhm the
optimum Vcc range is 4.5 to 5.5V.
Q4a
R2
3k
The delta-VBE appears across R2, given by
Q4b
Q3 the 24:1 emitter ratio of Q3 to Q1 (about 83mV at
24
room temperature) and is multiplied by R1 to about
Q1
SUB
1.2 Volts. The difference here is the placement of
R2 in the collector circuit of Q1, thus subtracting
rather than adding the delta-VBE. One VBE is that
Fig. 7-17: A different of the lateral (split-collector) PNP transistor Q2, the
design for a bandgap other the NPN transistor Q1. Lateral PNP
reference (2.5V).
transistors generally have a narrower variation in
VBE, but work only over a limited current range.
The error signal is picked up by a Darlington transistor (Q4, one
collector region, two base-emitter patterns).
Variation in production over a temperature range of 0 to 100oC is a
mere ± 1.6%. As always, the values given here are for a specific process,
with fairly large dimensions (the resistors are 4um wide). You may need to
adjust R1 for other processes (and certainly for other emitter ratios).
The circuit is stable with a load capacitance of less than 50pF or
greater than 200nF. With a 330nF capacitor at Vref power supply rejection
is -60dB, increasing further above 10kHz. The output impedance is 25
Ohms. The circuit is intended as a reference only, but it can sink several
milliamperes. If more sourcing current is needed, you simply decrease the
value of R3.
It is possible to modify this circuit for 1.2 Volts but, as a
consequence the performance suffers a bit (which is almost always true
when you move to lower voltages).
In figure 7-18 only a single diode-connected transistor (Q1) is used.
A second one mirrors one-third of the current, which is compared with the
mirrored current of Q4. Here the emitter ratio is 20:3. A second stage (Q5)
increases the loop gain, lowering the output impedance to about 1.7 Ohms.
R3 is optimized for operation from 3 to 3.6 Volts, consuming 90uA.
o
100 C is ± 2.2%. R3
22.5k
Frequency compensation is a
bit tricky. With a load capacitance of Vref
R1
500pF or less the circuit is stable and 42.75k Q3
1.1k 5.25k
Q10 Q4
current source discussed in
20 chapter 5, figure 5-12. The
Q8
R4
300
Q1 Q2
last transistor of the bandgap
reference, Q5, diverts the
SUB
unneeded current from Q6. Q6
supplies about 100uA. With a
Fig. 7-19: Three-terminal version of figure 7-18. maximum hFE of Q7 (at high
current) of 100, the output
current is limited to about
10mA (depending on the size of Q7), which prevent burn-out.
Production variation (3-sigma) over a temperature range from 0 to
100oC is ± 2.4%. The output impedance is 1.5 Ohms and the circuit is
stable with any load capacitance.
One of the circuit elements is a diode, which has a voltage drop of about
600mV. The second circuit element is a multiplied delta-VBE, which also
amounts to about 600mV. Therefore the minimum reference voltage
possible is about 1.2 Volts.
This is only true if we add these two voltages. There are other
approaches which avoid addition. Let's look at two of them.
Vcc Here is the ferment mind of Bob Widlar
I1 again. As early as 1978 he suggested a circuit
100u which works down to 1 Volt supply, a single
battery.
R1
9.75k
R5
800
First off he uses just about the largest
Q2
emitter ratio practical, 50:1 between Q2 and Q1.
Q1
50
This gives a delta-VBE of about 100mV at room
R2
Vref temperature.
50.25k R4
3.75k
The VBE appears at the base of Q1 to
ground. Thus the voltage at the entrance of the
R3
9.75k
current source is higher by a fraction of a VBE.
SUB Now assume R5 to be zero. Thus the voltage at
Vref is that fraction of a VBE plus the delta-VBE of
the 50:1 ratio in emitters of Q2 and Q1. If R1 is
Fig. 7-20: 200mV Vref
design by Widlar.
dimensioned such that the fraction of the VBE
amounts to about 100mV, then we have a
temperature-stable Vref of 200mV.
R5 provides some compensation for changes in I1 and connecting
R4 to a tap at R2/R3 rather than ground creates a minor amount of second
order temperature compensation.
A Vref of 200mV is just about the maximum value you can get from
this design. Even with a much larger emitter ratio, say 200:1, delta-VBE
only amounts to 138mV, i.e. Vref would be about 272mV.
depends on the R1
Q8
R2
R3
22.5k
VBE of Q8 and 7.65k 56k
SUB
the value of R2.
Again, the loop
has a limited Fig. 7-21: Bandgap reference with a minimum supply voltage of
and well- 0.9 Volts.
controlled gain
(the emitter ratio of Q10/Q9 and the 2:1 collector ratio at Q11/Q12), but a
small frequency compensation capacitor is still required. The collector
currents of Q11 and Q12 have a negative temperature coefficient and one
collector of Q12 feeds the output resistor R3. The sum of the two currents
flowing through R3 cause a voltage drop of 250mV, with a temperature
coefficient near zero.
The two currents can be adjusted independently with the values of
R1 and R2, allowing fine-tuning of the temperature coefficient. The
magnitude of the output voltage can be selected with the value of R3
without affecting the temperature coefficient.
Note that the currents depend on the resistor values. They will vary
in production but R3 tracks these variations and the output voltage depends
only on resistor matching.
The output impedance is that of R3. Unless the load draws only a
very small current you will need an output buffer.
This bandgap reference works down to 0.9 Volts supply and the
change in output voltage from 1 to 1.5V Vcc is 0.25%. Power supply
rejection is -55dB up to 10kHz. To keep this low at higher frequencies you
will need an external capacitor (10nF) at the output.
Production variation is ± 3.6% from 0 to 100oC, which illustrates
that the lower the supply voltage the more difficult it is to get high
performance, even if a more elaborate circuit is used.
8 Operational Amplifiers
The majority of op-amps have three stages. The first stage converts
the differential signal into a single-ended one; the second one provides the
bulk on the gain and the third one the required output power. There is no
law that says it has to be this way, it just turned out to be an approach that
works well.
Bipolar Op-Amps
In our first circuit Q1 +V
collectors of Q2 and Q3 we
Q7 Q8
have a high impedance, limited Q6 Q9
only by the base current of Q4
and the Early effects in Q2 and -V
Q3. The second stage, Q4, has Fig. 8-1: Simple 3-stage op-amp.
as a load the current sink Q8
and the base current of Q5, thus
its gain is also limited only by those two. The output stage is a simple
emitter follower with a pull-down current sink.
The operating currents for all three stages are derived from I1
through a multiple current mirror. Q9, with two emitters, delivers two times
I1, an arbitrary choice; more emitters could be used if more current is
required at the output for negative-going signals.
Q7 and Q8 are identical, which is no arbitrary choice. By making
the collector currents of Q4 and Q3 (both halves) equal, Q3 takes as much
base current from Q1 as Q4 does from Q2. Thus there is no designed-in
base current error and the offset at the input is zero (for ideal matching).
Common Mode Range: Describes the minimum to maximum DC level at which the
two inputs are functional. Desirable is rail-to-rail, i.e. from the negative supply to
the positive one, but many op-amps only work with the inputs a volt or two above -V
(or ground for a single supply) to some voltage below +V.
Common-Mode Rejection: If you connect both inputs together, bias them at a
functional DC level and superimpose on the bias a small AC voltage, no signal should
ideally appear at the output. In reality a small signal leaks through and the measure
describes how much smaller this signal is (in dB) compared to the input.
the AC signal. You want to choose the inductor and capacitor large enough
so they have no effect at even the lowest frequency of interest for the
circuit. Neither 1 million Henrys nor 1 Farad are practical values; they
don't need to be.
And here is what we
Y2 Y1
get: The open-loop gain is 180
Phase / deg
100
Gain / dB
80 80
gain decreases steadily and 60
Gain
60
reaches unity (0dB) at about 40 40
12MHz. At that point the 20 20
But, as pointed out in Fig. 8-3: For stable operation the gain of a loop
chapter 6, a phase-margin must reach 0dB before the phase reaches 0
analysis is not the real test of degrees.
stability. AC analysis uses
infinitely small signals (even if it says the signal is 1 Volt) and the operating
currents and voltages are not disturbed. So, to be certain, we would have to
repeat this analysis for the DC conditions over the entire (large-signal)
range of the circuit, a tedious task at best.
We get a more immediate picture by
+V
Vin
In+
+V observing a large-signal pulse.
5
Fig. 8-1
-V
Out
To do this we eliminate the
In-
inductor and capacitor in the feedback
-V
0.8
Fig. 8-4: Buffer connection. 0.6
0.4
path and connect the signal to the 0.2
Output
-0
-0.6
gain of one; with the entire open- -0.8
160 160
In+ +V 5
140 140
Fig. 8-1 Out Phase
-V 120 120
In- -V
Phase / deg
100 100
Gain / dB
5 80 80
1Meg 60 60
R2 R1
40 40
1K 99k L1 Gain
1
20 20
C1
0 0
Vac
Frequency / Hertz
Fig. 8-6: Simulation of loop gain and Fig. 8-7: With the lower gain the loop shows
phase with closed-loop gain of 100. greater phase margin, even though the
compensation capacitor (C1 in figure 8-1) is
reduced to 2pF.
1.5
The gain now shows a 45dB
maximum instead of 85dB and thus
1
drops to 0dB at a lower frequency. We
can now reduce the value of the
Output Voltage / V
0.5
Let's now consider a design in which the inputs can be operated all
the way down to the level of the negative rail and the output swings
(almost) rail-to-rail. The input stage in figure 8-11 is a configuration known
as the folded cascode stage. With an operating current (I1) of 10uA the
voltage drop across R1 and R2 is a mere 50mV, thus the inputs can go
about 250mV below the negative supply rail without saturating Q1 or Q2.
The collector currents of Q1 and Q2 upset the balance of the Wilson current
mirror Q3, Q4 and Q5 and the difference signal is picked up by Q6.
The output Q15 Q16
+V
Q14, a grounded Q6
Q7
C1
emitter amplifier. Q1 Q2
In-
Out
In+ 5p
All other transistors Q4
Q8
in this block serve
its antipode, Q13. I1
Q3 Q5
Q9
10u
Note the Q12
Q14
three diode R1
5k
R2
5k
R3
50k
connected SUB
-V
transistors Q7, Q8
Fig. 8-11: Op-amp with folded-cascode input stage and
and Q9. They set a (almost) rail-to-rail output.
voltage for the base
of Q11. If you follow the emitter-base junctions of Q11, Q12 and Q14, you
notice there are also three diodes in series to the V- rail. Thus, as the input
signal to the output stage moves up and down (by a few millivolts), the
current in Q11 fluctuates. It is this current, amplified by the size ratio of
Q13 to Q10 (here about 6) that becomes the pull-up portion at the output.
Q7 to Q9 are deliberately made larger than Q11, Q12 and Q14 so that the
idle current in the output is small. This creates a small "dead-band" (see
chapter 16) but, because of the large loop gain the distortion is very small
(0.0004% for a ± 4.7Vp signal at 1kHz).
In this circuit we are
Y2
180
Y1
180
fortunate to find a node ideally
160 160 suited for the connection of a
140 140 compensation capacitor to the
120 120
Phase output: at the base of Q6 the signal
Phase / deg
100 100
Gain / dB
0 0
represent a high impedance; and
-20 -201 10 100 1k 10k 100k 1M 10M 100M there is substantial voltage gain
Frequency / Hertz between it and the output. Which
Fig. 8-12: Phase margin of figure 8-11. all says that this op-amp can be
compensated (at unity gain) with a
single 5pF capacitor, even though
the loop gain is 110dB. Not all designs behave that well.
Out / V
bipolar transistor that there is a 0
Q11
Q23 Q7
Q1
In+ In-
Q6 C1
Q2 Out
Q22 Q8 5p
Q4
Q20
I1 Q19 Q21 Q24
10u
Q3 Q5 Q9
Q12 Q14
R1 R2 R3
Q25 Q26 5k 5k 50k
SUB
V-
Fig. 8-14: Op-amp of figure 8-11 with base-current compensation for the input stage.
In the last bipolar op-amp the goal is not an ultra-low input current,
but low-noise performance with a reasonably low input current.
Q18 V+
1 Q3 Q4
Q5
3
Q11 Q12
200u C1 Q6
I1
Q13
10 Q15 5p
SUB
V-
CMOS Op-Amps
CMOS devices have two advantages for op-amps over bipolar ones:
there is no input current (at least not at DC) and, when the transistor is fully
turned on, there is only a simple resistance between drain and source (not
some complex cancellation of two junctions, resulting in an offset voltage
and a resistance). M1 M3 M11 M13 M17
V+
Let's again
first look at a simple W=30u
L=5u
W=30u
L=5u
W=30u
L=5u
W=30u
L=5u
W=90u
L=5u
design. As in figure
M2 M4
8-11 a "folded M12 M14
C1
3-24, so the V-
operating current Fig. 8-16: Op-amp with folded-cascode input stage and a
for the input pair is simple (and limited) output stage.
a constant 20uA.
The four transistors M1-M4 also steer M11/M12 and M13/14,
producing two more accurate currents of 20uA each. The drains of the
input transistors are connected to the sources of M12 and M14, which have
a potential about 200mV below V+; thus the inputs can operate up to (and
about 100mV above) the positive supply.
At balance (In+ = In-) the input pair diverts half of the 20uA current
produced in M11 and M13, i.e. M12 and M14 are left with only half the
current, about 10uA each. The current out of M12 is mirrored in M15/M16
and opposed to that flowing out of M14. With a large input signal the two
currents become unbalanced and each can vary between zero to 20uA. The
voltage created by this unbalance is amplified by the output stage (M18 and
a simple pull-up current source, M17). The idle current of the output stage
is set by the ratio of the channel widths of M1 to M17, i.e. 60uA.
The circuit is Y2 Y1
Phase / deg
load capacitance is small. 40
Gain / dB
50
0
this chapter assume a split power
supply of 3 Volts total, or ±1.5V;
-0.5
they are operational down to
-1 ±0.8V, though with reduced
-1.5 0
performance.
0.2 0.4 0.6 0.8 1
Time/mSecs
You need to be aware of
200µSecs/div
V+
M1 M3 M11 M13 M21
M22 W=10u
M2 M4 L=5u
M12 M14
W=20u
W=20u W=20u L=0.5u
L=2u L=2u W=20u W=20u M=2
M=5 M=5 L=2u L=2u M27
M=5 M=5
M9 M10 W=20u
In+ In-
1p L=0.5u
M19 M23 C1 M=15
W=20u W=20u
Out
L=5u L=5u
R1 M=5 M=5 W=100u W=100u 1p
20u 10k L=0.5u L=0.5u C2
I1 M29
M5 M7 M15 M17
W=10u
M25 L=0.5u
W=20u W=20u W=20u W=20u M=5
L=2u L=2u L=2u L=2u
M=5 M=5 M=5 M=5 W=20u
L=0.5u
The trick in designing a rail-to-rail output is in the biasing of the two output
transistors. You want a small but well-controlled idle current to minimize
any uneven behavior as the output signal is switched from one transistor to
the other. In this circuit there are eight transistors whose only job is to set
this idle current.
Follow M25 and M26, two "diode-connected" n-channel devices,
fed by the current source M24. At the gate of M23 we then have a DC
potential of about 1.2V above the negative rail. There is a second path from
this node to V-, through M23 and M29, also n-channel transistors. Thus the
current in M29, one of the output transistors, depends on the current
supplied by M24 (and derived from I1 through M1 and M2) and the channel
dimensions of M23, M25, M26 and M29. An identical arrangement is
provided for M27 by M19 to M22. With the dimensions shown the idle
current amounts to 70uA.
In the next circuit (figure 8-21) the polarity of the input stage is
reversed and the current mirror for the second stage (M11 to M14) is
designed to have the highest possible output impedance, resulting in an
increased loop gain. Also note that the primary current has been
(arbitrarily) reduced to 5uA.
The lower operating current level has only a minor effect on the
sizes of most devices; they still need to be large to obtain satisfactory
matching, much larger than the process (0.35u, or the higher-voltage portion
of a 0.18u process) would allow. The idle current of the output stage is now
reduced to 10uA.
Open-loop gain is 107dB at low frequency and with no load.
Capacitive loading is still a problem (but much reduced if the minimum
closed-loop gain is higher than 1) and, as before, the closed loop gain is a
function of load impedance.
The input operating range now extends from about +0.8V to 150mV
below the negative rail. If a single supply is used the inputs can function at
or below ground level.
Because of the large dimensions used for the input transistors the
white noise level is a relatively low 21nV/rtHz. Note, however, that they
are run at twice the level of I1.
V+
M5 M7 M11 M13 M21 M24
R1
C1
50k W=15u
1p L=0.5u
M9 M10 M19 M23 M=10
In+ In-
Out
C2
W=20u W=20u W=15u W=10u
L=3u L=3u L=0.5u L=0.5u 1p
M=10 M=10 M=10 M=5 M28
Now let's extend the operating range of the input by using both p-
channel and n-channel devices. In the circuit of figure 8-22 we are adding
an n-channel differential pair which takes over when the DC level at the
inputs reach about +0.8Volts and the p-channel devices get cut off. There
are three voltage regions for the input now: within 0.8 Volts of the negative
rail only the p-channel devices are active; from about -0.8 Volts to +0.8V at
the inputs, both pairs amplify and within 0.8V of the positive rail only the
N-channel devices amplify.
When both pairs are active, the open-loop gain is at a maximum,
reaching 120dB. When the common-mode level is either high or low, the
loop gain drops by 10dB. Many schemes have been offered in the literature
which hold this gain more constant (e.g. by allowing only one pair to
operate at a time), adding another dozen devices. For most applications the
benefits of this measure are limited; in fact simultaneous operation of both
pairs increases performance (noise, for example drops to 20nV/rtHz,
compared to 30nV/rtHz when only one pair is amplifying).
V+
M5 M7 M11 M13 M21 M24
M6 M8
M12 M14 M22
W=20u W=40u
L=2u L=2u W=20u W=20u W=15u
I1 M29 M30 L=2u L=2u L=0.5u
M=5 M=5 M=10 M27
5u
W=20u W=20u C1
L=3u L=3u W=15u
M=5 M=5 1p L=0.5u
M19 M23 M=10
M9 M10
In+ In- Out
W=15u W=10u C2
W=20u W=20u L=0.5u L=0.5u
L=3u L=3u M=10 M=5 M28
1p
M=10 M=10
M26
M2 M4 M32 M16 M18 W=1.5u
L=5u
W=10u
W=5u W=5u W=10u W=10u W=10u L=0.5u
L=5u L=5u L=5u L=5u L=5u M=5
V-
The two input stages work with the second stage (M11 to M18) as
folded cascodes. The lower part of the second stage (M15 to M18) is a
current mirror derived from M1 to M4, set here at about 10uA; the upper
part (M11 to M14) mirrors the current again, so that at M19/M23 the
currents cancel if there is no input signal. M19 and M23 set the bias current
of the output transistors (M27, M28)
With an input signal one or both input stages change the currents in
the second stage, resulting in a net positive or negative current through
M19/M23, which is translated into a larger current at the output.
A common problem in op-amps using two separate input stages is
created by the random nature of the offset voltage. Suppose one pair has an
offset voltage of +5mV the other -5mV. As the signal moves from one
stage to the other, this causes a jump of 10mV, creating distortion.
Auto-Zero Op-Amps
Auto-zero or chopper stabilized amplifiers have been around for
decades but continue to evolve. In a modern embodiment two amplifiers
are used, checking up on each other.
Each amplifier has a "Trim" input, i.e. a single node which changes
its offset voltage in both directions.
A built-in oscillator flips the two switches periodically at a rate of a
few hundred to a few thousand Herz. In position A the inputs of amplifier 2
are shorted together and its own offset voltage is amplified (with the open-
loop gain) and corrected by feeding the output to the trim input. The
required trim voltage is stored in capacitor Ca.
In the second phase of the oscillator the inputs of amplifier 2 are
connected in parallel to those of amplifier 1 and its output now feeds the
Feedback trim input of amplifier 1. With the
charge remaining across Ca, amplifier 2
In+
In+ continues to be nulled and thus corrects
Amp. 1 Out Out
the offset of amplifier 1. As the
In- In- Trim
oscillator switches back to phase A this
correction voltage remains across
capacitor Cb. With the high open-loop
A B
In+ gains of both amplifiers the offset
Amp. 2 Out voltage is now reduced to microvolts.
In- Trim Since the correction is done repeatedly,
A B
temperature drift is also much reduced.
Ca Cb
There is an additional benefit.
Anything sensed by amplifier 2 below
the switching frequency is treated as an
Fig. 8-23: Auto-zero op-amp.
offset. This includes flicker (1/f) noise,
which is completely eliminated. Above the switching frequency the
behavior of the auto-zero amplifier is identical to a regular op-amp.
Distortion in an Op-Amp
An op-amp is basically a non-linear circuit. The input stage, for example, can
accommodate only a small differential voltage before it is limited by the input devices,
both bipolar and CMOS.
Feedback reduces the distortion caused by these limitations. Increasing the
amount of feedback increases the linearity.
If you increase the open-loop gain by a factor of 10 (20dB), distortion drops
by a factor of 10, assuming that, by increasing the gain you have not added more
distortion.
In 1919 John M. Miller was physicist with the National Bureau of Standards
when he wrote a paper on how the grid capacitance of a vacuum tube was so much
larger in use than measured statically. The voltage gain, he said, multiplies the
capacitance between grid and plate. What he described has been known as the Miller
effect or the Miller capacitance ever since.
Miller went on to doing research at Atwater Kent, RCA and the Naval
Research Laboratory. In 1953 he was awarded the IRE Medal of Honor.
The exact same effect was found in both the bipolar and MOS transistor. In
most applications it is detrimental, limiting the frequency response; in IC op-amps,
however, it has been helpful, greatly decreasing the size of the compensation
capacitance.
John M. Miller: "Dependence of the input impedance of a three-electrode vacuum tube upon the
load in the plate circuit", Scientific Papers of the Bureau of Standards, 1920, pp. 367-385.
9 Comparators
small second-order error due to the fact that the collector-base voltage of Q4
is larger.
But don't let this observed accuracy fool you into believing that this
is what will happen in production. Move on to a Monte Carlo analysis and
you will find that the offset voltage of the differential pair moves the
switching point (by about ± 1mV, depending on the process and the size of
the transistors).
Vcc
W=20u
is now 110dB. 20u
M1 M2
L=5u
W=2u
There is no (DC) input In Ref
L=0.5u
Out
flop with precisely controlled gain, giving the circuit a snap action. In
addition, the diode connection of Q4 makes Q5 into a current mirror (each
collector sources one-half of the current). This current is fed into a 400
Ohm resistor, causing the reference
voltage appearing at the base of Q2 to 2.014
2.01
V
2.004
to about 25uA (the exact level depends Fig. 9-4: Switching levels with
on the output voltage because of the hysteresis.
Early effect). You now have to
increase the input voltage to 10mV above the reference level to turn the
output current off.
Figure 9-5 shows the same circuit in CMOS, with the current output
(M8) opposed by a current sink of half the level (M13), and a logic stage
Vcc
M3 M4 M5 M6 M7
W=1u
M11 M12 M13
L=0.5u
added. CMOS has an advantage here in that the custom sizing of the
transistors allows the amount of positive feedback to be set in precise
increments (M3-M6). Note that the operating current (Ibias) has been
Vcc
M17 M20 M5 M6 M7 M8
M15
W=20u
Ibias In M1 M2 Ref L=1u
M=2 W=2u Out
20u L=0.5u
W=20u W=20u
L=5u L=5u M16
M3 M4
W=20u W=20u
L=5u L=5u
In this example the active load of figure 9-5 was chosen, again with
sufficient positive feedback to give a snap-action (M5-M8). Note that M5-
M8 and M9-M12 have a considerably large w/l ratio compared to the
corresponding M1 - M2 and M3 - M4 to allow the input to go slightly
beyond Vcc and ground.
Current Comparators
I1 I2
Vcc
When you use the word comparator you
Q4
automatically assume that voltages are compared.
But this does not always have to be the case,
In1 In2
Out
sometimes it is useful to compare currents.
Q3 With a simple current mirror (Q1, Q2)
Ibias
even a small difference in the magnitudes of I1
Q1 Q2 and I2 will show up quite drastically at the base
SUB
of Q3, turning it on or off.
The base-current error is eliminated if
Fig. 9-8: Bipolar Current Ibias is set at twice the level of I1 and I2. The
comparator. only remaining error is due to the Early effect of
Q3, which is easily reduced by adding another
NPN stage and using a more sophisticated current mirror in place of Q4.
Vcc
The CMOS version is M4 M5
channel-shortening) apply. M3
M7
10 Transconductance Amplifiers
1
gm =
re
Thus with an emitter current of 100uA the transconductance is
(1/260) Ohms, i.e. a 1mV signal at the base cases a change in collector
current of 3.8uA. In a differential stage the transconductance is half of that
since there is an re in each transistor (and we double the total current so that
each emitter receives 100uA).
In figure 10-1 the currents are mirrored in a ratio of 1:1 so that the
collector currents of Q1 and Q2 appear unchanged at the output. With no
signal at the input they cancel each other but, as one input is moved up or
down, one current becomes larger and the other one smaller by the same
amount. Thus the total transconductance is doubled and we have the same
value as for a single transistor.
Without some DC resistance at the output a transconductance
amplifier is really quite impractical. Even the slightest mismatch in any of
the transistors would slam the output voltage into one of the supply rails; we
need some impedance like Rload to keep this voltage near the center. Rload
converts the current output into a voltage output, which means that we no
longer have a transconductance amplifier but simply a voltage amplifier
(with a high output impedance to boot). Very few of the OTAs are actually
used as transconductance amplifiers.
With Rload back in the circuit the total voltage gain is now simply:
Rload q
Av = = ∗ Ie∗ Rload
re k∗T
So, if you have to handle a signal greater than about 20mVp, this
circuit is a poor choice. You cannot use feedback or emitter resistors (for
Q1 and Q2) to linearize it, it would interfere with the variable gain.
There is another problem, not just for this circuit but for all such
schemes: offset. A mismatch in not only the input stage but all three current
mirrors will show up as an offset (and added distortion) at the output,
increasing in magnitude as Igain increases. In this circuit this amounts to
60mV worst-case at 100uA. Also, remember that bipolar transistors have
input currents.
There is help, though, and as usual you need to add a few more
devices. If we +5V
connect diodes to Q3 Q5 Q7 Q9
we have Q15
impedance is low I2
Igain
I3 Q12 Q14
because of Q16 20u SUB 20u
Gain / dB
linear change in current -20
results in a logarithmic
change of gain. Thus, for -30
0 unaffected by absolute
variations.
-0.5 Figure 10-4
shows the waveform that
-1 appears at the output.
There is a very large
-1.5 change in the level of the
0 0.2 0.4 0.6 0.8 1
Time/mSecs
signal, which is of course
200µSecs/div
the purpose of the circuit.
Fig. 10-4: Output waveforms (1kHz). Because of the offset
voltage a
"transconductance" amplifier is best suited for audio and filter applications
with the output capacitively coupled to the next stage.
The concept works in CMOS too, but the fundamentals are different.
A CMOS transistor naturally takes a voltage at the gate and delivers a
current at the drain, and this transconductance varies as the square of the
operating current.
Figure 10-5 is the
+1.5V
M3 M4 M7 M8
same configuration as figure
W=20u W=20u W=20u 10-1, with NPN transistors
W=20u
L=5u L=5u L=5u L=5u
M=2 M=2 M=2 replaced by N-channel
M=2
In the gain vs. operating current plot (figure 10-6) the range is
extended down to 1nA
just to show the wide 20
range achievable. You
10
notice, however, that at
the high-end the circuit 0
Gain / db
logarithmic behavior; to -20
straighten this line the -30
transistors would have to
-40
be even larger.
Transconductance -50
temperature dependent 1n 2n 4n 10n 20n 40n 100n 400n 1µ 2µ 4µ 10µ 20µ 40µ 100µ
decrease of gain (at any Fig. 10-6: Gain (in dB) vs. Igain.
current) of about 2dB from
0 to 100oC. Also, a CMOS transconductance amplifier has the same offset
problem as a bipolar one; at 100uA this amounts to ± 30mV. Unlike the
bipolar version, however, this circuit has no DC input current.
W=20u W=20u
L=2u L=2u W=20u W=20u W=20u W=20u
M=10 M=10 L=2u L=2u L=2u L=2u
M=10 M=10 M=10 M=10
Output
In- M1 M15 M16 M17 M18 M19 M2 In+
M11 M12
W=20u W=20u
M25
L=2u L=2u
M=5 M=5
W=10u
M20 M21 M22 M23 M24 L=5u M13 M14
M=2
-10
-20
line above 20uA.
Because of the many
-30 additional devices there
is also a slight
-40
deviation below about
-50 10nA. Even so, the
10n 20n 40n 100n 200n 400n 1µ 2µ 4µ 10µ 20µ 40µ
gain control has a range
Igain / A of more than 70dB.
Fig. 10-8: Gain (in dB) vs. operating current (Igain). Note that the
output impedance is
Summer of 1970. The economy was at the bottom of the cycle and
Signetics, the promising young company I had joined just two years before,
laid off half of its employees.
Disgusted with the turn of events, I decided it was time to strike out
on my own and rented space between two Chinese restaurants in downtown
Sunnyvale, California. Signetics (now Philips) lent me the equipment I
needed and gave me a one-year contract to develop a new IC.
The idea for the new IC came from the work I did at Signetics on the
phase-locked loop. I had needed an oscillator whose frequency could be set
by an external resistor and a capacitor and was not affected by changes in
either supply voltage or temperature. Several products resulted from the
basic design, among them the NE566 Voltage-Controlled Oscillator.
The oscillator
Vcc
Rext.
contained first of all a
voltage-to-current
R 5k
+
Comp. 1 converter. The reference
V to I Converter
voltage at the positive
- -
5/6 Vcc Q1 input terminal of the op-
+
amp is not regulated, it is
Flip-Flop
S simply a fraction of the
R supply voltage. Feedback
5k
C to the op-amp keeps the
Cext. Comp. 2
voltage across the
+
external resistor at the
-
5k
same level and thus the
Mirror current through the
resistor becomes
(1/6*Vcc)/Rext.
Fig. 11-1: The basic 566 Oscillator. In the actual
circuit the comparators and the flip flop are combined Depending on the
in one Schmitt trigger. state of the switch
controlled by the flip-
flop, the external capacitor is either charged with the current, or discharged
with a current of the same magnitude through a 1:1 current mirror.
Vcc
R1 R2 R3 R4 R12
4.7k 830 4.7k 1K R7 6.8k
5k
Q5 Q6 Q7 Q8 Q9 Q21
Q19
Q22
R13
R10 3.9k
Threshold FM 15k
Q1 Q4 Q23
R11
Q2 Q3
4.7k
Output
R5
Q15
Discharge 10k
R16 R14
Q24
100 220
R6 R9 R15
Q14 100k 5k 4.7k
(pinched) SUB
Gnd
Both comparators use Darlington input stages. This makes the timer
fairly slow, but allows an extreme range of external resistance. Comparator
1 consists of Q1 to Q8. The four PNP transistors form a current mirror with
gain, provided by the unequal emitter resistors.
The output of this comparator feeds into a 4.7kOhm resistor (R11),
which is part of the cross-connection in the flip-flop (Q16, Q17).
Comparator 2 (Q10 to Q15) resets the flip-flop.
The output stage, which must be able to sink or source some
200mA, is controlled by Q20. In the high state the Darlington pair
Q21/Q22 delivers the current, but at a cost of a voltage drop of about 2
Volts. In the low state Q24 receives sufficient base current to work alone
up to about 50mA; beyond that, as the voltage drop increases, Q23 feeds
extra current into the base circuit.
There are several flaws in this design, indicative of the early period
of IC design (and the inexperience of a rookie designer). Neither
comparator is well balanced, showing offsets of as much as 30mV. The
circuit can get away with that because the voltage swing is quite large.
The operating currents are quite large; the lateral PNP transistors run
at up to 1mA. That was acceptable at the time since the devices had 10um
geometries; today it would be excessive.
10
R1 Output
91k Vcc
8
Threshold FM Vcc RC
555
12 6
Discharge
V
Output
4
Trigger Reset R2
Trigger 100
Gnd Trigger
2
C1
Time/µSecs 20µSecs/div
Fig. 11-4: Timer connection of the 555. Fig. 11-5: Timer waveforms.
12
10
R1 Vcc
50k R3
Threshold FM 100 V1 8
555
12
V
6
Discharge Output
R2
50k 4
Trigger Reset
Gnd 2
1n
C1 0 120 140 160 180 200 220 240 260 280 300 320 340
Time/µSecs 20µSecs/div
Fig. 11-6: Oscillator connection of the 555. Fig. 11-7: Oscillator waveforms.
In the oscillator connection there are two external resistors and the
voltage across C1 moves between 1/3 Vcc and 2/3 Vcc with a frequency
and duty cycle of:
1.46 R2
f = DutyCycle =
( R1 + 2 R2)C1 R1 + 2 R2
Q3 R4
20k
R8
3.75k
555
Q6 Q9 Q10
Second Version
Q31
R1 Q4
75k FM Flip-Flop
Q32 Q33 Q34 Q35
Q11 Q14
Q5 R5 Q36
20k Discharge
Q1
Q17
R3 Q28 Q37
34k R6
Q2 Q7 Q8 Q15 Q16 20k Q38
Vcc
R13
R7
Q41 30k
7.5k
Threshold Q40 Q44 Q48
Q22 Q26
Q19
Q50
Q45
Q47
Output
R10 R11 R12 R14
Q20 7.5k 7.5k 7.5k 7.5k
Q23 Q42
Trigger Q18 Q25 Q51
Reset
R9
Q39
Q27 15k Q49
Fig. 11-8: An improved version of the 555 timer, 33 years after the original design.
First off, the new timer gets a proper bias circuit (Q1 to Q5) to hold
the operating currents more constant over the wide supply voltage range.
This (and a few other steps) extends the operating voltage down to 3 Volts.
Comparator 1 (Q6 to Q17) now has a balanced active load (Q15,
Q16) which reduces the error in the timer mode to about 0.5% and the
temperature drift to 3 ppm/ oC without any loss in speed. The change in
timing from 3 to 15 Volts is a mere 0.05%.
There are two changes in comparator 2 (Q18 to Q27): a small
operating current for the outer Darlington transistors, which greatly
improves switching speed, and a balanced active load, which makes the
trigger level considerably more accurate.
The flip-flop (Q28 to Q36) is a new design; it operates in a current-
mode for maximum speed at the lowest possible current. The two 50uA
currents generated by Q31 are split by a pair of lateral PNP transistors; one
quarter of the current is fed into the base of the opposite flip-flop transistor,
another quarter turns the reset transistor on and off and one half of the
current is used to steer the output stage. The voltage swing at the collectors
of the flip-flop transistors (Q30, Q36) is 2VBE.
The most significant change is in the output stage. The base current
for the lower output transistor (Q51) is no longer derived from a resistor. A
small amount of current is injected into the bases of three transistors, forced
to be equal by the three resistors R10, R11 and R12. This (plus an
additional current delivered by Q45) starts a positive feedback loop formed
by Q40, Q41 and Q42. Q40 is about seven times the size of Q41 and Q42
has one emitter while the output transistor has 24. This loop then provides
whatever current is needed to keep Q51 fully turned on.
Positive feedback loops are always dangerous, they can run away or
refuse to turn off. In this case the loop is contained by the collector
resistance of Q43 and can be opened up by turning Q43 off.
Replacing the Darlington configuration in the upper part of the
output stage with a compound (PNP/NPN) transistor reduces the voltage
drop. Base current for this part is provided by Q47. Q44, Q46 and Q49 aid
in turning the power devices off rapidly and eliminate the large transient
current.
With these measures the current consumption is now down to
0.85mA from 3mA (typical) at 5 Volts. At 15 Volts the circuit consumes
1.2mA (down from 10mA). Minimum operating voltage is 2.5 Volts (-40oC
to 100oC).
Shortly after the 555 came out Intersil announced a CMOS version.
It was (and still is) done in a 15-Volt process, which requires large
dimensions and is inherently slow. The circuit is not directly compatible
with the bipolar version, lacking high current outputs.
Except for this weakness, CMOS is ideally suited for a timer: there
is no input current and thus no need for Darlington stages.
Figure 11-9 shows a design using a more modern 5-Volt (0.5um)
process. The comparators are conventional (as discussed in chapter 9), with
the dimensions of the devices chosen so that the threshold and trigger inputs
can move rail to rail and their matching is adequate for precision operation
(3ppm/ oC).
M3 M4 M6 M8 M18 M19
R2
Vcc
40k
W=10u W=10u W=10u W=10u W=10u W=10u
L=0.5u L=0.5u L=5u L=5u L=5u L=5u
M13
Discharge
M15
Reset
Threshold M1 M2 M22 M25
FM W=20u
L=0.5u
W=5u
W=10u W=10u R3 L=0.5u W=4u W=20u
L=0.5u L=0.5u 40k L=0.5u L=0.5u
M9 M10
M=10
R1
100k Output
W=10u W=10u
Trigger L=0.5u L=0.5u
M17
M20 M21 M23 M24 M26
+12V
R1 R2 R14
5k 5k 13k R5 R6
375 375
Square
Q1 Q2
Q3
R9
7.5k
R3 R4 R7 R8
Q19 Q22
Rext
C Q16
65k A
Q18 Q21
Cext. Q13 Q14 R10
20p 7.5k
R Q23 Q25
R11
B
7.5k
Triangle Q24
Q7 Q20 Q26
Q6 Q8 Q17
Q15
Q12
SUB
6.5
square-wave can be obtained.
6
This is not an oscillator
5.5
of ultimate precision, but it
5
delivers a good-quality wave-
4.50 0.5 1 1.5 2 2.5 3 form up to at least 1MHz. The
Time/µSecs 500nSecs/div temperature coefficient is
190ppm/ oC and the change in
Fig. 11-12: Wave-form at 1MHz.
frequency from 9 to 15V
supply is 1.7%. As always, these results are based on one particular
process; it is a good idea to re-simulate the design for the process you are
using.
R15
R1, encounters attenuators at six Q9 1.2k
R20
different voltage levels. Follow 2k
Q10
R16
Q12
there is no attenuation; at about 200
R17
1.4k
0.6 Volts R11 kicks in, held at
Gnd
that level by Q6; at the next B
7.5
parallel to R11 and at the final
7
level an even smaller resistor, R2,
reduces the signal even further.
6.5
Using only three clipping
V
6
levels in each direction, the
5.5 resulting sine-wave has a
5 distortion of only 1%. Using more
4.5 levels reduces the distortion, but
40
is likely to require trimming.
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time/mSecs 200µSecs/div
Oscillators (and timers) can
Fig. 11-14: Triangle and sine-wave.
often be very simple for non-critical
applications. Suppose you wanted to create a brief pulse once a second, for
example to flash an LED. The frequency or pulse-width need not be
precise, i.e. the design does not require two sophisticated comparators; a
simple Schmitt Trigger will do.
180
resistor.
3.5
160 LEDs have a rather
3
140 large forward voltage drop
I(LED-anode) / mA
2.5
120
(about 2 Volts), so a supply
RC / V
100 2
80
voltage of at least 2.5 Volts is
1.5
60 required. The current through
1
40 LED Current the LED (40mA) is primarily
0.5
20
determined by the size of Q1;
0 0 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8
it operates in the high-current
Time/Secs 200mSecs/div
region, where the gain has
Fig. 11-16: Wave-forms of the pulse generator. already decreased but the
spread of hFE becomes
narrow. Average current consumption is 1mA.
Q18
+3.3V
It is interesting to
Enable consider a CMOS design
W=10u
L=0.35u for the same function.
M3 M4 M9
We can avoid using a
Rext
1Meg resistive divider (and
W=10u W=10u W=1u
L=0.35u L=0.35u
M=9
L=0.35u M11
LED thus save current) by
using two comparators
W=10u
M1 M2
L=0.35u and making two of the
RC
W=20u
L=0.35u
W=20u
L=0.35u
M12 M14
LED transistors in each
M=9
R4 W=1u W=1u
M16 comparator nine times
130k L=1u L=1u
W=20u
the size of the others.
M5 M6
L=0.35u
M=5 This results in an offset
W=20u
L=0.35u
W=20u
L=0.35u M17
of some 200mV. The
M=9 M10 M13 M15
reference potentials for
330n
Cext R1
20k
M7 M8
W=10u W=1u W=1u
W=10u
L=0.35u the comparators are the
L=0.35u L=1u L=1u M=4
W=10u W=10u
M=4 supply lines. As the
L=0.35u L=0.35u
M=9 voltage across the
external capacitor rises to
200mV below the
positive supply, the upper
Fig. 11-17: CMOS design for a pulse generator.
comparator (M1 to M4) sets the flip- 3
flop (M12 to M15); as it falls to Vcc - 200mV
1.5
R1 (20msec). The two times are
surprisingly accurate, exhibiting a 3% 1
The output current, on the other 1.8 2 2.2 2.4 2.6 2.8 3 3.2 3.4
A second example, a timer this time, shows just how low a power
dissipation can be achieved if a resistor divider is avoided: the circuit in
figure 11-19 draws 1uA at 1.8 Volts.
The entire circuit (save the output inverter) is powered from the
Start pulse, rising from ground to Vdd. The logic input must stay high
longer than the set
Start M3 M4 M5 M6 Vdd
timing. Alternately, if
W=5u W=5.2u W=5.2u W=5u
the Start terminal is
M10
R1
L=0.18u L=0.18u L=0.18u L=0.18u simply connected to Vdd,
M12
10Meg
M8
W=2u
the circuit becomes a
L=0.18u
W=10u
start-up timer.
RC M1 M2
W=2u
L=0.18u
L=0.18u M1 to M6 form a
Output
W=5u W=5u
comparator. By making
M7 M9 M11
L=0.18u
M=10
L=0.18u
M1 ten times as wide as
M13
W=0.18u W=3u W=3u M2, an 80mV offset is
3.3n
C1
L=60u L=0.18u L=0.18u
W=5u created. The gate of M2
L=0.18u
is connected at Vdd, thus
the comparator switches
Fig. 11-19: A 1.8-Volt timer which consumes just 1uA.
at 80mV below Vdd.
The switching action is
enhanced by employing a small
1.8
amount of positive feedback; M4
1.6
and M5 are slightly wider than +V - 80mV
1.4
M3 and M6 and deliver their
1.2
drain currents to the opposite
1
sides (see also figure 9-5)
V
Capacitor
0.8
The operating current for Output
0.6
the comparator is provided by
0.4
M7, a long and thin transistor.
0.2
The advantage of such a device is
00 20 40 60 80 100 120 140 160 180 200 220
size: it produces about 0.6uA
Time/mSecs 20mSecs/div
using a relatively small area; a
resistor doing the same job Fig. 11-20: Switching threshold of the 1.8V
(1.2MOhms) would be painfully timer.
large.
On the other hand an MOS transistor with its gate connected to the
supply is hardly a constant current source. With a 25% change in supply
(1.6 to 2 Volts) the current changes 70% (0.5 to 0.85uA). Here we can
afford to live with this shortcoming.
The two outputs of the comparator are level-shifted by M8 and M10
and the active load M9/M11 to fit the input of an inverter.
Simulation of Oscillators
You have just drawn up a great idea for an oscillator and start the
simulation. Nothing happens, you get nothing but DC levels.
This situation is all too common. The simulator is trained to first
find an operating point, i.e. set the DC voltages and currents so all the
device equations are satisfied (in simulator-speak: find convergence). Then
the transient analysis starts and the computer finds that all the voltages and
currents remain unchanged over time.
In real life the circuit may start at exactly the same point, but no
voltage or current stays unchanged, there is noise. No matter how tiny these
fluctuations are, they move the circuit to a slightly different operating point
and it becomes apparent that movement in one direction is the path to be
followed. Thus, gradually, the oscillation builds momentum.
Without any noise (or some sort of transient disturbance) no circuit
would oscillate; it would just sit there, precariously balanced.
It is of great advantage to have a simulator which allows real-time
noise (i.e. all currents and resistors actually produce the appropriate amount
of noise not only for a (small-signal) ac simulation, but during a transient
analysis as well). With this feature a properly designed oscillator will
always start. If you are using a simulator which has no real-time noise, you
may have to coax the oscillator by jarring it with a pulse, e.g. step the
supply voltage abruptly to a higher level.
But there is a second potential problem: it may take a long time for
the oscillation to build up. For the circuits discussed so far in this chapter
this is no great worry, but for the type of circuits we are about to encounter
this can be very frustrating.
Take a crystal oscillator, for example. A high-quality crystal can
take up to a second to start oscillating. At 10MHz, that amounts to 10
million cycles the simulator has to go through, in very small time steps to
catch any movement. If you are not aware of this nuisance, you may sit
there watching flat lines and come to the conclusion that your oscillator
does not work.
LC Oscillators
Q5
Q8 (about 12uA) is injected
Q11
Q4 Q6 into the bases of Q8 and
R1
30k
Q11 to bring them into a
SUB current level at which there
is sufficient gain. TX1 is
Fig. 11-21: Sine-wave oscillator with large in reality a center-tapped
amplitude.
inductor, shown as a
transformer with two windings for
the simulation.
20
The collectors of the 18
oscillating transistors start at the 16
supply voltage, 10 Volts. After a 14
+10V
few hundred cycles of gradually 12
10
6
emitters of Q5 and Q6 at the 4
negative end; at this point the peak- 2
to-peak amplitude has reached 21 0
550 600 650 700 750 800
Volts, more than twice the supply
Time/nSecs 50nSecs/div
voltage. The action of the center-
tapped inductor is that of a see-saw: Fig. 11-22: The voltage swing of the
one end dips to ground (or slightly oscillator extends to twice the supply
voltage.
below) while the other peaks at a little
above 20 Volts.
Crystal Oscillators
Let's start with the circuit commonly used in CMOS: the crystal is
connected between the input and the output of an inverter. Since an inverter
is ill-equipped to remain in a state
between low and high, R1 is M1 M3 M5
+3.3V
1
f =
2∗ π ∗ L1 ∗ C1
Gain / db
100 10
80
0 +3.3V
Phase M1
60
-10
40
W=2u
-20
20
L=0.35u
Phase / degrees
20
dB / db
oscillation) until about 10.015MHz. 0
-30
1.5
for the oscillation to reach full
amplitude. If you want to measure
1
the frequency accurately, you need
0.5
to do this in very fine steps (say
00 +5V
0.5 1 1.5 2 2.5 3 3.5 4
R1 R2 R3 C1 C2
Time/mSecs 500µSecs/div 11k 7k 2.5k 10p
25p
Pulse
Figure 11-28: Start-up of a crystal Crystal Sine
oscillator.
R4
Q1 Q3 Q4
10k
1nsec), i.e. you have to wait for
30,000 cycles before you can see R5
10k
the actual wave-form. Q2 Q5
SUB
Figure 11-29 shows a
different approach, for a bipolar Fig. 11-29: Alternate crystal oscillator with
process. Gain and a 180 degree bipolar devices.
(VCO).
20u
I1 The signal inputs are at a certain
DC bias level, say 1V or 1.5V, high
enough to exceed the threshold voltages.
Fig. 12-1: The phase detector.
Now imagine a square-wave at the
Diff Probe / mV
Output
200mV differentially, centered at 0
connected to R1 (through M1) and Fig. 12-2: With the VCO signal in-phase,
the drain of M6 to R2 (through the output is a rectified sine-wave with a
M4). During the second phase this positive average.
connection reverses: The drain
current of M5 flows through M2 and R2 and that of M6 through M3 and
R1. Thus, ignoring C1 for now, the output across the two load resistors is a
rectified sine-wave with a positive
200 average value.
150
Now, let's keep the
Output
100
frequencies the same but shift the
phase of the VCO signal by 90
50
Diff Probe / mV
Input
degrees (figure 12-3). The signal
0
is now chopped at the moment it
-50
reaches its peak amplitude and the
-100
output shows equal positive and
-150 negative excursions. Thus the
VCO
-200
0 0.2 0.4
average differential output voltage
0.6 0.8 1 1.2 1.4
Diff Probe / mV
away in frequency, e.g. 800kHz. 0
Since the two frequencies are not -50
synchronized, there is no phase Output
-100
relationship yet. At this point the
-150
phase detector is merely a mixer,
-2000
producing several new 0.2 0.4 0.6 0.8 1 1.2 1.4
Time/µSecs 200nSecs/div
frequencies, such as the
difference between the two Fig. 12-4: At 180 degree phase shift
frequencies and various the average of the output is negative.
combinations of harmonics. The
one of interest is the difference,
200kHz; it is still too high to pass
through the filter.
As we move the input
signal gradually higher in
frequency, there comes a point
where the difference frequency is
low enough so that some of the
signal passes through the filter and
Fig. 12-5: Block diagram of a simple phase-
locked loop. starts influencing the VCO. The
signal is not rectified yet, it is still
AC, but the VCO starts to jitter
around its free-running
frequency. Move the input signal 1.3
Loss of Lock
just a little higher in frequency
and suddenly the jitter disappears 1.2
1.1
continue to track it, until the loop Fig. 12-6: Locking behavior of a phase-
finally runs out of control voltage. locked loop.
This is illustrated in figure 12-6
where signal frequency is swept from low to high over a 5msec period.
3. The error signal (i.e. the output of the low-pass filter) is a measure of
frequency deviation. If the input signal is frequency modulated, this output
is the demodulated signal.
Output
W=5u W=5u W=5u W=5u W=5u W=5u W=5u W=5u W=5u W=5u
L=0.35uL=0.35u L=0.35uL=0.35u L=0.35uL=0.35u L=0.35uL=0.35u L=0.35uL=0.35u
I1
2u
M4 M5 M9 M10 M14 M15 M19 M20 M24 M25
W=1u W=1u W=1u W=1u W=1u W=1u W=1u W=1u W=1u W=1u
L=0.35u L=0.35u L=0.35u L=0.35u L=0.35u L=0.35u L=0.35u L=0.35u L=0.35u L=0.35u
Fig. 12-8: Current-controlled oscillator for operation between 6MHz and 300MHz.
10u
I1 W=20u W=20u W=20u M22 M23
L=0.35u L=0.35u L=0.35u M25 M26 M28
M=2 M=2
W=2u W=2u
M9 M11 M13 L=0.35uL=0.35u W=2u W=1u W=1u
L=0.35u L=0.35u L=0.35u
1.5V
W=5u W=5u
R2 L=0.35u L=0.35u
5k
M1 M2 M14 M16
C1 R6
1V 20k M29
W=10u W=10u W=20u W=20u
L=5u L=5u L=10u L=10u 2p
Vin
M=2 M=2 W=0.5u
L=0.35u
R7
10k M6 M7 R3 M15 M17 M24 M27
10k
+3V
3
R1 R2
2.8
20k 10k Output
2.6
2.4
M1 M2 Output
Input 2.2
V
2
Input
W=3u W=3u 1.8
L=0.35u L=0.35u
1.6
1.4
R3
5k 1.2
1
0 2 4 6 8 10
Time/µSecs 2µSecs/div
13 Filters
1
f 3 dB =
2πRC
Fig. 13-1: Single-
pole RC low-pass For the
filter.
values 0
shown in -2
-6
10kHz. Below about 1kHz there
Attenuation / dB
-8
is no attenuation. At 10kHz the -10
signal at the output is down by -12
of the curve (figure 13-2) upward 100 200 400 1k 2k 4k 10k 20k 40k 100k
low-pass filter is said to have an attenuation of 20dB per decade or 6dB per
octave (doubling of the frequency).
We don't need to be R1 R2
+ +
satisfied with just one RC 15.9k C1 15.9k C2
network, we can connect 1n 1n
several of them in series (i.e.
cascade them). But we need Fig. 13-3: A poor way to sharpen filter response.
to put a buffer in between the
stages, otherwise the network 0
little over 2kHz. 100 200 400 1k 2k 4k 10k 20k 40k 100k
0
single RC network (i.e. 40dB per
-5
decade or 12dB per octave) and
the -3dB point has remained at -10
Attenuation / db
10kHz. -15
-20
8.2n 1.5n
R7 R6 R11 R10
+ +
3.6k 8.58k 7.82k 21.5k
0
1n Bessel
1n C5 -5
C3 Sallen & Key Chebyshev
Butterworth
-10
C12 C2 -15
Butterworth
1.5n 3.3n
dB
R19 R18 R3 R2
+ + -20
5.1k 16.2k 3.32k 9.03k
1n -25
1n C1
C11 Sallen & Key
Bessel -30
-35
C13 C15
deg
capacitors. But there is a -200
Butterworth
80
its pulse response. In figure 13-11 a
60
40
1.2
20 Chebyshev
Bessel
0
1
1k 2k 4k 10k 20k 40k Bessel Butterworth
Frequency / Hertz 0.8
0.4
100usec pulse was applied to the
input. We expect a rounding of 0.2
Input
the corners at the output but, 0
considering that all three filters 0 20 40 60 80 100 120 140 160 180
have the same cut-off frequency,
Time/µSecs 20µSecs/div
the Bessel filter does the best job.
Fig. 13-11: Pulse response of the three
filters.
How do we get the values for
the resistors and capacitors? If you open up a text-book on filters, you will
see elaborate tables giving you coefficients for Butterworth, Bessel and
Chebyshev functions. This is no longer necessary. There are a multitude of
programs available on the web (many of them at no cost), which calculate
these values for you. Search for "active filter software".
C2 C4
Let's look at two more 1.08n 1n
low-pass filters, using designs R1
11.76k
R2
11.76k
+
R5
12.24k
R6
12.24k
+
High-Pass Filters
R2 R4
There is no mystery
6.1k 14.7k
to converting a low-pass C1 C2
+
C3 C4
+
filter into a high-pass one: 1n 1n 1n 1n
0
frequency end, but at the same rate as
-10 that of a low-pass filter, 80dB per
-20 decade for a fourth-order filter.
-30
dB / db
Band-Pass Filters
Take the R2
0
3.3k
second-order low- -2
R1 C2 -4
pass filter of figure + -6
83.1k
13-5 and convert 1n -8
dB / db
-10
C1 R5
one RC network to R3 -12
3.18k 10k
1n
high-pass. You R4
5k
-14
-16
40 50 60 70
high and low Fig. 13-16: Sallen & Key band- Fig. 13-17: Second-order
frequencies. pass filter. band-pass filter response.
The filter of figure 13-18 has two Twin-T stages. The first stage is
a second-order low-pass notch
configuration, the second stage is 10
-10
bandwidth 2kHz. Just outside the
bandwidth the attenuation reaches -20
Frequency/kHertz 2kHertz/div
It must be clear to you by
now that active filters are costly. Fig. 13-19: Response of a fourth-order
elliptic band-pass filter.
Not only do they require precision
components, but the values of most
capacitors and some of the resistors are such that they cannot be integrated.
A fourth-order low-pass or high-pass filter requires at least eight external
components and five pins. For a band-pass filter with only modest
performance 14 external components and pins are needed.
Q = C ∗V = I ∗ t
Fig. 13-20: Making a resistor
out of a capacitor by switching
i.e. the charge in a capacitor (in Coulombs) at a rapid rate.
is given by either the capacitance times the
voltage or the current flowing into the capacitor for a certain period of time.
In the case of figure 13-20, the current flowing between the two terminals
over one period is
C R ∗ (V 1 − V 2)
I= = C R ∗ (V 1 − V 2)∗ f clock
t clock
(V 1 − V 2)
I=
R
1
R=
C R ∗ f clock
1
R= = 2∗10 6 = 2 MegOhms
10 ∗5∗10 −12
5
1
f 3dB = S1 S2
2∗ π ∗ R∗ C CR C
f clock ∗ C R C
f 3dB =
2∗ π ∗ C
greatly reduce the values of the capacitors and then replace the resistors
with a capacitor and switches.
The switched-capacitor filter requires lateral switches, which are
easily implemented in CMOS, but cumbersome (and slow) in a bipolar
process. For this reason, this approach has become exclusively CMOS
territory.
To minimize the influence of stray Ph1
CR
Ph2
1. No matter how carefully you design the switches, there is always some
switching noise.
4. The output has sampled noise, which is present even if the input is zero.
14 Power
Linear Regulators
Let's say you have 12 Volts available but need 3.3. Your 3.3-Volt
load consumes up to 500mA. The 12-Volt source (e.g. a car battery)
fluctuates between 10 and 14 Volts; the lower voltage needs to be within
5%.
The immediate choice to effect this change in voltage is a linear
regulator. Look at it as a variable resistor, dropping whatever voltage is not
needed.
Q1
Vcc The unwanted
Q9
voltage is dropped in
Q12 an NPN transistor. In
Q2
Repi Q7
C1
Q13 figure 14-1 this is a
50
10p
Vreg
Darlington
Q15
Q14
Q10
R3 configuration to
26.25k
R1
Q5 Q6
minimize the drive
3k
Q3
R4 current; it requires at
Q16 15k
Q4 Q8 least 2.2 Volts
R2
Vref
difference between
6k Q11
1.2 SUB
Vcc and Vreg, but it
is an easy and simple
design.
Fig. 14-1: Linear regulator with NPN power stage. The regulator
uses a 1.2-Volt
bandgap reference
(see chapter 7), whose voltage is compared with a fraction of the regulated
output by the differential amplifier Q5, Q6, Q7 and Q10. Once the circuit is
in balance the voltages at the bases of Q5 and Q6 are equal, so the regulated
voltage is:
Vref ∗ ( R3 + R4)
Vreg =
R4
Output Voltage / V
2
drop between supply and output.
For this reason such a circuit is 1.5
through the load also flows Fig. 14-2: Drop-out voltage of NPN
through the output transistor. So, at regulator.
500mA, the load consumes 1.65
Watts, the regulator 4.36 Watts
4.5
(with 12-Volts in), which is
4
simply converted into heat.
3.5
Output Transistor This the main disadvantage of
Power Dissipation / W
3
a linear regulator. The heat is
2.5
produced mainly by one
2
device: Q13. Thus there will
1.5
Load be a hot-spot on the chip and
1
0.5
resulting temperature
00
gradients, even with an
2 4 6 8 10 12
adequate heat-sink. These
Supply Voltage/V 2V/div
temperature gradients are
Fig. 14-3: In a linear regulator the energy not bound to influence other
required by the load is converted into heat. circuitry on the chip, including
the regulator's own reference.
A linear regulator with an NPN output transistor is relatively easy to
compensate. Despite the fact that the loop gain is high (which results in an
output impedance of a mere 4mOhm) the circuit is rendered stable with a
-30
3.34 -40
No Output Capacitance
-50
-60
3.3 -70
-80
3.28
-90
100uF
3.26
-100
0 1 2 3 4 5 6 7 8 9
1k 2k 4k 10k 20k 40k 100k 400k 1M 2M 4M 10M
Time/µSecs 1µSecs/div
Frequency / Hertz
Fig. 14-4: The regulator is stable, even Fig. 14-5: Power supply rejection
with a filter capacitor at the output. with and without a filter capacitor.
100 100
together with Rext form a zero at
dB / db
80 80
-20 -201
degrees.
10 100 1k 10k 100k 1M
The external capacitor
Frequency / Hertz
frequencies above about 5kHz the 1m 10m 100m 1 10 100 1k 10k 100k 1M 10M
High Currents in an IC
There are two factors which limit how much current an IC can carry. The
first is electro-migration. The force of huge numbers of electrons rushing through a
conductor can become so large that the electrons begin to move atoms, physically
push them along. For pure aluminum this happens at about 500'000 Amperes/cm2.
The effect is slow, it may take months, but eventually there will be an area where
there is no aluminum left. Electro-migration is aggravated by high temperature and
depends on the composition and grain structure of the metal.
Half a million Amperes may seem large and safe, but when you consider that
you are dealing with very thin layers, the limitation becomes real. For example, for a
thickness of 10'000 Angstroms (10'000Å = 1um) and a width of 1um a current density
of 500'000A/cm2 is reached with just 5mA.
The second limitation is resistance. Pure aluminum has a resistivity of
2.8uΩcm. Thus a layer 1um thick has a sheet resistance of 28mOhms/square. Make
this run 100um long and you have a resistance of 2.8Ohms.
Let's say you want to carry 1 Ampere over a distance of 1000um on a chip.
With a thickness of 1um, the aluminum stripe would have to be at least 200um wide
to avoid electro-migration. It then would have a resistance of 140mOhms, i.e. drop
0.14 Volts.
And don't forget to check how much current contacts and vias can take in
your process, as well as the thickness required for bonding wires.
Vdd
Figure 14-11 shows
M5 M6
M7 a CMOS version of figure
M9
W=20u W=20u 14-7. Also dimensioned
L=1u L=1u W=20u
I1 L=1u W=20u for 20mA, the P-channel
L=0.5u Vreg
20u
C1 M=20 output device is smaller
0.1p R1
than the previous lateral
M2 M4
10k PNP transistor. A low
Cext dropout voltage is,
1.2
W=1u
L=1u
W=1u
L=1u
10u however only present at
M8
R2
low current; to get the
M1
Vref
M3 W=10u
20k Rext
1
same value at 20mA, M9
L=1u would need to be 20 times
W=10u
L=1u
W=10u
L=1u
the indicated size, or a total
width of 8000um.
Fig. 14-11: Low-drop out CMOS regulator. 1.8
1.7
1mA
The circuit was designed for a 1.6
10mA
regulated output of 1.8V. 1.4
1.1
Again an external capacitor with a resistor
11
in series is necessary at the output to 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
Vdd/V 200mV/div
create a zero and turn the phase up.
Fig. 14-12: Drop-out voltage.
Y2 Y1
180 180
-40
160 160
-45
140 140
Power Supply Rejection / dB
Phase -50
120 120
-55
Phase / degrees
100 100
dB / db
-60
80 80
-65
60 60
40 40
Gain -70
20 20 -75
0 0 -80
-20 1k 10k 100k 1M 10M 100M
10 100 1k 10k 100k 1M 10M 100M
Fig. 14-13: An adequate phase margin is Fig. 14-14: ...but Rext limits power
achieved with Rext .... supply rejection at high frequency.
current.
With the chosen
device for Q6, the maximum Fig. 14-15: Low drop-out regulator with external
current is around 500mA. At PNP transistor.
3.3
this level the supply voltage can drop to
3.25 10mA within 200mV of the output (3.3V).
3.2 500mA
100mA
There is, however, a compromise
3.15
in the loop gain, which effects the output
Vreg / V
3.1
3.05
impedance (33mOhm); in order to achieve
3 stability, the gain has to be reduced (R1,
2.95
R2). Also, the same scheme as used in the
2.9
2.85
two previous circuits is required: an
3 3.2 3.4 3.6 3.8 4
output capacitor with a resistor in series to
Fig. 14-16: Drop-out voltage.
VP/V 200mV/div
keep the phase from reaching zero before
the gain does.
Y2 Y1
180 180 -20
160 160
-30
140 140
Power Supply Rejection / dB
100 100
Gain / db
-50
80 80
-60
60 60
40 40 -70
Gain
20 20
-80
0 0
Fig. 14-17: Phase margin can only be Fig. 14-18: ... which again impairs
kept high by a resistor in series with power supply rejection at high
the output capacitor .... frequency.
Switching Regulators
Assume again that you have a supply voltage of 12 Volts, but you
need 3.3 Volts. Your load consumes 1 Ampere.
A linear regulator acts as a resistor which drops the unneeded 8.7
Volts. In the process it converts 8.7 Watts into heat. 3.3 Watts are used by
the load; a rather dismal efficiency.
Enter the switching regulator: instead of creating a resistance
between input and output, it connects an inductor between the two for short
periods of time.
Vcc The switch, S1, is driven
S1
PWM by a pulse generator (PWM, or
pulse-width modulator). The
L1
Out pulses are rapid, so that the
27u C1
RLoad inductor value can be small.
D1
4.7u
10 The inductor, together
with C1, smoothes out the
switching pulses.
Fig. 14-19: Reducing a supply voltage with a
series switch and inductor. When the switch is
closed, the left node of the
inductor is at Vcc (assuming the
switch has no resistance), but when the switch opens, this voltage jumps
abruptly to a large negative value, created by the energy stored in the
inductor. It is the purpose of D1 to catch this negative spike so it does no
harm to the switch and provides a path for the current during the off period.
Figure 14-20 shows the 8
4
remains of the switching frequency
20
(100kHz). 3
But when you add some resistance to the switch and inductor and a forward
voltage drop for the diode, the efficiency drops. For example, with a total
resistance of just 50mΩ
Vcc
and a diode drop of
S1
0.3Volts (a Schottky
L1 diode) the efficiency is
Vreg
D1 27u
R1
94%.
Error Amp.
32k The circuit of
RLoad figure 14-19 is not a
R3 R2 4.7u 10
100kHz 100k 27k C1 regulator; we have to add
47n feedback to make the
C2 Vref
Triangle
1.2
output voltage immune to
supply fluctuations. This
is accomplished by
Fig. 14-21: "Buck" regulator.
amplifying the difference
between a fraction of the
output voltage (R1, R2)
and a reference voltage in an error amplifier. S1, an abstract simulation
symbol, is now used as both a switch and a comparator (with the on/off
thresholds set just a few millivolts apart). The output of the low-pass filter
(R3, C2) following the error amplifier is thus compared with a triangle
wave (100kHz, 2Vpp). In this way the regulator finds the duty cycle which
gives the desired output voltage. Such a circuit is generally called a Buck
Regulator.
There are a few items to 3
consider, which are peculiar to a
switching regulator: 2.5
charge and discharge 1nF. If you push the switching frequency to 500kHz,
this current increases to 0.5 Amperes.
Second, the current level that the switching transistor needs to
handle is always larger than the average output current. If you use a small
inductor, the peak current can exceed the average by a factor of three or
more; with a large inductance this factor is between 1.1 and 1.4.
Third, the voltage drop (and switching speed) of the diode is just as
important as that of the switching transistor, their peak currents are roughly
equal.
Fourth, the output LC filter (L1, C1) form a pole, which makes
frequency compensation (R3, C2) more challenging.
V ∗t
I=
L
Fig. 14-23: By using inductive charge the
As soon as the switch is output voltage can be made higher than
turned off, a positive voltage appears the supply voltage.
at the anode of the diode, created by
the stored current. This voltage is averaged by C1.
4 300
3.5 250
70 Percent Switch Current
3
200
D1-cathode / V
2.5
50 Percent
mA
150
2 Output Current
100
30 Percent
1.5
50
1
0
0 1 2 3 4 5
1.435 1.44 1.445 1.45 1.455 1.46
Time/mSecs 1mSecs/div
Time/mSecs 5µSecs/div
Fig. 14-24: Output voltage for three Fig. 14-25: Currents through switch and
different duty cycles. load.
1.5
an overshoot.
1
5 22x29 1
20 30x37 1.7
40 52x70 5.7
100 152x220 52.4
R5
2.7k emitter length some 200
200
Q6
times that of a minimum
Q9
R6
5k
10
geometry transistor. Speaker
Input
Q1 Q2
R1 Ideally we would
Speaker
R2
1K
want one of the two output
29k 8
Q7
devices to be a PNP
Q11
transistor, to exploit the
Q12
10
I1 I2
R3
10k
complementary nature of
200
Q14
500u 1m
the "push-pull" output. But
Q13
SUBNPN transistors carry a
much higher current than
-12V
Gain / dB
compound PNP transistor Q11 to Q14, 0
100m
an ideal Class B amplifier is 76%.
10m
For this circuit, with its 2-Volt drop
in each output device, the maximum
1m efficiency amounts to 62%. Thus
the output transistors produce 1.7
0.5 1 1.5 2 2.5 3 3.5 4 4.5
Watts of heat each (for a 5.6 Watt
Frequency/kHertz 500Hertz/div
output).
Fig. 14-31: Spectrum of the output signal
at full power. 70
60
+24V
A better solution is the Bridge
Output. In essence there are two
Output
amplifiers, 180 degrees out of phase. With
R1
no input signal, both output rest at 1/2 Vcc.
R2
As the signal appears, one output moves up,
Speaker
8
the other one down.
Input
In this configuration we have in fact
doubled the output swing. With the same
R4 Output
total supply voltage, 25 Watts of output are
generated (which requires four output
R3
transistors with a capability of 2.5A each).
Efficiency is unchanged at 62%, which
Fig. 14-33: Class AB amplifier with produces a power dissipation of 15.3 Watts.
bridge output.
-10V 4
2
Fig. 14-34: Bidirectional
V
0
switching arrangement. Output Voltage
-2
-4
To start with, let's use
-6
two power supplies. The two -8
switches connect the inductor -10
assume that there is no dead Fig. 14-35: Switching and output wave-
time or overlap and that this forms.
switching action is instantaneous.
The value of the inductor is fairly large for the chosen switching
frequency (200kHz); it is never fully charged or fully discharged. Despite
this, there is still a substantial ripple at the output.
The average output voltage
900 is a function of the duty cycle. At
800 50% the output is zero; 75%
700
produces +5 Volts and 100% +10
Inductor Current / mA
600
Volts. Duty cycles of less than
500
50% cause the output to be
400
negative.
300
200
Notice that the current
100
gradually builds up (figure 14-36);
00 5 10 15 20 25 30
the time constant of this effect is
Time/µSecs 5µSecs/div
given by L1 and the 8-Ohm load (a
speaker), a factor which will
Fig. 14-36: Current through S1. become important when we close
the loop with feedback.
Let's now take the next step
+10V and modulate the duty cycle with a
S1 sine-wave signal, making a Class
D amplifier. As in the switching
Input
30u
Speake r regulators, the switch symbols also
act as comparators (i.e. the
L1
270n RLoad thresholds of the control terminals
S2
Triangle C1 8
are set so that the switches turn
from off to on (and from on to off)
within a few millivolts. Also (for
-10V
4
delay and insignificant resistance.
Output Voltage / V
-4
but increases the build-up delay -6
mentioned above. -8
10 1
1 100m
Output Spectrum / V
Output Spectrum / V
100m 10m
10m 1m
1m 100µ
0.5 1 1.5 2 2.5 3 3.5 4 4.5 150 200 250 300 350 400 450
Fig. 14-39: Frequency spectrum in the Fig. 14-40: Frequency spectrum in the
signal range. switching range.
level of distortion can only be brought down with feedback. And that is
somewhat of a problem.
In order to reduce the
Y2 Y1
0
high-frequency components at
0 the output we used an LC filter.
-20
Phase It is dimensioned to be effective
-2
-40 at 200kHz, but it causes a phase
Phase / degrees
Amplitude / db
-60
-4 shift already in the audio range.
Amplitude Because of this, the amount of
-6
-80
feedback possible, using a single
-100 -8 feedback loop, is limited to
about 20dB. With two or even
-120 1k 2k 4k 10k 20k 40k 100k
three nested feedback loops this
Frequency / Hertz
figure increases to about 35dB.
Fig. 14-43: Amplitude and phase response of
Also, a loud-speaker is
the output filter. not really a simple resistor, there
is some inductance as well,
making the phase relationship in the feedback path even more complicated.
We could, of course, increase the switching frequency, which would
allow us to push the cutoff frequency of L1 and C1 higher, but the penalty
would be lower efficiency and an increase in drive requirements for the
switching transistors.
We have been assuming that we want a faithful (albeit larger)
reproduction of the input signal at the load. Strictly speaking, this is not
really true. In the case of an audio amplifier, the human ear cannot hear
200kHz, so filtering out high frequencies makes little difference. If the
application is a servo amplifier, the load is unlikely to respond to such rapid
fluctuations.
But there is radiation. Do we want to connect a square-wave of
200kHz (and its harmonics) across a long speaker cable and let it radiate
into AM receivers and other electronic equipment? The answer is a clear
no, and rules and regulations limiting such radiation have been written.
There are ways to reduce radiation. First, we can keep the speaker
wires short, moving the amplifier next to the speaker. Second we can vary
the switching frequency in a random fashion, creating a spread spectrum.
Although this does not reduce the total radiation, it at least makes it less
noticeable and allows meeting radiation limits.
either connected to +V on S1
D1 D4
S3
the left side and -V on the
right, or vice versa. This Input Output
L1 L2
output
RLoad
effectively doubles the 15u 8 15u
+10V
But now
let's change the
S1 S3
D1 D4 circuit a little.
Input Output Output Instead of having
L1
RLoad
L2
the two outputs
8
15u
C1 C2
15u
move in opposite
S2
270n 270n
S4 direction, invert one
D2 D3 of the drives so that
-10V they move up an
down together. If
Imverter the input signal is
Triangle
zero, the two
outputs will move
at exactly the same
Fig. 14-45: Class D amplifier which suppresses the
time. Each output
fundamental of the switching frequency. then carries a
200kHz square-
wave, but between them there is no signal. As the input signal goes
positive, the duty-cycle of one output increases while the duty-cycle of the
other output decreases by the same amount. Thus, between the two outputs,
there is now a square-wave with a duty cycle amounting to the difference.
10
The effect on the
frequency spectrum is quite
1 drastic: the fundamental of the
switching frequency has
Output Spectrum / V
15 A to D and D to A
resistors.
Note that the full reference voltage does not get to the output. This
quirk is caused by the fact that a string of eight resistors has nine nodes.
We need to include zero, so three bits only reaches 7/8 of the total voltage.
To represent bipolar values some special codes are used. In the
"sign + magnitude" code a bit is added which represents just the polarity.
This is not the most efficient way and it is somewhat awkward (there are
two values for zero, 0000 and 1000).
The offset binary code Sign + Offset Twos
simply starts at the most Number Magnitude Binary Complement
negative number and counts up. +7 0111 1111 0111
Note that there is only one +6 0110 1110 0110
+5 0101 1101 0101
value for zero, but the full +4 0100 1100 0100
reference voltage is still not +3 0011 1011 0011
+2 0010 1010 0010
present. +1 0001 1001 0001
In the twos complement 0 0000 1000 0000
-1 1001 0111 1111
code, positive numbers are the -2 1010 0110 1110
same as in a binary code, with -3 1011 0101 1101
-4 1100 0100 1100
the additional sign bit. -5 1101 0011 1011
-6 1110 0010 1010
Negative numbers are the -7 1111 0001 1001
inverse or complement of the -8 0000 1000
positive ones, with a 1 added Fig. 15-2: Codes representing bipolar
and the sign-bit changed. values.
Other codes are also
used for DACs. In the BCD (binary coded decimal) code each decimal
digit is represented by four binary digits. BCD is primarily used for digital
voltmeters. The Gray Code changes only one bit at a time, a feature useful
in shaft encoders.
In any DAC, the output is strictly proportional to the reference
voltage; double it and the output will double. Thus if you treat the reference
terminal as an input, you have what is known as a multiplying DAC.
delivered to the output. Notice that in this approach, too, the top of R9 is
not connected; a seventh bit would be required to allow that.
In such a segmented DAC the resistors in the first string (the most
significant bits) are the most critical. They should be largest in size to
obtain the best matching (or be trimmed).
It is crucial that a DAC be monotonic, i.e. as you step through the
code from low to high, the output always increases (it may not increase by
precisely the same amount, but at least it will never decrease). A divider
DAC is always monotonic. The same holds true for the segmented DAC,
provided each segment is monotonic (which is the case if we use dividers).
Vref +
+
R1 -
R5 -
R9
+ Analog Out
R2
R6 R10 -
R3
R7 R11
R4
R8
+ R12
+
-
-
Vref
Analog Out
MSB LSB
1 0 0 1 1 0
resistors with 1R
-Vref
identical segments M1
1/2 R
to 63 resistors, one 4R Analog Output
less than a simple M3
resistor string. In M4
8R
-
addition, the resistor +
16R
for the most M5
amplifier, so the
output goes R
negative with a
positive reference -
voltage. +
Analog Out
Using
MSB LSB only two resistor
Fig. 15-5: DAC with R-2R ladder.
values improves
matching and
trimming is
easier. Note, however, that the MOS transistors carry current and their
resistance is critical. Moving from left to right, the current drops by 50% in
each stage. Thus, to get the smallest error, the transistor size should be
doubled for each stage moving from right to left.
In many DACs the analog voltage is created not by voltage taps but
by currents. An example of a much simplified current DAC is shown in
figure 15-6. A primary current is generated by R10 from a reference
voltage; with Vref at 1 Volts, this current amounts to 200uA. Q1 through
Q6, being biased from Q1, each produce a fraction of this current, Q2
+V 100uA, Q3 50uA, Q4
R11
10k 25uA and Q5 12.5uA.
Vref Analog Out
Q6 is used to terminate
R10
5k the ladder.
S w i t c h e s These binary-
weighted currents are
+
Q1 Q2 Q3 Q4 Q5 Q6 then switched to either
-
16 8 4
R11 or +V (by, for
example, a differential
R1
1k
R2
2k
R4
2k
R6
2k
R8
2k
R9
2k pair, acting as logic
-V
R3 R5 R7 inputs). Note that the
1k 1k 1k
currents are flowing all
Fig. 15-6: Bipolar current DAC with R-2R ladder. the time, which makes
this a fairly power-
hungry approach.
You can, of course use the current directly as the output. But note
that, if R10 and R11 are both inside the IC, their temperature coefficients
and absolute variations will cancel.
Don't let this simplified example mislead you; there are many
sources of error which require fine attention to detail. In a bipolar circuit
there are base currents which must be compensated lest they subtract as
much as 1% from the ideal values of the binary (collector) currents. And
each collector must be at exactly the same potential as Q1 (here ground).
Lastly, if bipolar
+V
switches are used, they, R11
10k
too, will have base Analog Out
Vref
currents which also R10
must be rendered 5k
S w i t c h e s
Although it +
M1
16
M2
8
M3
4
M4
2
M5
1
M6
1
started out that way, a -
R1 R2 R4 R6 R8 R9
current DAC is no 1k 2k 2k 2k 2k 2k
longer primarily a -V
R3
1k
R5
1k
R7
1k
CMOS has some significant advantages here. There are no base currents,
therefore no base current errors. However, the drain voltages still need to
be at identical levels, otherwise the Early effect will cause substantial
deviation.
S w i t ch e s
M1 M2 M3 M4 M5 M6
+
16 8 4 2 1 1
-
R1 R2 R4 R6 R8 R9
1k 2k 2k 2k 2k 2k
R3 R5 R7
-V
1k 1k 1k
the last transistor is only used for terminating the resistor ladder. It carries
the same current as the least significant bit, 12.5uA. Use it and split it into
16 equal parts, 8 used for 5th bit, 4 for the 6th, 2 for the 7th and 1 each for
the 8th bit and a dummy transistor.
Simple transistor ratios are used for this extension. We could have
also employed another R-2R resistor ladder but, since these are the least
significant bits, the accuracy is most likely sufficient.
R6
reference voltage with an equal but
+ negative value.
- All comparators operate
R7
+
simultaneously, so the speed of the
- converter is given by the speed of the
R8
+
comparators alone.
-
The disadvantage of this approach is
again complexity with large number of bits,
Fig. 15-9: Divider ADC.
with an accompanying high power
consumption. At eight bits 512 resistors
and comparators are required (assuming a bipolar input); at 12 bits the
number increases to 8192. Also note that all comparator inputs are in
parallel, which makes for a rather large input capacitance.
The Successive
Approximation ADC DAC
Vref
In
the number of resistors
+
Sample
-
and Hold
remains unchanged (Figure
15-10). Here a sample of Encoder
DAC. The control logic sets the DAC through a register to a likely value
initially. If the value is too high, the register is moved down; if the initial
guess is too low, the register is moved up. After a few steps the correct
setting is found and the conversion stops.
All this guessing and stepping takes time, thus a successive
approximation ADC is considerably slower than the divider approach,
though it consumes less current and takes up a smaller area.
In both approaches the accuracy is limited by the resistor (or
capacitor) divider, as was pointed out in the DAC section.
C1
R1
50k 10p
Analog In
R2 - Comp.
Pulse
50k
Train
+
Timer
Integrator
S1
Bias
1usec
0
Vref
2
falling flanks of the triangle wave are equally affected and thus cancel out).
A high loop gain in the integrator op-amp assures that the voltage
fluctuation at its input is down to a few microvolts (but its offset voltage
still matters).
The performance of a delta-sigma ADC can be improved by adding
one (or more) feedback
loops (Figure 15-13,
shown again in the
conventional way). Be
aware, however, that
the higher the order, the
less stable the design
becomes. A third-order
delta-sigma ADC can
oscillate in some Fig. 15-13: Second-order Delta-Sigma ADC.
unexpected ways.
The significant advantage of the delta-sigma ADC is its capacity for
resolution. These circuits are often called "oversampling ADCs", because
they must sample the incoming waveform above the Nyquist rate (twice the
maximum input frequency). For example, if you want to capture a 1kHz
signal with an 8-bit resolution, the maximum frequency must be at least
2kHz times 256, or 512kHz. At 12 bits this frequency increases to 8.2MHz.
However, this presumes that we do nothing else than counting
pulses at the output, which ignores much of the concept's capability. The
delta-sigma ADC was made for CMOS and with today's small geometries a
great deal of digital signal processing can be done once the pulses exist.
Apart from increasing the resolution, the sampling noise (which is already
centered around a rather high clock frequency) can be brought down to
stunningly low levels with a sophisticated digital low-pass filter (called a
"decimation" filter).
With these additional measures, a second-order delta-sigma ADC
with a signal bandwidth of 4kHz can achieve a 14-bit resolution with 85dB
signal to noise ratio, using a clock frequency of 1MHz.
In this, the last circuit design chapter, we look at six functions which
did not fit well into the previous subjects.
As pointed out before, be aware that you will need to re-simulate
these circuits with models specific to the process to be used.
Q3
the inputs for the next
Q2 Q2 Q3
cell. Each subsequent
cell is biased at a
I2 I1 I3
50u 1m 50u
I2 I1 I3 higher DC potential to
50u
50u 1m
avoid saturating any of
the transistors.
In both forms
Fig. 16-2: Second form of Fig. 16-3: And the third there is a small error
the Gilbert cell. form of the Gilbert cell. due to the base
34
Fig. 16-4: A practical application of
the Gilbert cell. 32
With Gilbert Cell
Differential Gain / dB
30
This problem is made
Differential Pair Only
worse by stacking several cells, 28
Multipliers
+V
We have seen a similar circuit
R3 R4
60k 60k before, used as a phase detector for a
Out PLL. While accuracy in that application
Out was of minor importance, in a multiplier
it is the main feature,
V2 Q3 Q4 Q5 Q6
The circuit requires a split power
R5
10k
R6
10k
R7
10k
R8
10k
supply (e.g. ± 5 Volts), so that at least
one input (V1) can be at ground level.
The second input (V2) is biased safely
V1 Q1 2.5V
Q2
higher (2.5V) to avoid saturating Q1
and Q2.
R1 R2 It is the insertion of resistors in
10k 10k
the emitters of all six transistors that
I1
100u SUB gives this multiplier its accuracy. Their
-V
values need to be large compared to the
dynamic emitter resistance (re, see page
Fig. 16-6: Simple four-quadrant
multiplier. 4-1).
Such a circuit is 10
called a four-quadrant 8
V2 = 100mV
multiplier because it 6
R5 R6
Accuracy is unchanged for
9k 9k untrimmed operation; with
100u
I1
100u
I2
100u
I3
100u
I4
trimmed thin-film resistors
SUB
-5V
and additional temperature
Fig. 16-8: Four-quadrant multiplier with both input
compensation (see reference)
voltages at ground level. such a circuit can be brought
to within 0.1%.
Figure 16-9 +1.5V
to ± 100mV.
This circuit M1 M2 M3 M4
illustrates the V1 V2
can, of course, change the resistor ratios so that the relationship between
inputs and output is multiplied by a constant.
Peak Detectors
currents of the Q2 Q5 R3
inner ones. This In
10k
R2
limits the Q1 Q6
Peak
20k
temperature range D1
Q9
(to about 100oC)
C1
but lowers the Q7 Q8 R1
10Meg
100n
input current. SUB
Notice the
discharge resistor
R1. Without it, Fig. 16-11: Single-supply peak detector with a lower input
current.
the base current of
Q6 would charge C1 (it flows out of the base). Thus, in this circuit, a
controlled rate of discharge through R1 is essential.
+1.8V
In both examples the supply
I2 voltage must be lower than the
10u M6 emitter-base breakdown voltage of
I1
the output transistor. If it is to be
10u
W=10u
L=0.35u
higher, use an additional diode in
R1 series with the emitter.
1k
In M1 M2 CMOS devices are much
Peak better suited for peak detector
W=10u
L=2u
W=10u
L=2u M5 design than bipolar ones for two
C1 reasons: 1) there is no (DC) input
M3 M4
W=5u
L=5u
100n current and 2) you can reset a
M=2
capacitor to zero volts (the
W=5u
L=5u
W=5u
L=5u
collector-emitter voltage of bipolar
transistors does not go to zero, there
is always a remaining voltage of
Fig. 16-12: CMOS peak detector. about 100 or 150mV).
As in the bipolar examples,
the feedback loop is compensated with a resistor in the output path, working
together with Cext.
Since there is no input current, C1 can be made quite small, to the
point where it can be internal. But be aware that the smaller C1 the more
difficult it becomes to compensate the feedback loop.
accuracy. Specifically Q2 Q5
Out
designed circuits avoid this -
AC Q1 Q6
and require only a single C1
Q9
R1
1 2k
0.6 Q7 Q8
0.2
In / V
SUB
-0.2
-0.6
-1
Fig. 16-14: Bipolar half-wave rectifier.
1.8
1.4
In the circuit of figure 16-14
Out / V
1
0.8 the output is at ground (without an
0.4
input signal), held there by R1. If the
00 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
input moves above ground, the output
Time/mSecs 200µSecs/div
follows. But if the input goes
Fig. 16-15: Input and output waveforms. negative, there is nothing in the output
stage that can pull it below ground, so
it just stays there.
The value of R1 must be low enough to keep the voltage drop due to
the base current of Q6 low. This resistor cannot be replaced with a current
sink; the minimum collector-emitter voltage of an NPN transistor is too
high.
You can capacitively couple the input signal, with a resistor
connected from the input terminal to ground to provide a dc path.
Minimum required supply voltage for a 1-Volt input range is 3.5V.
In figure 16-16
an inverting op-amp R1
10k
Vcc
-0
-0.2 frequency compensation of the
-0.4
-0.6 op-amp becomes very difficult.
-1 Minimum supply voltage
1.8
1.6
for a 1Vp input is 2 Volts.
1.4
Out / V
1.2
0.8
Both of these circuits can
0.6
0.4
be readily translated into
00
CMOS. The half-wave rectifier
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
in figure 16-18 uses two
Time/mSecs
Fig. 16-17: Input and output waveforms200µSecs/div
for the advantages of CMOS: first,
full-wave rectifier.
there is no base current, so the
output pull-down impedance can be quite high; second, a current sink can
be used at the output instead of a resistor.
With a 1-Volt input range this circuit works down to 1.8V supply at
-40 C, 1.6V at 0oC.
o
I1 I2 I3 I4
10u 10u 10u 10u
M8
M2 M3
1p W=10u
W=10u W=10u C1 L=0.35u
AC M1 L=0.35u L=0.35u M6 Out
W=10u W=10u
L=0.35u L=0.35u
M7 M9
M4 M5
W=5u W=10u
W=5u W=5u L=2u L=0.35u
L=2u L=2u M=2
+3.3V
I2 I4
10u
10u
M6
C1
W=10u
L=0.35u
M1 M2 1p
R1
20k
W=10u W=10u
L=0.35u L=0.35u M5 Out
M3 M4 R2 W=5u
20k L=0.35u
M=2
W=5u W=5u
L=0.35u L=0.35u
AC
Averaging the obtained rectified fundamentally takes time; the longer the
time constant, the smaller
the ripple (but there will
700
always be a ripple).
600
Full-Wave
Take the most
Rectifier simple approach, a single
500
low-pass filter (RC)
400 connected to the output. A
mV
Thermometers
k∗ T A1∗ I 2
deltaVBE = ∗ ln
q A2∗ I1
I1 and I2 are identical (as produced by Q5, Q6, R3 and R4) and the
area ratio is 24. Thus deltaVBE amounts to roughly 83mV at 300K. Since
T is in Kelvin, the current increases linearly from zero at absolute zero to
22uA at 300K. This current is then mirrored by Q7 and causes a voltage
drop across R2. With the values chosen (and perfect matching), the output
voltage amounts to 1mV per Kelvin. Note that any temperature coefficient
or absolute variation in R1 is eliminated by a matching R2.
Although the design is relatively insensitive to power supply
variation, accuracy is maximized by powering the thermometer from a
reference voltage.
Matching is the all-important factor here. A ±1% resistor matching
variation will result in an error of ±3oC at room temperature. Adding
mismatching of VBE and hFE, you must expect a variation of up to ±5oC
untrimmed. With trimming an accuracy of ±0.5oC is possible.
2.5V
room temperature. You can improve this
M3 M5 M7
by using a larger emitter ratio for Q2/Q1
W=20u W=20u W=20u and generally larger devices.
L=5u L=5u L=5u
Zero-Crossing Detectors
waveform.
V
1.5
The effect used to create the window does not work well in CMOS.
Instead we can employ two comparators, one biased at about +200mV (M9,
M10) and the other at ground (figure 16-27). Their outputs are then
supplied to an "and" gate (M17, M18), which drives the output.
The AC waveform is clamped in the positive direction by two
"diode-connected" transistors (M1, M2) and in the negative direction by a
substrate diode.
Since there is no base current, you could theoretically make Rext.
very large. But also consider that the devices connected to the input
(including the pad and the ESD protection device) have a small amount of
capacitance. The time constant formed by the external resistor and this
capacitance must be smaller than the desired pulse-width.
1.8V
M3 M4 M8 M12 M14 M16
W=2u
M9 M10 L=0.5u
17 Layout
The layout of analog ICs has so far remained an art, there are no
computer programs which could design, place and route the components in
an intelligent, competent way. And, more often than not, the person who
created the circuit diagram needs to (or should) get involved.
This chapter is by no means a complete guide; it would take an
entire book to do the subject justice. Look at it as some hints stemming
from practical experience.
Bipolar Transistors
almost always requires wider metal runs for both the emitter and collector
(see box on page 14-5).
An alternate design is
shown in figure 17-8, with
the uniform contact openings
required by dense processes.
There are two collector
contacts (on the outside),
three emitters and four base
contact columns.
If you increase the
size of such a transistor
further, there comes a point
where it is of advantage to
taper both the emitter and
collector metal, gradually
increasing the width as the
Fig. 17-8: Alternate high-current design. currents from more and more
contacts are added.
voltage). This field plate improves the gain of the transistor at low current
by keeping p-type charges away from the surface. (In a CMOS process, the
poly layer is used as a field plate, also connected to the emitter).
Although a circular emitter results in a uniform base-width and thus
(theoretically) produces the highest possible gain, there is actually very little
enhancement over the more simple square one.
The third terminal (at the bottom) is the base contact, identical to a
collector contact for an NPN transistor.
For a lateral PNP transistor (in a bipolar process) the presence of a
buried layer is essential. Without it, the substrate (connected to the most
negative supply) would be just as attractive a collector as the intended one;
i.e. about half the emitter current would flow to the collector, the other half
to the substrate.
The dual pattern at the bottom of figure 17-9 should be avoided. It
looks attractive, especially since lateral PNP transistors often have common
bases (e.g. in current mirrors), but the two devices influence each other,
especially in saturation.
Resistors
be small enough to ignore for resistors with 200 Ohms/square (about 0.2%),
but for implanted resistors this error is almost always significant. Note that
this is an initial error only; it is not subject to change during production.
To avoid this effect, you can place each resistor in its own tub and
connect the tub to the positive end of the resistor.
CMOS Transistors
The third belief holds that you should add dummy devices at
the periphery. There appear to be two different explanations to justify
this practice: 1. shadows or reflections during exposure act differently
on the remote edges than on devices in close proximity, or: 2. the etch-
rate for wide spaces is different from narrow ones.
I found no difference between groups of resistors with and
without dummy devices at the periphery. It appears that you might be
better off using the extra space to make the devices larger.
Cross-Unders
Kelvin Connections
To fit into small, shallow packages, wafers are often thinned down
by back-lapping (a somewhat messy, wet grinding operation). This
removes not only the oxide layer on the back but any diffusions which may
have taken place there, giving direct access to the substrate material. If you
add a gold-plating step you get a low-resistance connection directly to the
substrate.
Ordinarily such a connection is not essential (the substrate is also
contacted from the top). But if you have sinned and allowed high substrate
currents, you may be able to suppress the resulting effects in this way.
* * *
Thus ends this book of the minority field in the world of semiconductors. A
field past glamour, often neglected, but undeniably essential. And a field of
great satisfaction for those who know it.
Chapter 1
Shockley, W.: "The Invention of the Transistor", National Bureau of Standards Publication
# 388, May 1974
Shockley, William: "The Path to the Conception of the Junction Transistor", IEEE
Transactions on Electron Devices, July 1976, pp. 597-620
Brattain, Walter H.: "Genesis of the Transistor", The Physics Teacher, March 1968, pp.
109-114
Wolff, Michael F.: "The Genesis of the Integrated Circuit", IEEE Spectrum, August 1976,
pp. 45-53
Interviews with Phil Ferguson, Victor Grinich, Jean Hoerni, Eugene Kleiner and Robert
Noyce, 1983
Transistor Design:
Muller, Richard and Kamins, Theodore: "Device Electronics for Integrated Circuits", John
Wiley and Sons, 1977
Chapter 2
Kundert, Kenneth: "The Designer's Guide to Spice and Spectre", Kluwer Academic
Publishers, 1995
Fig. 5-7: This circuit is usually attributed to Bob Widlar, but in his paper and patent he
considered only 1:1 emitter ratios. Widlar, "Some Circuit Design Techniques for Linear
Integrated Circuits," IEEE Transactions on Circuit Theory, Dec. 1965, pp. 586-590.
Widlar, "Low-value current source for integrated circuits," US Patent 3,320,439, 1967
Fig. 5-14: George Erdi, "Starting to Like Electronics in Your Twenties", p 172, in
Williams, "Analog Circuit Design", Butterworth-Heinemann, Stoneham, MA, 1991. Erdi,
US Patent 4,837,496, 1989
Chapter 6
dB: Martin, W.H., "DeciBel - The New Name for the Transmission Unit, Bell System
Technical Journal, January 1929
Steinmetz: Wagoner, C.D., "Steinmetz Revisited", IEEE Spectrum, April 1965, pp.82-95.
Kline, Ronald R., "Steinmetz", Johns Hopkins University Press, 1992
Chapter 7
Hilbiber, D.F., "A New Semiconductor Voltage Standard", International Solid State
Circuits Conference, 1964 (ISSCC 1993 Commemorative Supplement, pp. 34-35)
Brokaw, A. P., "A Simple Three-Terminal IC Bandgap Reference", IEEE Journal of Solid-
State Circuits, December 1974, pp. 388-393. Brokaw, "Solid-State Regulated Voltage
Supply", US Patent 3,887,863, June 3, 1975
Gunawan, M., Meijer, G., Fonderie, J. and Huijsing, J., "A Curvature-Corrected Low-
Voltage Bandgap Reference", IEEE Journal of Solid State Circuits, June 1993, pp. 667-670
Chapter 8
Solomon, James E.: "The Monolithic Op Amp: A Tutorial Study", IEEE Journal of Solid
State Circuits, December 1974, pp. 314-332
De Langen, Klaas-Jan and Huijsing, Johan H.: "Compact Low-Voltage Power Efficient
Operational Amplifier Cells for VLSI, IEEE Journal of Solid State Circuits, October 1998,
pp. 1482-1496
Chapter 10
Figure 10-2 is derived from the National Semiconductor LM13700. A similar concept is
used in the RCA (now Intersil) CA3280
Chapter 11
Low-Voltage 555: Camenzind, Hans R.: "Redesigning the old 555", IEEE Spectrum,
September 1997, pp. 80-85
Matthys, Robert J.: "Crystal Oscillators", revised edition, Krieger Publishing Company,
1992
Chapter 12
Gardner, Floyd M.: "Phaselock Techniques", John Wiley and Sons, second edition, 1979
Chapter 13
Sallen, R.P. and Key, E.L.: "A Practical Method of Designing RC Active Filters", IRE
Transactions on Circuit Theory, March 1955, pp. 74-85
Butterworth, S.: "On the Theory of Filter Amplifiers", Wireless Engineer, 1930, pp. 536-
541
Brodersen, R.W., Gray, P.R. and Hodges, D.A.: "MOS Switched Capacitor Filters",
Proceedings of the IEEE, January 1979, pp. 61-75
Camenzind, H.R.: "Modulated Pulse Power Amplifiers for Integrated Circuits", IEEE
Transactions on Audio and Electroacoustics, September 1966, pp. 136-140
Attwood, Brian E.: "Design Parameters Important for the Optimization of Very-High
Fidelity PWM (Class D) Audio Amplifiers", Journal of the Audio Engineering Society,
November 1983, pp. 842-853
Chapter 15
Boser, Bernhard E. and Wooley, Bruce A.: "The Design of Sigma-Delta Modulation
Analog-to-Digital Converters", IEEE Journal of Solid-State Circuits, December 1988, pp.
1298-1308
Candy, James C. and Temes, Gabor C.: "Oversampling Delta-Sigma Data Converters",
IEEE Press, 1992
Chapter 16
Gilbert, Barrie: "A New Wide-Band Amplifier Technique", IEEE Journal of Solid-State
Circuits, December 1968, pp. 353-365
Index
Thermometer 16-10
Threshold voltage 1-26
Timers 11-4, 11-14
Transconductance 1-26
Transconductance Amplifier 10-1