0 ratings0% found this document useful (0 votes) 741 views362 pagesIntroduction To CMOS OP-AMPs and Comparators-1
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content,
claim it here.
Available Formats
Download as PDF or read online on Scribd
INTRODUCTION TO
CMOS OIE
_AND COMPARATORS
ior. GREGORIANCONTENTS
PREFACE
1 INTRODUCTION 1
1d
12
Classification of Signal Processing Techniques / 1
Examples of Applications of Op-Amps and Comparators
in Analog MOS Circuits / 6
Problems / 16
References / 16
2 MOS DEVICES AS CIRCUIT ELEMENTS 17
21
22
23
24
25
26
27
28
Semiconductors / 17
MOS Transistors / 21
MOS Transistor Types: Body Effect / 27
‘Small-Signal Operation and Equivalent Circuit of MOSFET
Transistors / 30
Weak Inversion / 39
Impact Ionization / 40
Noise in MOSFETS / 41
CMOS Process / 44
Problems / 45
References / 47
viiviii. = CONTENTS
3 BASIC ANALOG CMOS SUBCIRCUITS 48
3.1
32
3.3
34
35
3.6
Bias Circuits in MOS Technology / 48
MOS Current Mirrors and Current Sources / 55
MOS Gain Stages / 63
MOS Source Followers / 74
MOS Differential Amplifiers / 77
Frequency Response of MOS Amplifier Stages / 84
Problems / 92
References / 94
4 CMOS OPERATIONAL AMPLIFIERS, 95
4.1
4.2
4.3
44
45
46
47
48
49
4.10
Operational Amplifiers / 95
Single-Stage Operational Amplifiers / 100
Two-Stage Operational Amplifiers / 106
Stability and Compensation of CMOS Amplifiers / 112
Dynamic Range of CMOS Op-Amps / 126
Frequency Response, Transient Response. and Slew Rate of
‘Compensated CMOS Op-Amps / 132
Noise Performance of CMOS Op-Amps / 137
Fully Differential Op-Amps / 140
CMOS Output Stages / 149
Op-Amps with Rail-to-Rail Input Common-Mode Range / 164
Problems / 170
References / 173
5 COMPARATORS 175
5.1
5.2
5.3
54
5.5
56
6.1
6.2
Circuit Modeling of a Comparator / 175
Single-Ended Auto-Zeroing Comparators / 177
Differential Comparators / 182
Regenerative Comparators (Schmitt Triggers) / 192
Fully Differential Comparators / 198
Latches / 205
Problems / 212
References / 213
DIGITAL-TO-ANALOG CONVERTERS. 214
Digital-to-Analog Conversion: Basic Principles / 214
Voltage-Mode D/A Converter Stages / 21863
6.4
65
6.6
CONTENTS = ix
Charge-Mode D/A Converter Stages / 231
Hybrid D/A Converter Stages / 234
Current-Mode D/A Converter Stages / 238
Segmented Current-Mode D/A Converter Stages / 244
Problems / 252
References / 254
ANALOG-TO-DIGITAL CONVERTERS 255
TA
72
73
7
1S
76
TT
Analog-to-Digital Conversion: Basic Principles / 255
Flash A/D Converters / 263
Interpolating Flash A/D Converters / 270
Two-Step A/D Converters / 273
Successive-Approximation A/D Converters / 282
Counting and Tracking A/D Converters / 294
Integrating A/D Converters / 295
Problems / 300
References / 301
PRACTICAL CONSIDERATIONS AND DESIGN EXAMPLES 303
81
8.2
8.3
Practical Considcrations in CMOS Op-Amp Design / 303
Op-Amp Design Techniques and Examples / 316
Comparator Design Techniques and Examples / 349
Problems / 355
References / 355
INDEX 357CHAPTER 1
INTRODUCTION
Operational amplifiers (op-amps) and comparators are two of the. most important
building blocks for analog signal processing. Op-amps and a few passivé components
can be used to realize such important functions as summing and inverting amplifiers,
integrators, and buffers. The combination of these functions and comparators can
‘result in many complex functions, such as high-order filters, signal amplifiers, ana-
Jog-to-digital (A/D) and digital-to-analog (D/A) converters, input and output signal
buffers, and many more. Making the op-amp and comparator faster has always been
one of the goals of analog designers. In this chapter the basic concept of digital and
analog signal processing is introduced. Then a third category of signal processing,
the sampled-data analog technique, which is in between the two main classifications,
is described. Finally, a few representative examples are given of circuits and systems
utilizing CMOS op-amps and comparators, to illustrate the great potential of these
components as part of an MOS-LSI chip.
1.1 CLASSIFICATION OF SIGNAL PROCESSING TECHNIQUES [1-4]
Electrical signal processors are usually divided into two categories: analog and digital
systems. An analog system carries signals in the form of voltages, currents, charges,
and so on, which are continuous functions of the continuous-time variable. Some
typical examples of analog signal processors are audio amplifiers, passive- or active-
RC filters, and so on. By contrast, in a digital system each signal is represented by
a sequence of numbers. Since these numbers can contain only a finite number of
digits (typically, coded in the form of binary digits, or bits) they can only take on
discrete values. Also, these numbers are the sampled values of the signal, taken at
discrete time instances. Thus both the dependent and independent variables of a
12 INTRODUCTION
signal are discrete. Since the processing of the digital bits is usually performed
synchronously, a timing or clock circuit is an important part of the digital system.
The timing provides one or more clock signals, each containing accurately timed
pulses that operate or synchronize the operation of the components of the system.
Typical examples of digital systems are a general-purpose digital computer or a
special-purpose digital signal processor dedicated to (say) calculating the Fourier
transform of a signal via the fast Fourier transform (FFT), or a digital filter used in
speech analysis, and so on.
By contrast, analog signal processing circuits utilize op-amps, comparators, resis-
tors, capacitors, and switches to perform such functions as filters, amplifiers, recti-
fiers, and many more. To understand the basic concepts of the most commonly used
configurations of an analog circuit, consider the simple analog transfer function
Vout (5) b
Vi) 24 as +b" ap
It is easy to verify that the RLC circuit shown in Fig. 1.1@ can realize this function
(Problem 1.1). Although this circuit is easy to design, build, and test, the presence
of the inductor in the circuit makes its fabrication in imegrated form impractical.
In fact, for low-frequency applications, this circuit may well require a very large
valued, and hence bulky, inductor and capacitor. To overcome this problem, the
designer may decide to realize the desired transfer function using an active-RC
circuit. It can readily be shown that the circuit in Fig. 1.1b, which utilizes three
operational amplifiers, is capable of providing the transfer function specified in
oo)
Figure 1.1, Second-order filter realization: (a) passive circuit; (b) active-RC circuit.1.1 CLASSIFICATION OF SIGNAL PROCESSING TECHNIQUES [1-4] 3
qa 7h e (84g
Figure 1.2. Switched-capacitor realization of a resistive branch.
Eg. (1.1). This circuit needs no inductors and may be realized with small discrete
components for a wide variety of specifications (Problem 1.2). It turns out, however,
that while integration of this circuit on a bipolar chip is, in principle, feasible (since
the amplifiers, resistors, and capacitors needed can all be integrated), there are some
major practical obstacles to integration. These include the very large chip area needed
by the RC components, as well as the stringent accuracy and stability requirements
for these elements. These requirements cannot readily be satisfied by integrated
components, since neither the fabricated values nor the temperature-induced varia-
tions of the resistive and capacitive elements track each other. The resulting
pole-zero variations are too large for most applications.
Prior to mid-1970s, analog circuits such as the one shown in Fig. 1.1 were imple-
mented using integrated bipolar op-amps and discrete passive components. In the
1970s two developments made it possible to fully integrate analog circuits in metal-
oxide semiconductor (MOS) technology. The first development was the emergence
of a technique called switched-capacitor (SC) circuits [6], which is an effective
strategy for solving both the area and the matching problems by replacing each
resistor in the circuit by the combination of a capacitor and a few switches. Consider
the branches shown in Fig. 1.2. Here, the four switches S,, Sz, 53, and S, open and
close periodically, at a rate which is much faster than that of the variations of the
terminal voltage vy and vg. Switches 5; and 5, operate synchronously with each
other but in opposite phase with 5, and 53. Thus when 5S; and S3 are closed, 5, and
Sq are open, and vice versa. Now when S, and S; close, C is discharged. When S>
and S3 open, $, and S, close, and C is recharged to the voltage tie = v4 — vp. This
causes a charge y = C(u, — vg) to flow through the branch of Fig. 1.2. Next, C
is again discharged by S2 and S3, and so on. If this cycle is repeated every T seconds
(where T is the switching period or clock period), the average current through the
branch is then
le = 4 = Fon = wy). (2)
Thus ig, is proportional to the branch voltage v4 — vp. Similarly, for a branch
containing a resistor R, the branch current is i = (//R)(va — vg). Thus the average
current flowing in these two branches are the same if the relation R = T/C holds.4 INTRODUCTION
C4
Sa op
‘e9 Sols Ses | ty
: |
YY yy
Figure 1.3. Second-order switched-capacitor filter section.
Physically, the switches transform the capacitor C, a nondissipative memoried ele-
ment, into a dissipative memoryless (i.c., resistive) onc.
It is plausible therefore that the branch of Fig. 1.2 can be used to replace all
resistors in the circuit of Fig. 1.16. The resulting stage [3] is shown in Fig. 1.3. In
this circuit, switches that belong to different “resistors” but perform identical tasks
have been combined. Furthermore, the second operational amplifier (op-amp) in
Fig. 1.1b, which acted merely as a phase inverter, has been eliminated. This was
possible since by simply changing the phasing of two of the switches associated
with capacitor C3, the required phase inversion could be accomplished without an
op-amp.
As Fig. 1.3 illustrates, the transformed circuit contains only capacitors, switches,
and op-amps. A major advantage of this new arrangement is that now all time
constants, previously determined by the poorly controlled RC products, will be given
by expressions of the form (T/C;)C2 = T(C2/C,). Here the clock period T is usually
determined by a quartz-crystal-controlled clock circuit and hence is very accurate
and stable. The other factor of the time constant is C2/C,, that is, the ratio of two
on-chip MOS capacitances. Using some simple rules in the layout of these elements,
it is possible to obtain an accuracy and stability on the order of 0.1% for this ratio.
‘The resulting overall accuracy is at least 100 times better than what can be achieved
with an on-chip resistor and capacitor for the RC time constant.
A dramatic improvement is also achievable for the area required by the passive
elements. To achieve a time constant in the audio-frequency range (say 10 krad/s),
even with a large (10-pF) capacitor, a resistance of 10-MQ is required. Such a
resistor will occupy an area of about 10° j:m?, which is prohibitively large; it is
nearly 10% of the area of an average chip. By contrast, for a typical clock period
of 10 ys, the capacitance of the switched capacitor realizing a 10-MQ resistor is
C= TR = 10-710" = 10-'?F = 1 pF. The area required realizing this capacitance
is about 2500 um?, only 0.25% of that needed by the resistor that it replaces.1.1 CLASSIFICATION OF SIGNAL PROCESSING TECHNIQUES 1-4) 5
‘The second development that made the realization of the fully integrated analog
MOS circuits possible was the design of the MOS op-amp. Perhaps the most gener-
ally useful analog circuit function is that of the operational amplifier. Prior to about
1977, there existed a clear separation of the bipolar and MOS technologies, according
to the function required [1,5]. MOS technology, with its superior device density,
was used mostly for digital logic and memory applications, while all required analog
functions (such as amplification, filtering, and data conversion) were performed
using bipolar integrated circuits, such as bipolar op-amps. Since that time, however,
rapid progress made in MOS fabrication techniques made it possible to manufacture
much more complex and flexible chips. In addition, new developments occurred in
communication technology (such as digital telephony, data transmission via tele-
phone lines, adaptive communication channels, etc.) which required analog and
digital signal processing circuitry in the same functional blocks. The analog functions
most often needed are filtering (for antialiasing, smoothing, band separation, etc.),
amplification, sample-and-hold operations, voltage comparison, and the generation
as well as precise scaling of voltages and currents for data conversion. The separation
of these analog functions from the digital ones merely because of the different
fabrication technologies used is undesirable, since it increases both the packaging
costs and the space requirements and also, due to the additi interconnections
required, degrades the performance. Hence there was strong motivation to develop
novel MOS circuits, which can perform these analog functions and which can also
share the area on the same chip with the digital circuitry.
Compared with bipolar technology, MOS technology has both advantages and
disadvantages. MOS device has extremely high impedance at its input (gate) termi-
nal, which cnables it to sense the voltage across a capacitor without discharging it.
Also, there is no inherent offset voltage across the MOS device when it is used as
a charge switch. Furthermore, high-quality capacitors can be fabricated reliably on
an MOS chip. These features make the realization of such circuits as precision
sample-and-hold stages feasible on an MOS chip [1]. This is usually not possible
in bipolar technology.
On the negative side, the transconductance of MOS transistors is inherently lower
than that of bipolar transistors. A typical transconductance value for a moderate-
sized MOS device is around 2.5 mA/V; for a bipolar transistor, it may be about 50
times larger. This leads to a higher offset voltage for an MOS amplifier than for a
bipolar amplifier. (At the same time, however, the input capacitance of the MOS
transistor is typically much smaller than that of a bipolar transistor.) Also, the noise
generated in an MOS device is much higher, especially at low frequencies, than in
a bipolar transistor. The conclusion is that the behavior of an amplifier realized on
an MOS chip tends to be inferior to an equivalent bipolar realization in terms of
offset voltage, noise, and dynamic range. However, it can have much higher input
impedance than that of its bipolar counterpart.
‘As a result of these properties, the largest use of the MOS op-amp is expected
to be as part of an MOS-LSI (large-scale integration) chip. Here the design of the
op-amp can take advantage of the important performance specifications that are
needed. The loading of the op-amp is often very light and usually only a small-6 INTRODUCTION
valued capacitor has to be driven by these op-amps. Switched-capacitor circuits fall
especially into this category, where element-value accuracy is important but the
signal frequency is not too high and the dynamic range required is not excessive.
Voice- and audio-frequency filtering and data conversion are in this category and
represent the bulk of the past applications.
In addition to frequency-selective switched-capacitor filtering introduced in Fig.
1.3, which has been the most common application of MOS op-amps, there are many
other functions for which op-amps and comparators can be used. These include
analog-to-digital (A/D) and digital-to-analog (D/A) data conversion, programmable-
gain amplification for AGC and other applications, peak-detection, rectification,
zero-crossing detection, and so on. They have also been used extensively in large
mixed-signal analog/digital systems such as voice codecs, high-speed data communi-
cation modems, audio codecs, and speech processors. This range will expand contin-
uously as the quality (bandwidth, dynamic range, power consumption, etc.) of the
components, especially op-amps and comparators, improves.
1.2, EXAMPLES OF APPLICATIONS OF OP-AMPS AND
COMPARATORS IN ANALOG MOS CIRCUITS
In this section, a few selected examples of practical analog MOS circuits are given
where CMOS op-amps and comparators are used extensively. Of course, the reader
should not expect to understand the details of these systems at this stage. However,
the diagrams may give an idea of the potentials of these components in analog signal
processing.
As mentioned earlier, one of the most important applications of CMOS op-amps
is in switched-capacitor filters. Figure 1.4a shows the circuit diagram of a seventh-
order switched-capacitor filter. Its measured frequency response is shown in Fig.
1.4b, The measured passband variation for the device is less than 0.06 dB. This
represents a superior performance, which could not have been achieved without
‘extensive trimming using any other filter technology.
An obvious application of a CMOS op-amp is the realization of charge-mode
digital-to-analog converters (DAC). It can be obtained by combining a programmable
capacitor array and an offset-free switched-capacitor gain stage. An example of an
N-bit charge-mode DAC is shown in Fig. 1.5, where V,,; is a temperature-stabilized
constant reference voltage. The output of the DAC is the product of the reference
voltage and the binary-coded digital signal (by, bz, bs,..., by). In Chapter 6 the
design of such circuits is discussed in some detail.
Modulators, rectifiers, and peak detectors [6] belong to an important class of
nonlinear circuits, which can be implemented with a combination of op-amps and
comparators. In an amplitude modulator the amplitude of a signal x(t) (usually called
the carrier) is varied (modulated) by m(,), the modulating signal. Hence the output
signal y(t) is the product of x(r) and m(t), or y(t) = x(t\m(t). A periodic cartier
signal, which is readily generated from a stable clock source, is a square wave
alternating between two equal values + V. An easy way to perform modulation with-raujy ssed-mo] sonsedeo-poyortms
Jopwo-tpvanas v Jo asuodses Aouanbaxy pamnsvau (q) ‘wex8eIp wna (P) “pT aanBiy
ty8 INTRODUCTION
o 4000 8000 12000 18000
frequency (Hz)
o
Figure 1.4. Continued
top plate
bottom plate
Figure 1.5. Multiplying digital-to-analog converter.1.2. EXAMPLES OF APPLICATIONS OF OP-AMPS & COMPARATORS INANALOG MOSGIRGUITS = 9
Figure 1.6. Switched-capacitor modulator with two clock signals.
a square-wave carrier is to switch the polarity of the input signal m(1) periodically.
A stray-insensitive switched-capacitor modulator circuit which performs according
to this principle is shown in Fig. 1.6. The clock phases , and ¢ are operated at
the fast clock rate w,, while the phase ¢, changes at the slow carrier-frequency rate
4. Normally, w, is much larger (by a factor of 30 or more) than w,4.
Another nonlinear circuit is a full-wave rectifier that converts an input signal viq(t)
to its absolute value |via(t)]. A simple way of implementing a switched-capacitor full-
wave rectifier is to add a comparator to an amplitude modulator. The circuit of a
switched-capacitor full-wave rectifier based on the modulator of Fig. 1.6 is shown
in Fig. 1.7a. Here A is set to “1” if uj, > 0 and to “0” if 4, < 0, while B is set
to A by the comparator and the latch that follows it each time ,, goes high. The
signals A and B then set the polarity of the transfer function so that it inverts the
negative input signals, but not positive ones. Figure 1.7b shows an auto-zeroing
comparator, which is discussed in detail in Chapter 5.
A peak detector is a circuit whose output holds the largest positive (or, if so
specified, negative) voltage earlier attained by the input signal. An MOS peak detec-
tor is shown in Fig. 1.8. The op-amp acts as a comparator, with vou = Vmax and
Up as its inputs. If y > Vmax, the op-amp output goes high and M, conducts,
charging C until vow ~ vin is reached. If ti, < Vinax, the op-amp output is low, M;
is cut off, and vx = Vmax is held by C.
One of the most important applications of the comparators is in A/D converters.
A successive-approximation A/D converter is one type of medium-speed Nyquist-
rate converter that can be realized using a programmable capacitor array (PCA) and
a voltage comparator. A 5-bit converter is shown in Fig. 1.9. For high-speed opera-
tion, flash A/D converters can be used. In this configuration an array of 2” compara-
tors are used for an N-bit A/D converter. A conceptual diagram of an N-bit flash10 INTRODUCTION
(a)
%
ry -
+ og
‘n Yout
%24|
)
Figure 1.7. Switched-capacitor full-wave rectifier: (a) complete circuit; (b) offset-compen-
sated comparator.
AID convener is shown in Fig. 1.10. Analog-to-digital converters are discussed in
detail in Chapter 7.
With the recent rapid progress made in MOS fabrication techniques and the
emergence of the submicron CMOS technology, many intricate systems containing
analog and digital functions have been combined in a fully integrated form. One
drawback of the submicron CMOS technology is the reduction in the power supply
voltage, which results in a reduced signal swing and hence a lower dynamic range
To improve the performance of the system and reduce the effects of noise injection
from the power, ground, and clock lines, most modern high-performance mixed-1.2, EXAMPLES OF APPLICATIONS OF OP-AMPS & COMPARATORS IN ANALOG MOS CIRCUITS 11
Yop
m1
Yout
I
Figure 1.8, Continuous-time peak detector.
Q,- -2¢v,,
e to the digital logic
@
Figure 1.9. Five-bit successive-approximation A/D converter.12 INTRODUCTION
Figure 1.10. Conceptual diagram of an N-bit A/D converter.
signal integrated circuits make use of fully differential signal paths. With op-amps
and comparators, the fully differential signal paths require fully differential outputs
as well as inputs. and they are known as fully differential op-amps and comparators.
Since this technique uses symmetrical layout, many of the noise voltages (power
supply noise, clock-feedthrough noise, offset voltages) appear as common-mode
signals. They are to a considerable extent canceled in the differential output voltage
Zoe at all frequencies. A high-frequency high-Q switched-capacitor bandpass filter
that uses a fully differential signal path is shown in Fig. 1.11. This filter is typically
used in a radio-frequency (RF) receiver system, which requires high selectivity at
high frequencies [7}. The two complementary switch blocks (X; and X2) are shown
in Fig. 1.12. The filter uses fully differential single-pole transconductance folded-
cascode op-amps with source-follower common-mode feedback as illustrated in Fig.
1.13 [8]. This op-amp achieves 100-MHz unity-gain bandwidth and 60 dB of gain
with 1 mA of total current consumption. Fully differential op-amps are discussed
in detail in Chapter 4.“ray ssedpueq ajod-|je 2oy2edes-payowas xopso-WIXIS & Jo wesBeIP OMEWOHDS “TTT aN14 INTRODUCTION
X
fs
+
=
Figure 1.12. Two switch blocks for a double sampling.
‘Another application of the fully differential op-amps is in oversampling, or delta-
sigma A/D converters. The oversampling converters operate at sampling rates of 16
to 512 times the Nyquist rate and increase the signal-to-noise ratio by subsequent
filtering. The oversampling techniques lend themselves most favorably to applica-
tions that require a relatively low frequency (<1 MHz) and high resolution (>12
bits). The most obvious application of delta-sigma converters is in digital telephony
ss
Figure 1.13. Wideband op-amp for the filter.1.2, EXAMPLES OF APPLICATIONS OF OP-AMPS & COMPARATORS IN ANALOG MOS CIRCUITS 15.
Veet] [Vref Veet" [rot
_ i j $4 ' s,| °F en Li
x | ¥
“so i}
fet 2 fem 2
war! toe wat by
@
(o)
Figure 1.14. (a) Fully differential CMOS implementation of a second-order delta-sigma
modulator; (b) two-phase clock scheme.
and digital audio. Figure 1.14 shows a fully differential, switched-capacitor CMOS
implementation of a second-order delta-sigma modulator [9]. It consists of two para-
sitic insensitive integrators, a comparator that serves as a 1 bit A/D converter, and
a two-level (1-bit) D/A converter. Use of a fully differential configuration attenuates
power supply noise, clock feedthrough, and even-order harmonic distortion. The
modulator operates on (wo-phase nonoverlapping clocks consisting of a sampling
phase and an integration phase. It achieves 16-bit dynamic range with an oversam-
pling ratio of 256 and a signal bandwidth of 20 kHz.16 INTRODUCTION
As the examples above illustrate, present-day CMOS op-amps and comparators
and their use in analog MOS circuits have reached a certain level of maturity. Al-
ready, almost any analog signal processing task in the voice- or audio-frequency
range has a possible solution using such circuits. As fabrication technology and
circuits design techniques continue to advance, the speed and dynamic range of
these circuits will increase, allowing their use in such large-volume applications as
video and radio systems, image processing, high-speed transmission circuits, and
so on.
PROBLEMS
1.1. Show that the circuit of Fig. 1.1a can realize the transfer function of Eq. (1.1).
What should be the clement values R, L, and C?
1.2. Calculate the transfer function of the active-RC circuit of Fig. 1.1b. Assume
that the circuit is to realize the transfer function of Eq, (1.1). Write the available
equations for the element values. How many element values can be chosen
arbitrarily?
REFERENCES
1. R. W. Brodersen, P. R. Gray, and D. A. Hodges. Proc. IEEE. 67. 61-75 (1979)
2. Y. Tsividis, Proc. IEEE, 71, 926-940 (1983),
3. R. Gregorian, K. W. Martin, and G. C. Temes, Proc. IEEE. ~! 941-966 1983.
4. D. J. Allstot and W. C. Black, Jr., Proc. IEEE, 71, 967-986 11983
5. P.R. Gray and R. G. Meyer, Analysis and Design of Analog Ince eraed Circutts. 2ad ed.
Wiley, New York, 1984.
6. R. Gregorian and G. C. Temes, Analog MOS Integrated Circutts for Signal Processing,
Wiley, New York, 1986.
. Bang-Sup Song and P. R. Gray, JEEE, J. Solid-State Circuits. SC-21(6), 924-933 (1986).
8. T.C. Choi, R. T. Kaneshira, R. W. Broderson, P. R. Gray. W. B, Jett, and M. Wilcox,
IEEE J. Solid-State Cirewits, SC-18(6), 652-664 (1983),
9. B. P. Brandi. D. E. Wingard, and B. A. Wooley. IEEE J. Solid-State Circuits, SC-26(4),
618-627 (1991)
_CHAPTER 2
a
MOS DEVICES AS CIRCUIT
ELEMENTS
In this chapter the physics of MOS (metal-oxide semiconductor) devices is discussed
briefly. The most important and simplest current-voltage relations are given, and
simple models introduced for MOS transistors in linear operation. The discussion
here is in the simplest possible terms, aimed at providing some physical understand-
ing of the highly complex device operation for the circuit designer. Precision and
depth have regretfully been sacrificed in the process. The ambitious reader is referred
to the excellent specialized works listed as references at the end of the chapter.
2.1. SEMICONDUCTORS
In metals (e.g., aluminum, copper, silver) that are good electrical conductors, the
atoms are arranged in a regular crystal array. The electrons from the outer (valence)
shell of the atoms are free to move within the material. Since the number of atoms,
and thus the number of free electrons, is very large (on the order of 107° cm™*), even
a small electric field results in a large electron current—hence the high conductivity
observed for these metals.
The picture is quite different for an insulator such as silicon dioxide (SiOz).
Here the valence electrons form the bonds between adjacent atoms and hence are
themselves tied to these atoms. Thus no frec clectrons are available for conduction
and the conductivity is very low.
Semiconductors (such as silicon or germanium) are in between conductors and
insulators in their electrical properties. At very low temperatures, the valence elec-
trons are bound to their atoms, which again form a regular lattice. However, as the
temperature is raised, due to the thermal vibrations of the atoms, some bonds will
be broken, and an electron escapes from each of these bonds. Such electrons are
718 MOS DEVICES AS CIRCUIT ELEMENTS
capable of conducting electricity. Furthermore, each fugitive electron leaves a charge
deficit (called a hole) behind in the bond. A valence electron in a bond close to a
hole can easily move over, filling the hole and leaving a new hole in its own bond.
The effect is the same as if the hole had moved from one bond to the next. Since
the hole ‘‘moves’’ in a direction opposite that of the moving valence electron, in
an electronic field it behaves like a positively charged particle.
Electrical conduction is thus possible for a semiconductor at room temperature.
The density of thermally gencrated electrons and holes is, however, much smaller
than that of the free electrons in metal. Typical numbers are 10'° charge carriers
per cubic centimeter for silicon and 10" in germanium. In what follows, the currently
dominant material, silicon, is discussed exclusively.
Adding foreign elements (dopants) to the pure silicon can raise the number of
free charge carriers in a semiconductor. Silicon (and germanium) has four valence
electrons. If an atom of an element with five valence electrons (such as arsenic,
phosphorus, or antimony) is added to the semiconductor, it may take the place of
a silicon atom in the crystal lattice. Thus four of its valence electrons will participate
in the four bonds tying the atom to adjacent semiconductor atoms in the lattice. The
fifth valence electron of the foreign atom, however, will not have a place in any
bond and will thus be free to move away from its parent atom. Hence such a dopant
element (called a donor, since it contributes free electrons to the semiconductor)
enhances the conductivity of the material.
Adding atoms of an clement with three valence electrons will also contribute to
the conductivity. Now there will be a bond lacking a valence electron for each dopant
atom. Thus each such atom creates one hole. These dopants (e.g., boron, aluminum,
and gallium) are called acceptors, since the holes will propagate by accepting bound
valence electrons from adjacent semiconductor atoms.
In doped semiconductors there will be carriers due to thermal effects as well as
to the donor (or acceptor) atoms. Materials containing donors will thus have both
free electrons and holes, but there will be more electrons than holes, Such semicon-
ductors will be called n-type, where n stands for negative. Materials containing
acceptors will have a majority of holes; they are called p-type semiconductors, where
p stands for positive.
A semiconductor structure can also be fabricated that contains two adjacent re-
gions of different types (Fig. 2.1). The surface joining the two regions is called a
pn junction. When the junction is fabricated, the random thermal motion of the
solder joint
a
type
wire
Z
Figure 2.1. A pn junction diode. pn junction2.1. SEMICONDUCTORS 19
++++ 4+ +
3
“|
Figure 2.2. [on layers in a pn junction.
majority carriers (elecurons in the n-type region, holes in the p-type region) will
cause electrons to spill over from the n-type region to the p-type region. Vice versa,
holes will move from the p-type region to the n-type semiconductor. Thus this
random motion (called diffusion) results in the p-type semiconductor being charged
negatively while the n-type region is charged positively. The effect will be strongest
near the junction: There, in the p-type region, the negatively charged acceptor atoms
will no longer be neutralized by holes, and (in the ri-type region) free electrons will
no longer surround the positively charged donor ions. Hence in this area a dipole
layer of fixed ions will be formed (Fig. 2.2). The electric field € created by the
dipole opposes further majority-carrier diffusion, but it helps the thermally generated
minority carriers (electrons in p-type regions, holes in n-type regions) to migrate
from one region to another. Thus, after a short transient, equilibrium will be obtained.
Four different carrier currents will flow: Majority carriers will move hy diffusion
from region to region despite €, and minority carriers will flow aided by €. These
currents cancel each other in equilibrium, since the effects of € compensate for the
larger number of available majority carriers.
The equilibrium will be upset if a voltage source is connected to the wires soldered
to the semiconductor (Fig. 2.3). Assume first that the polarity of the source is such
that it makes the p-region more positive with respect to the n-region; that is, v > 0
in Fig. 2.3. Then v will reduce € and thus increase the current of majority carriers
held back by € from spilling over the boundary. Even a small reduction of € caused
by, say, a battery of v = 0.8 V can result in a large majority-carrier current (say,
depletion
region
Figure 2.3. Circuit for testing a pn diode,20 MOS DEVICES AS CIRCUIT ELEMENTS
Nama)
Figure 24, Current versus voltage 0.5 “
characteristics of a pn junction diode. 's
i = 1 A) in the circuit. Hence v with the polarity indicated will be called forward
voltage and i forward current.
Let us now reverse the polarity of the voltage source so that v < 0 in Fig, 2.3.
Now v will aid & in obstructing the flow of majority carriers from region to region.
If vis large enough, the majority current is essentially eliminated, and only the flow
of minority carriers (electrons moving from the p region to the n region, and holes
moving in the opposite direction) remains. Since the number of minority carriers
is, however, small and nearly independent of v, the resulting net current will be
small and nearly constant. This is the case of reverse voltage and current. With the
reference directions used in Fig. 2.3, now i < 0. Figure 2.4 illustrates the overall
behavior of i as a function of v. A detailed theoretical analysis [1, Sec. 4.3; 2, Sec.
6.6] reveals that the describing equation is, to a good approximation,
i = Ife — 1), (2.1)
Here J is the saturation current, determined by the geometry and the material
properties of the device, g ~ 1.6 X 10-9 C is the electron charge, and k = 1.38
X 107” J/K is Boltzmann’s constant. T is the temperature of the semiconductor,
in Kelvin. At room temperature (T = 300 K), kT/q ~ 26 mV. Is is usually very
small, on the order of 10~ A or less. Thus i increases exponentially with v for v
> 0, while i= —Jg and is very small if v < 0 (Fig. 2.4).
The behavior of the region directly adjacent to the boundary between the p and
n regions is of prime importance. As mentioned earlier, the majority carriers are
very sparse in this area; some have immigrated into the other region, and the others
have been pushed back into the inside of their native region by the field €, Hence
the border area contains only the fixed ions, charged negatively in the p-type region
and positively in the n-type material (Fig. 2.3). This area is hence called the depletion
region. Its width increases with increasing €; hence it will be greater (smaller) for
reverse (forward) voltage v.
Due to the field €, a voltage 4; (often called the built-in voltage) appears across
the depletion region for v = 0. The total potential across the junction, for v¥ 0, is
thus ¢;'— v. Typically, @; = 0.5 to 1 V.
For v < 0, the pn junction can be regarded as a capacitor, since only a small22, MOS TRANSISTORS = 27
saturation current /s flows for a de voltage v, and since adjacent positive (+ Q) and
negative (— Q) charges are stored in the depletion region (Fig. 2.3). Since the charge
stored is a nonlinear function of v, the capacitance is nonlinear. We shall define the
capacitance C by the incremental relation C = dQ/du It can then be shown [1, Sec.
3.3; 2, Sec. 6.5] that for the device illustrated in Fig. 2.3,
rr
c- (eta + uNath 22)
o + lf
holds. Here 1 KHz) signal rather than a constant voltage.
Consider next the structure shown in Fig. 2.7. A new feature is the presence of
two n* (ie. heavily doped n-type) regions in the p-type material. The one on the
left will be called the source; a voltage us is connected to it. The n* region on the
* Often, the oxide thickness / is measured in angstroms (1 A = 10~® em), Usual values of / are between
50 and 200 A.22. MOS TRANSISTORS 23.
Figure 2.7, MOS transistor.
right will be called the drain; its voltage is denoted by vp. The top metal electrode
will be called the gate; its voltage is vg. The body of the semiconductor is usually
called the substrate or bulk. The overall device is the MOS transistor. Its operation
is discussed briefly next.
Let the source be grounded, so that vs = 0. Also, let up have a sutall positive
value, say 0.5 V. We will consider the behavior of the drain current ip as vg is
raised from zero to higher positive values. Since the gate is insulated from the rest
of the device by the oxide layer, it will not conduct any current. The n° drain
region and the surrounding p-type substrate form a pn junction. Since the substrate
is grounded, while vp > 0, this junction is reverse biased. Hence for ug = 0,
in ~ 0.
As ug is increased, the region R under the gate will first be depleted, then inverted,
as discussed earlier in connection with Figs. 2.5 and 2.6. When R is depleted, ip
remains Zero, since the area around the drain is still reverse biased. However, the
situation changes when ug is so large that inversion occurs, so that R is filled with
electrons. Now, a layer containing mobile electrons, called an inversion layer or
channel, connects the drain to the source. Since the drain is positive with respect
to the source, electrons will flow from the source to the drain and a positive current
ip > 0 will be observed. The smallest voltage vg necessary to produce a channel
is called the threshold voltage and is denoted by Vr. Usually, Vz is given as the ve
value needed for ip = 1 wA; it may range from a fraction of a volt to several volts.
It should be noted that for the structure of Fig. 2.7, most of the electrons in the
channel do not originate from thermal effects in the bulk; instead, they are drawn
by the electric field due to vg out of the source. Some electrons are also drawn from
the drain; however, since vp > 0, the drain—substrate junction is more reverse biased,
and hence it is harder for electrons to escape from the drain.
Since a potential difference vp exists between the two ends of the channel, the
electrons in the channel will be attracted to the drain. Therefore, in addition to the
random thermal motion of the electrons, a steady motion (called drift) will occur,
which causes the current flow. For small vp, the channel will therefore behave as
a resistor, and hence ip ~ vp/R, where the channel resistance R is given by
L
R= Wald (2.4)24 MOS DEVICES AS CIRCUIT ELEMENTS.
Figure 2.8. Pinch-off in an MOS transistor.
Here L is the length and W the width of the channel, and y,, is the mobility of the
electrons in the channel,* defined by the relation (electron drift velocity) = (mobil-
ity) X (electric field). Finally. Q, is the charge density (in C/cm”) of the electrons
in the channel. Since vg can be considered as the sum of two terms,V; (necessary
to maintain the depletion region under the channel) and vg — Vz (necessary to
maintain the channel), we have
Q, = —Cox (We — Vr), (2.5)
where Cox = €ox/l is the capacitance (per unit arca) of the oxide layer separating
the gate from the channel. Hence, for small vp (i.¢., vp < ug — Vz), the relation
Ww
In = Wi€ox 7 (a — Vr 26)
holds. Thus the transistor acts as a resistor, with resistance R = [p,Cox W/L(we —
Vz)I7! controlled by ug.
If up is increased so that it is no longer negligible compared to vg, Eq. (2.6) will
become inaccurate. Since the potential of the channel at the grounded source is zero
while at the drain it is vp, we can assume that its average potential is vp/2. Hence
the average voltage between the gate and channel is (vg — vp/2). Replacing ug by
(vq — vp/2) in Eq. (2.6) gives
. w
ip = PsCon T (x = Vy = hon 27
Equation (2.7) remains a good approximation for up < vg — Vz. This range is
called the linear region (or triode region) of operation of the MOS transistor.
When vp = uc — Vr, anew phenomenon occurs. Consider the situation illustrated
in Fig. 2.8, where only the structure near the semiconductor surface is shown, magni-
* The mobility in the bulk of the semiconductor is higher. since 1 decreases with the concentration of
ionized impurities. Typical values are pty = 1000 cm?/V - s for Np = 10! cm7*, while tq = 100 cm*s
Vs for Np = 10" cm™*.22. MOS TRANSISTORS = 25
fied. As the figure indicates, due to the variation of the potential along the channel,
the charge density Q,, decreases near the drain. If vp = vg — Vz, at the drain the
gate-to-channel voltage is no longer sufficient to maintain the channel. Thus the
depletion region surrounding the source, the channel, and the drain extends all the
way to the surface. This phenomenon is sometimes called pinch-off, and the region
where it occurs is the pinch-off point (Fig. 2.8). If vp is increased further, the pinch-
off point will move toward the source, since the area where ug — up = Vy will
increase. Hence the channel will now extend only from the source to the pinch-off
point, the latter being somewhere under the gate. The region between the pinch-off
point and the drain is depleted. Electrons from the channel are injected into this
depletion region at the pinch-off point and are swept to the drain by the field created
by the potential difference between the drain and the pinch-off point. The voltage
ups & up — ts is thus divided between the two series-connected regions: the channel
between the source and the pinch-off point, and the depletion region between the
pinch-off point and the drain. Clearly, the latter has a higher resistance, and hence
most of ups in fact appears across it. Any increase of vp will, to a good approxima-
tion, result in an equal voltage increase across the depletion region and will hardly
change ip. Thus, for vp > ug — Vr, from Eq. (2.7),
in(ep) = insa 2 ipW@nsst) (2.8)
= be Fug — vn.
This phenomenon is called saturation; ups. = ug — Vr is the drain saturation
voltage and ipsa: as given by Eq. (2.8), is the drain saturation current.
The drain current does, in reality, increase somewhat with increasing vp. This
can be attributed to the move of the pinch-off point toward the source for increasing
up and hence to the shortened channel; as Eq. (2.8) indicates, ip increases as L is
reduced. As an approximation, this effect (often called channel-length modulation)
can be included in the formula for ip (vp) in the form of an added factor (1 + dup).
Here is a device constant that depends on L, on the doping concentration of the
substrate, and on the substrate bias (discussed in the next section). For L ~ 10 ym,
typically X ~ 0.03 V-); generally, & WL.
It is usual to introduce the abbreviations k”
the saturation current given by Eq. (2.8) becomes
HnCon/2 and & 2 K (W/L). Then
ip = kag — VrPC + dep), ug = Vn (2.9)
which incorporates channel-length modulation. Figure 2.9 shows the variation of ip
with vg for constant vp. Figure 2.10 illustrates its dependence on up for various ug
values, Where ¥c1 < Ug2 < Vox ~
All derivations of this section were performed for the structure shown in Fig.
2.7, whose source, drain, and channel were all n type. This device is called an n-
channel MOS, or NMOS transistor. A similar arrangement can be constructed by
creating p* drain and source diffusions in an n-type substrate. Now a negative ug
is needed to create a p-type channel under the gate, and a negative up is used to26 MOS DEVICES AS CIRCUIT ELEMENTS,
‘or
Figure 2.9. Drain current versus gate
voltage characteristics of an MOS tran- * ‘a
sistor. T
attract the holes in the channel to the drain. Also, ip will be negative if the reference
direction of Fig. 2.7 is used. The resulting device is called a p-channel MOS or
PMOS transistor. Formulas (2.3) to (2.9) remain valid if some small changes are
made. The mobility 41, of electrons must be replaced by p1», the hole mobility in
the channel. As would be expected from the more elaborate mechanism of hole
conduction, jt, < }1,. Typical mobility values in the channel region for an impurity
concentration of 10'° cm™? are y,, = 1000 cm7/V « s and pp, = 400 cm7/V - s. The
electron charge density Q, in the channel is to be replaced by @,, the hole charge
density; also, a negative sign must be included in Fags. (2.6) to (2.9) to account for
the change in the charge of the carriers. Finally, vp must be replaced by |vp| in Eq.
(2.9), since now vp < 0. In conclusion, Eq. (2.7) becomes
ip = —2k (x —Vr- “| vp. (2.10)
Here k A ,C,,WI2L and Vr <0, Equation (2.10) describes the drain current charac-
teristics in the linear range. The behavior of ip in the saturation region can be
obtained by modifying (2.9):
ip = —k(0g — VrP(L + > fen)). (11)
'0 Vp = Yq Vr
linear -— | > saturation region Yee
region
Figure 2.10. Drain current versus
drain-to-source voltage characteristics V@1 \p
of an MOS transistor.23. MOS TRANSISTOR TYPES: BODY EFFECT 27
rain drain
gate ----« bulk or gate bulk
(a) NMOS
drain
or gate bulk.
‘source
(8) PMOS
drain
gate —F--.+ bulk
‘source
(e) Mos
Figure 2.11. Transistor symbols.
The circuit symbols used for NMOS and PMOS transistors are shown in Fig. 2.11a
and 6, respectively. If the type is unimportant, the simplified symbol of Fig. 2.11c
may be used for both NMOS and PMOS devices.
Since the operation of the devices described in this section is dependent on the
electric field induced by the gate voltage, they are called field-effect devices (FETs),
or MOSFETs.*
ince PMOS transistors are more easily fabricated than NMOS transistors, they
ally predominant. However, later, when the techniques for the reliable
production of NMOS devices were developed, the latter became standard. The main
reason for this is the higher mobility of electrons, which makes the NMOS transistors
faster than PMOS transistors.
2.3. MOS TRANSISTOR TYPES: BODY EFFECT
The MOS transistors described in Section 2.2, both NMOS and PMOS types, share
several features. In the structure, the gate is insulated electrically from the rest of
* Since the charge carriers here are either electrons or holes (not both), FETs are also sometimes called
unipolar devices, to contrast them with bipolar transistors, in which both electron and hole currents exist.28 MOS DEVICES AS CIRCUIT ELEMENTS:
—{ Ho depletion MOSFET
Figure 2.12, Symbols for depletion-
mode transistors.
the device by the SiO, layer under it, Hence it is often called an insulated-gate
field-effect transistor (IGFET). Also, the voltage vg induces and enhances the drain
current. Thus the devices described operate in the enhancement mode.
Its also possible to fabricate an MOS transistor that conducts drain current when
vg = 0. For example, an n-type layer can be introduced by doping, which connects
the source and drain of an NMOS device. With such a doped channel, the field of
the gate is not needed to produce an inversion layer: the region R (Fig. 2.7) now
has a ‘‘built-in’’ conducting n-type channel.
However, if a negative gate voltage is applied, the field thus created will repel
electrons and create a depletion layer in the channel adjacent to the SiO, surface,
thereby reducing the conductivity and thus the drain current. If the magnitude of
the negative gate voltage is sufficiently large, the channel becomes completely de-
pleted and ip ~ 0 results. The value ve at which this occurs is again called threshold
voltage and is denoted by V7. Now, however, Vr < 0. Such a device is called a
depletion-mode FET.
It should be noted that even without a doped layer. an NMOS transistor can
conduct for vg = 0 due to oxide charges [1, Sec. 7.4] if the bulk is very lightly
doped. It is also possible to create a depletion-mode PMOS device, with V; > 0,
by establishing a p-type doped channel.
The relations (2.6) to (2.11) remain valid for depletion-mode devices if the value
and sign of Vr is chosen appropriately, as described above. Two symbols often used
to denote depletion-mode MOSFETS are shown in Fig. 2.12.
A totally different structure can also be used to produce a depletion-mode field-
effect transistor (Fig. 2.13). Here, a lightly doped n-type layer (channel) connects
the n* source and drain regions and the gate is a p* region implanted in this layer.
Hence, for vs = vg = 0 and v¢p > 0, a drain current will flow. If ug is made nega-
tive, the p* implant acting as the gate will be surrounded by a depletion layer;
the greater |uc|, the deeper the layer. The mobile electrons in the channel cannot
Figure 2.13. Junction field-effect transistor.2.3. MOS TRANSISTOR TYPES: BODY EFFECT 29
enter this depletion layer, or the one along the pn junction between the channel and
the substrate. Hence the effective cross section of the channel is reduced as |vg| is
increased, At some value u = Vp (Vp < 0), ip becomes zero (in practice, < 1
pA). Thus Vp plays the same role as Vz for a depletion-mode MOSFET; it is called
the pinch-off voltage. It can be visualized as the gate voltage, which causes the two
depletion regions in the channel to merge, leaving no conductive path between source
and
The device described and shown in Fig. 2.13 is called an n-channel junction field-
effect transistor FET), since its gate is separated from the rest of the device by a
reverse-biased pn junction rather than by an SiO, layer as for the MOSFET (IGFET).
Since the JFET is hardly ever used in analog MOS integrated circuits, it is not
discussed in detail here. A clear description of its physics and current—voltage charac-
teristics is given in Sec. 2.5 of Ref. 1.
Next, a key limitation (called the body effect) of MOSFETs used as analog circuit
elements is described. In the discussion in Section 2.2 it was always assumed that
both the bulk and source are grounded, so that vp = ws = 0 held. Often, circuit
considerations make this convenient arrangement impossible and us # vy must be
used. Obviously, the voltage uy — vg must be such that the source—bulk junction
is reverse biased, otherwise, a large junction current will flow inside the transistor.
This current may damage the device, and in any case will impede its proper operation.
Thus, say, in an NMOS transistor, the bulk must be biased to make it negative with
respect to both source and drain.
If the source potential is not zero, the voltages vg and vp must be replaced in all
equations by rigs = tg — us and tp5 = Up — wy, respectively. In addition, the
depletion region around the channel (Fig. 2.8) will become wider if the reverse
voltage between the bulk and the source (and hence the channel) is increased. Since
the voltage ue = Vy is the gate voltage necessary to maintain the depletion region
(without creating a channel), Vr will increase in magnitude. The dependence of Vr
on the voltage vsz 4 vs — vp can be shown [1, Sec. 8.2] to be in the form
[vr] = [rol + YV2l6r1 + [esa] — V21bp))- (2.12)
Here, V7 is the threshold voltage for usg = 0 and ¥ is a device constant given by
V2€.qM imp
(2.13)
Ox
In Eq. (2.13), és 18 the permittivity of silicon: ey = €oKs, Ks ~ 11.7. Also, Nimp
is the density of the impurity ions in the bulk. For NMOS, Nimp = Na, the acceptor
ion density; for PMOS, Ninp = Np, the donor ion density. For example,
for Ninp = 10'S cm™* and 800 A oxide thickness (i.e., Cy, * 4.4 x 107 F/em?),
y ~ 0.423V!”, Finally, $, is a material constant of the bulk; its value is around
03 V.
* , 2 (E, — jg, where £, is the intrinsic Fermi energy and E, the Fermi energy of the semiconductor
U1, p. 318}30 MOS DEVICES AS CIRCUIT ELEMENTS
TABLE 2.1. Key Units and Constants for MOS Transistors
1pm = 107¢em = 10°A
1 mil = 25.4 pm = 0.0254 mm
Electron charge (magnitude): ¢ = 1.6 x 107 C
Permittivity of free space: € 86 x 10-' F/em
Permittivity of silicon: ¢ = Ks = 1.04 x 107 !? Ffem; Ky = 11.7
Permittivity of silicon dioxide: €,, = 6K. = 3.5 X 10-" Flem; K,. = 3.9
Oxide capacitance: Cy. = éx/fox 5X 107 87et Flem*
Intrinsic carrier concentration: n, = 1.5 X 10! em~?, T = 300 K
Boltzmann’s constant: k = 1.38 x 10-* J/K; kTVq (at T = 300 K) = 0.026 V
Electron mobility in Si (Nimp = 10'7 cm, T = 300 K): 670 cnv/V - s
Hole mobility in Si (Nimp = 10'’em=, T = 300 K): 220 om2/V - s
Body-cffect coefficient: y = \/7MaMen fox 1.67 x 10-3 sp VNB V"
Bulk potential: y = — jy Mine = 0.026 In (0.67 x 10>" NE)
This phenomenon, the body effect, is a major limitation of MOS devices operated
with vs # vp; its evil influence will be lamented repeatedly later in the book. As
Eqs. (2.12) and (2.13) show, to reduce the body effect, Nimp Should be made small.
However, for very small Ny values (say, N4 < 10!3 cm~7), an NMOS may behave
as a depletion-mode device, as explained earlier. Thus the body effect cannot be
climinated completely. Some key constants and formulas on MOSFETs are summa-
rized in Tables 2.1 and 2.2.
2.4. SMALL-SIGNAL OPERATION AND EQUIVALENT CIRCUIT OF
MOSFET TRANSISTORS
Earlier the physical principles and basic operation of MOS transistors were discussed
briefly, and formulas derived that gave the drain current as a function of the voltages
and/or currents at the various terminals of the device. In these earlier discussions it
was assumed that all voltages and currents were constant or that they varied suffi-
ciently slowly so that all capacitive currents (and hence all capacitances themselves)
could be neglected in the discussions. On the other hand, the formulas derived were
valid for large as well as small voltage and current variations.
In many important linear applications (such as operational amplifiers, discussed
in Chapter 4) the voltages and currents of the transistor vary so rapidly that capacitive
effects cannot be ignored, and hence the capacitances of the device must be included
in the analysis. At the same time, in such circuits the signals are sufficiently small,
so that linear approximations may be used in all nonlinear relations. This simplifies
the equations and permits the use of linear models (simple linear equivalent circuits)
for these nonlinear devices.(ble A — 852 — | @lzAyh — e-"¢4A) = 4A
“ctTNen-CA — 2%2—),_T © X ary
“ Te.
CN — 1) oth — $0) Fasoag = OF
(*p pue & jo sonqea up 405 1°7 9198 998)
(IZA — #8 + L2@1Z AA + 9 C4A) = 4A
‘TNen CA + 2%), 7 2» ¥ a1
saay + 1) ta — 8) —Te_ =
(stay TCA ) aA "
Fal - [2a] < [saa]
‘Hal < [|
oar uonemieg
Fal ~ [$a] > baal
‘Hal < [$a]
sa, (& a =) Joye = 4 onl -!n a) Tega = a4 :uorax apouy.
SOWd SOWN uogesdQ jo uoI3zay
wopeiedg Kouambery-ao7y [eusg-a81e7] Uy SLAISOW 409y suopepy w9sINy-UPIG “TZ ATAVL
a32 MOS DEVICES AS CIRCUIT ELEMENTS:
In the following discussion, we concentrate on the linearized approximation and
modeling process for MOSFETs operating in their saturation regions, which is the
usual condition for linear (analog) operation. Afterward, we give a brief overview
of the linearization and modeling for devices that operate in their triode (nonsatu-
rated) or cutoff region. Assuming an NMOS transistor, and combining Eqs. (2.9)
and (2.12), the relation
ip = Kues — Vy ~ ¥WV20_ + vse + ¥V2b,)°CL + dvps) (2.14)
results, Here we used ucs & ug — us and ups A up — ws to replace vg and up,
since in general vs # 0. For small variations of ip, vcs, Ups, and vsg, the nonlinear
expression (2.14) can be replaced by a first-order Taylor approximation. Specifically,
near a constant bias point i9, = f (ves, v2s, vs) we can write
9 dip
= 72
B+ Mip=B+
iD ip = Ip (22
(2.15)
Here (dip/Augs)° and so on denote the partial derivatives evaluated at the bias point.
Aip is the deviation (increment) of ip from its bias value; Aves, Aups. and Ausa
are the increments of vgs, ups, and usp. All deviations must be small for Eq. (2.15)
to hold. If only the incremental (small-signal ac) components are of interest, Eq.
(2.15) can be written as
Min = 8m Aves + 84 Avps + 8m» Ausp, (2.16)
fe
Figure 2.14. Low-frequency equivalent circuit of a MOSFET.2.4. SMALL-SIGNAL OPERATION AND EQUIVALENT CIRCUIT OF MOSFET TRANSISTORS 33.
where
(2.17)
dip
&mS (2) ,
Here g, is the (incremental) drain conductance: g,, and g,,. are transconductance,
that can be represented by voltage-controlled current sources (VCCSs). Hence an
equivalent-circuit model, shown in Fig. 2.14, can be constructed. The values of gm,
Rmb. and gq can be found from Eq. (2.14):
= 2k obs — Vo — ¥V26, + Bo + YV26)U + dvds)
= 27k + AWB), (2.18)
dip
Bmp pe) = —Kuks — Vio — ¥W2bp + Ve + V2)
x (L + Av) = (2.19)
°s V2b, + vfs
= 8m/2
© V2b, + Be”
on
wot (2) = ks — Vio — YV2b, + Be + YV20,)°A (2.20)
Oups;
= a 9,
T+ pas?
Hence, to a good approximation, gq and g,,» are proportional to ‘Vip, while gy is
proportional to i}.
The other important components of the complete small-signal model of the MOS-
FET are the capacitors representing the incremental variations of stored charges with
changing electrode voltages. These play an important role in the high-frequency
operation of the device. The intrinsic components of the terminal capacitances of
the MOSFET devices (associated with reverse-biased pn junctions and with channel
and depletion regions) are strongly dependent on the region of operation, while the
extrinsic components (due to layout parasitics, overlapping regions, etc.) are rela-
tively constant. Assuming again that the transistor operates in the saturation region,
it can be assumed that the channel begins at the source and extends over two-34 = MOS DEVICES AS CIRCUIT ELEMENTS
thirds of the distance to the drain. In this region of operation. the most important
capacitances are the following:
Cys: Gate-to-Drain Capacitance. This is due to the overlap of the gate and
the drain diffusion. It is a thin-oxide capacitance. and hence to a good approxi-
mation can be regarded as being voltage independent.
2. Cys: Gate-to-Source Capacitance. This capacitance has two components:
Cgsov. the gate-to-source thin-oxide overlap capacitance, and Cy,., the gate-to-
channel capacitance. The latter (in the saturation region) is around 3C,., where
C,. is the total thin-oxide capacitance between the gate and the surface of the
substrate. In the triode region, C;, = Cox. Ces is nearly voltage independent
in the saturation region.
3. Cyp: Source-to-Substrate Capacitance. This capacitance also has two compo-
ents: Cyn, the pn junction capacitance between the source diffusion and the
substrate, and C,,, which can be estimated as two-thirds of the capacitance
of the depletion region under the channel. The overall capacitance Cy, has a
voltage dependence which is similar to that of an abrupt pn junction.
4. Cy: Drain-to-Substrate Capacitance. This is a pn junction capacitance and
is thus voltage dependent.
5. Cy»: Gate-to-Substrate Capacitance. This capacitance is usually small in the
saturation region; its value is around 0.1C,,-
Figure 2.15 illustrates the physical structure of an NMOS transistor and the locations
of the capacitances in the cutoff (Fig. 2.15a). saturation (Fig. 2.155). and nonsatura-
tion or triode (Fig, 2.15c) regions. Table 2.3 lists the terminal capacitors of the
NMOS device and their estimated values in the three regions of operatian The
notations used are those shown in Fig. 2.15u to c. Figure 2.16 depicts the complete
high frequency (ac) small-signal mode! of the MOSFET. In analyzing the small-
signal behavior of MOSFETs. the model of Fig. 2.14 can be used if only low-
frequency signals are present: if the capacitive currents are also of interest. the circuit
of Fig. 2.16 must be applied
From the models of Figs. > 14 and 2.16 and accompanying discussions, a number
of general statements can be made about the desirable construction of a MOSFET:
1, For high ac gain. g,, should be large. This will be the case, by Eq. (2.18). if
£ 2B (WrCoxWV/2L 1s large. Thus the oxide should be thin to maximize C,,
(which is the oxide capacitance per unit area); also, W/L should be as large
as possible. These measures, however, tend to increase the size and thus the
cost of the integrated circuit. Also, by Eq. (2.18), the quiescent (bias) current
i, should be as large as the allowable de power dissipation permits.
2. As the negative sign in Eq. (2.19) indicates, the body effect reduces the gain.
To minimize g,., by Eqs. (2.19) and (2.13), we need large Cyx. small Nimp
(ie.. lightly doped substrate), and a large bias voltage v%g for the source.2.4. SMALL-SIGNAL OPERATION AND EQUIVALENT CIRCUIT OF MOSFET TRANSISTORS = 35,
S t
ception | EsL-2Lov
region Cebpn p - substrate
@
source side —,~ drain side
t
inversion
layer —_p - substrate
©
Figure 2.15, Parasitic capacitances in MOSFET in the (a) cutoff region, (b) saturation region,
and (c) triode region.
(Of course, if usp is constant, no incremental body effect occurs and these
requirements are irrelevant.)
3. Ideally, the MOSFET in saturation should behave as a pure current source.
Hence, as Fig. 2.14 illustrates, g, should be small. By Eq. (2.20) this requires
a small bias current i%, a large bias voltage «ps, and a small A. Since A is
introduced by channel-length modulation, it can be reduced by increasing L
and also [1, Sec. 8.4] by increasing Nimp. A summary of the formulas derived
in this section is given in Table 2.4.“IZ — 1 SUP (Ty
A" TMG Ay“ “O/TMt oar
+ PAMory + AM OW 0 (IE + “TOM. + TMH (@porn) paremesuony
CAO TMS ayy +9
Ary + AO" (PAYOO TAR TM uorgex tonemreg
wary aory “OK “OTM “TM wordar JJOI)
wy "2 #9 ae “D uonesadg jo uolsay
aouensede3,
pMopEtedO Jo suoyBay UP) 224, 24) UN) LAASOW # JO saouNpedsy jeaULEL, “¢-7 ATAVL
362.4. SMALL-SIGNAL OPERATION AND EQUIVALENT CIRCUIT OF MOSFET TRANSISTORS 37.
@
Figure 2.16. High-frequency equivalent circuit of a MOSFET.
Next, we discuss briefly the linear model of a MOSFET that is biased in its nonsatu-
rated (triode) region. Usually, such a device is used as a switch that is opened or
closed (tumed on or off) by a large gate voltage or as a fairly linear large-valued
variable resistor. Hence, here we derive its model only to analyze its behavior in
such applications. We assume that ves is constant and that 2m» < ga and hence
negligible. Under these conditions it can be seen from Eq, (2.7) that when the device
conducts drain current, it behaves like a resistor connected between the drain and
source terminals. The equivalent small-signal drain-to-source resistance for the case
of ups near zero is given by
1 1
ra 94, ~ Walls = Veo) (2.21)
and is thus controlled by the gate voltage overdrive v%s — Viv.
In high-frequency application, the device capacitances must also be included in
the model. A simple equivalent circuit is shown in Fig. 2.17. Since now a continuous
channel extends from the source to the drain, the gate-to-channel capacitance Ci,
is connected to both the drain and the source. An accurate high-frequency representa-
tion of the channel should include a distributed resistive line extending from the
source to the drain and capacitively coupled to both the gate and the substrate.
However, as a first approximation, C4, can be treated as a lumped capacitance
partitioned equally between C,, and C,¢, as indicated in the last row of Table 2.3.
Finally, if the device is cut off, no channel exists and the model contains only
Cyd Db Cap
6 tds B
Ci
a (Csb Figure 2.17. High-frequency model of a MOSFET in its
s triode region.“Se[musoy [fe UI pouinsse st | > FAY “OTT “Biy eag,,
“gery — 1A “eure - 1h
“ouig + A “oye + IN
(1?9)""2 aoueyoedes
oD . OD Wi Wi ‘yinq-01-(ureup Jo) eamog
depzoao P*D, dejsano 8 coumpoedes uresp-o1-21e5)
“OME “TMS *%y sooysedva s0sn0s-0- 910)
SON = 1 SHY +1 S00 = ny
ea oR ae
:goueionpuoo urea
— “eA, a 4 a0 =
ng BABA wg AA ne 5 mg
tuk wk Ne
rsouepnpuossuen 1995}2-Apog,
song =
= = 44 — 5) 2 1 = (4 — 8p) 7 ne Sue
eA = 9) ao Taroe|\ > U8 — Goo aig ©
:eoueronpuoosues
SOWd SOWN RULE,
MONEINES UI SIAASOW JO SiaysmEseY [eusis-[ewS “PZ TTAVL2.5. WEAK INVERSION 39
Cad p Caw
Cap
a B
Soe Cob Figure 2.18. High-frequency model of a MOSFET in its
s cutoff region.
the capacitances, with the values listed in the first row of Table 2.3. A simplified
model of a MOSFET jin the cutoff region is shown in Fig. 2.18, where the
drain—sourve resistance is infinity and C,, and Cy, capacitors are due w gate overlap
and fringing capacitances. The gate-to-substrate capacitance C,a in the cutoff region
is, however, highly nonlinear, and for the gate-to-source voltage around zero its
value is approximately equal to WL'C,,, where L’ = L — 2, and Loy is the length
of the overlap between the gate and the source-drain diffusion regions.
2.5. WEAK INVERSION
‘The triode and saturation regions of operation discussed earlier in this chapter assume
that the device is operated in strong inversion and Ves — Vz = 100 mV (for an
NMOS wansistor). If Ves — Vr < 100 mV, the device is in the weak inversion
region (also called the subthreshold region) and operation of the n-channel MOS
transistor with the source connected to substrate is more accurately described by
the following exponential relationship between the gate-to-source control voltage
and drain current:
Ip= © igetteet, (2.22)
where
1 nk: . — (gV/nkT + I)
Ion = wobor mg ("q) orem 2.23)
and the parameters m and n are defined in terms of the various capacitances in the
device [3]. Equation (2.22) assumes that the device is operated with Vps > kT/q.
The MOS transistors operating in the weak inversion region, similar to bipolar
devices, have an exponential relationship. However, since Ino is very small (on the
order of 10 to 20 nA), the available current to charge and discharge capacitances is
also small, resulting in poor frequency performance. In practice, MOS transistors
are operated in weak inversion only in low-frequency applications when low power
consumption is desired.40 MOS DEVICES AS CIRCUIT ELEMENTS
2.6. IMPACT IONIZATION [4]
One of the severe problems in submicron MOS technologies operating at supply
voltages around 5 V is impact ionization. Figure 2.19 illustrates an n-channel MOS
device cross section showing the impact ionization current flow and the J-V charac-
teristic as a result of impact ionization. As depicted in the figure. when the drain-
to-source voltage is increased, the strength of the electric field at the drain end of
the channel eventually becomes high enough to induce significant impact ionization
current which originates from the drain depletion region and flows into the substrate.
Once this happens the current that flows into the drain terminal has two components.
One component is the MOS transistor channel current that flows from the drain to
the source, and the other is the impact ionization current that flows from the drain
to the substrate. The impact ionization current is not a function of the wansistor
channel length, and the magnitude of the current is not reduced dramatically simply
by making the length longer. The current is largely determined by the peak electric
field, which in tum is a function of the gate oxide thickness, drain junction depth,
doping concentration in the substrate, the voltage between the drain terminal and
the drain end of the channel region, and the gate-to-drain voltage. In technologies
with feature sizes in the range of 2 jum, for an n-channel MOS device. the impact
ionization current equals 1% of the drain current when the voltage between the drain
and the drain end of the channel is in the range 4 to 9 V, and the device is biased
los
wo
Figure 2.19, (a) An n-channel MOS device cross section showing umpax wmuzaon >rrent
flow; (b) -V characteristic observed as a result of impact ionizazoe,2.7. NOISEINMOSFETS 41
in the saturation region. In p-channel MOS devices the effect occurs at substantially
higher field strengths.
‘The impact ionization has several potentially damaging side effects. One serious
negative consequence shown in Fig. 2.19b is degradation of the transistor output
impedance, which results in reduced gain in amplifier stages that use transistors as
active loads. One way w deul with this problem is dough circuit wechniques, where
a shielding n-channel device is placed in series with the tansistor, preventing it
from having a Vs greater than half the supply voltage [4]. The second damaging
effect is the possibility of triggering latchup duc to the ohmic drop induced by the
ionization current that flows into the substrate. Latchup is a phenomenon caused by
the parasitic lateral pnp and npn bipolar transistors created on the chip. The collectors
of each transistor feed the other’s base, and this creates an unstable device similar
to a pnpn thyristor [5]. This causes a sustained de current that may cause the chip
to stop functioning and may even destroy it. Latchup may be prevented by proper
substrate strapping and using guard rings to surround some critical transistors on
the chip. Another strategy is to reduce the substrate resistance. In this method the
p- and n-channel transistors are formed in a lightly doped epitaxial layer that is
grown on a low-resistivity substrate. Finally, the third serious side effect of impact
ionization is the threshold shift of the MOS device due to the continuous operation
in the impact ionization mode. This phenomenon is due to the high electric field
which creates high energy carries that can be trapped in the gate oxide, resulting in
long-term threshold shift. Several process modifications have been proposed that
are effective at raising the voltage at which impact ionization becomes a problem.
One technique is to lower the impurity gradient in the drain junction using lightly
doped drain (LDD) structures.
2.7. NOISE IN MOSFETS
There are three distinct sources of noise in solid-state devices: shot noise, thermal
noise, and flicker noise.
Shot Noise
Since electric currents are carried by randomly propagating individual charge carriers
(electrons or holes), superimposed on the nominal (average) current J, there is always
a random variation i,s. This is due to fluctuation in the number of carriers crossing
a given surface in the conductor in any time interval. It can be shown that the mean
square of i,s is given by
iE, = ql Af, (2.24)
where q = 1.6 X 107! C is the magnitude of the electron charge and Af is the
bandwidth. This formula only holds, Lowever, if the density of the charge cantiers
is so low and the external electric field is so high that the interaction between the42 MOS DEVICES AS CIRCUIT ELEMENTS
DI tar
oe (b) fe)
Figure 2.20. Thermal noise in a resistor: (a) noisy resistor; (b) and (c) equivalent circuits.
camers 1s negligible. Otherwise, the randomness of their density and velocity is
reduced due to the correlation introduced by the repulsion of their charges. The
noise current is then much smaller than predicted by Eq. (2.24).
In a conducting MOSFET channel, the charge density is usually high and the
electric field is low. Therefore, Eq. (2.24) does not hold. The noise current due to
random carrier motions is hence better described as thermal noise, which is discussed
next.
Thermal Noise
In a real resistor R, the electrons are in random thermal motion. As a result, a
fluctuating voltage v,r appears across the resistor even in the absence of a current
from an external circuit (Fig. 2.20a). Thus the Thévenin model of the real (noisy)
resistor is that shown in Fig. 2.20b. Clearly. the higher the absolute temperature T
of the resistor, the larger v7 will be. In fact, it can be shown that the mean square
of ur is given by
ar = AKTR Af. (2.25)
Here k is the ubiquitous Boltzmann’s constant, and Af is the bandwidth in which
the noise is measured. in hertz. (The value of 4k7 at room temperature is about 1.66
x 107 V Cy
If Eq. (2.25) was true for any bandwidth, the energy of the noise would be infinite.
In fact, however, for very high frequencies (~10'? Hz) other physical phenomena
enter, which cause ver to decrease with increasing frequency so that the overall
noise energy is finite.
The average value (dc component) of the thermal noise is zero. Since its spectral
density v2p/Af is independent of frequency (at least for lower frequencies), it is a
white noise. Clearly, Fig. 2.20b may be redrawn in the form of a Norton equivalent
that is as a (noiseless) resistor R in parallel with a noise current source i,r (Fig.
2.20c). The value of the latter is given by
By = 4kTG Af, (2.26)
where G = UR.2.7. NOISE INMOSFETS 43
noiseless
transistor
(a)
wo
Figure 2.21. Equivalent models of the thermal noise in a MOSFET.
Since the channel of a MOSFET in conduction contains free carriers, itis subject
to thermal noise. Therefore, Fas. (2.25) and (2.26) will hold, with R given by the
incremental channel resistance. The noise can then be modeled by a current source,
as shown in Fig. 2.21a. If the device is in saturation, its channel tapers off (Fig.
2.8) and the approximation R ~ 3/2g,, can be used in Eq. (2.26).
In most circuits it is convenient to model the effect of i,r caused by a voltage
source connected to the gate of an (otherwise noiseless) MOSFET (Fig. 2.21). This
“gate-referred”” noise voltage source is then given by
(2.27)
Both i,7 and v,r depend thus on the dimensions, bias conditions, and temperature
of the device. As an example of their orders of magnitude, fora transistor with W
= 200 pm, L = 10 wm, and Cop = 4.34 X 1078 F/em? (corresponding to an oxide
thickness of 800 A) which is operated in saturation at a drain current i, ee 200 pA,
the gate-referred noise voltage at room temperature is about 9 nV/\/H:
If the device is switched off, R becomes very high, and the equivalent noise
circuit will be a current source with a value given by Eq, (2.26). Clearly, 27 is very
small; hence for usual (low or moderate) external impedance levels, the MOSFET
can be regarded as a noiseless open circuit if it is turned off.44 — MOS DEVICES AS CIRCUIT ELEMENTS
Flicker (1/f) Noise
In an MOS wansistor, extra electron energy states exist at the boundary between the
Si and SiO,. These can trap and release electrons from the channel, and hence
introduce noise [6.7]. Since the process is relatively slow, most of the noise energy
will be at low frequencies. As before. a possible model of this noise phenomenon
is a current source in parallel with the channel resistance. The de value of noise
current is again zero. [ts mean-square value increases with temperature and the
density of the surface states: it decreases with the gate area W X L and the gate
oxide capacitance per unit area C,.. For devices fabricated with a “‘clean”” process,
the gate-referred noise voltage is nearly independent of the bias conditions and is
given by the approximating formula
K_ sf
ty = a
= GWE (2.28)
Here K depends on the temperature and the fabrication process; a typical value [8,
p. 31} is 3 x 10°74 V° - F. For the transistor described in the preceding example,
the formula gives a noise voltage of 83 nV/\/Hz at f = 1 KHz. As before, the
equivalent channel current noise is related to vr by the formula iy = gritty.
The noise process described is usually called flicker noise or (in reference to the
1/f factor in v3,and i3,) Lf noise. As the example given illustrates, at low frequencies
(say. below | kHz) it is usually the dominant noise mechanism in a MOSFET.
In conclusion, the channel noise in a MOSFET can be modeled by an equivalent
noise current generator, as in Fig. 2.214. In the small-signal model this generator
will be in parallel with the current sources gnUes and gmotas (Fig. 2.16). Its value
can be chosen as the root-mean-square (RMS) noise current, which from Eqs.
(2.26)-(2.28) is
Kaz
= AKTG + WH) Af.
Note that the mean squares of the noise currents are added, since the different noise
mechanisms are statistically independent. Alternatively, the noise can be represented
by its gate-referred voltage source (Fig. 2.216), in series with the gate terminal. The
value of the source is i,,/g,,, with i, given by Eq. (2.29).
(2.29)
2.8. CMOS PROCESS
The CMOS process provides the most flexibility to the circuit designer, due to the
availability of complementary MOS devices on the same chip. The original motivation
for developing the CMOS technology was the need for low-power and high-speed
logic gates for digital circuits. The required isolation between the two different device
types is accomplished by the use of ‘‘wells,”” that is, large, low-doping-level deep
diffusions, which serve as the substrates for one of the twa device types. Asan example,
Fig. 2.22 shows part of an n-well CMOS chip, where high-resistivity p-type substrate
is used for the n-channel devices, and diffused n wells for the p-channel devices.2.8. CMOS PROCESS 45
deposited oxide metal
Pp
\ fold * " well
field implant
polysilicon 2}
polysilicon 4
> si
Figure 2.22, Device structure fabricated in a high-performance n-well CMOS process.
As will be shown in later chapters, a CMOS circuit can be operated with a single
power supply, and it can be used to realize high-speed, high-gain, low-power analog
amplifier stages. An additional advantage is that for the devices in the well (in Fig.
2.22, the PMOS transistor), the source can be connected to the well, thereby eliminat-
ing the body effect, and if the device is used in an amplifier, increasing the gain of
the circuit. This, however, results in a large stray capacitance between the source and
the substrate, due to the large size of the well-to-body interface. Another important
advantage of the CMOS process is the availability of transmission gates madc of a
parallel connection of complementary transistors that can be used as switches. When
such transmission gates are used, the signal is no longer limited to a level, which
is a threshold voltage below that of the high clock signal, as is the case when single-
channel switches are used. In addition, in CMOS chips a bipolar transistor can be
fashioned from a source diffusion, the well, and the substrate. This can be used in
an emitter-follower buffer stage (described later), in a bandgap voltage reference
circuit, and so on.
In addition to transistors, analog MOS circuits usually require on-chip capacitors,
and sometimes also on-chip resistors. In a silicon-gate ‘‘double-poly”’ process, a
second layer of low-resistivity polysilicon is available for usc as an interconnect or
for the formation of a floating gate for memory applications. These two layers of
polysilicon can also be used as the top and bottom electrodes of a monolithic capaci-
tor. Figure 2.22 shows the construction of a capacitor with two polysilicon electrodes.
Resistors can be created on an MOS chip using a diffused or implanted layer on
the surface of the substrate. Since the sheet resistance of these resistive layers is
relatively low (typically 25 to 70 for a square layer), the size of the resistors
obtainable on a reasonably small area is limited to about 100 k®. The higher resistiv-
ity well diffusion is also available as a resistor. This resistor, however, has much
higher voltage and temperature coefficients compared to diffused or implanted ones.
PROBLEMS
2.1. A pn junction diode is connected to an external voltage vin the forward direction
(Fig. 2.3), Reversing the polarity of the voltage reduces the current by a factor46
Figure 2.23. Circuit for Problem 2.8.
2.2.
2.3.
24.
2.8.
2.6.
2.7.
2.8.
2.9.
MOS DEVICES AS CIRCUIT ELEMENTS
Yoo wo
1
10°. Assume that the diode satisfies Eq. (2.1) and is at room temperature. What
is v?
For a pn junction (Fig. 2.3), Ns = Np = 10'° ions/cm, | = 5V,A = 0.34
mm?, and the measured value of C is 27 pF. How much is x4, the width of the
depletion layer? How much is 6,?
Using the definition R = 1 /(ip/8Vps), calculate the channel resistance of an
NMOS transistor from (a) Eq. (2.6), (b) Eq. (2.7), and (©) Eq. (2.9).
For an NMOS wansistor, p,, = 10° cm?/V-s, the thickness of the gate oxide is
10° A(1A = 107% cm), W = 25 jum, and L = 5 ym. The threshold voltage
is 4 V. Calculate ip for us = vp = 0 V and ug = 6 V, and (a) up = 0.1 V,
(b) vp = 2 V, and (c) wp = 4V.
Repeat the calculations of Problem 2.4 if uv; = 0, vs = —3 V, and $, = 03
V. What conclusions can you draw from your results regarding the body effect?
For an NMOS transistor,” = 2 wA/V2, W = 30 um, L = 10 wm, bp = 03
V, y = 15 V'?, and X = 0.03 V~!. Find the incremental conductances Sm»
Gus and By for usg = OV, vs = 5 V, and 7% = 10 WA. Repeat your
calculations for vga = 2 V!
‘An NMOS switch transistor has a gate-to-source voltage ucs > Vr. Its drain
is open-circuited. How much is vps? Why?
In the circuit of Fig. 2.23, the switch 5 is opened at 1 = 0. (a) Is the transistor
operating in its linear or saturation region? (b) Neglecting body effect and
channel-length modulation, find u(t) by solving the appropriate differential
equation for the circuit.
In the circuit of Fig. 2.24 a noise voltage v, is generated due to thermal and
ka.
wv
Figure 2.24, Circuit for Problem 2.9.2.8. CMOSPROCESS 47
@i ®@i @i
’ eg
®
®@
+
+
:
Figure 2.25, Circuit for Problem 2.10.
shot-noise effects. For what value of R will the two noise voltages v,r and v,s
be equal?
2.10. Calculate the incremental impedance Ou/di seen at node A of the circuits
shown in Fig. 2.25.
2.11. Show that the transconductance g,, in the saturation region is equal to the
drain conductance in the triode region for a given device and a fixed Ve
REFERENCES
1, R. S."Muller and T. I. Kamins, Device Electronics for Integrated Circuits, Wiley, New
York, 1977.
2. A.S. Grove, Physics and Technology of Semiconductor Devices, Wiley, New York, 1967.
3. Y.P. Tsividis and R. W. Ulmer, A CMOS voltage reference, IEEE J. Solid-State Circuits,
SC-13(6), 774-778 (1978).
4. C. A. Laber, C. F. Rahim, 8. F. Dryer, G. T. Uehara, P. T. Kwoh, and P. R. Gray, /EEE
J. Solid-State Circuits, $C-22(2), 181-189 (1987).
5. S. M. Sze (Ed.), VESI Technology, McGraw Hill, New York, 1983.
6. M. B. Das and J, M. Moore, IEEE Trans. Electron, Devices, ED-21(2), 247-257 (1974).
7. P. R. Gray and R. G. Meyer, Analysis and Design of Analog Integrated Circuits, Wiley,
New York, 197.
8. PR, Gray and D. A. Hodges, and R. W. Brodersen (Eds.), Analog MOS Integrated
Circuits, IEEE Press, New York, 1980.CHAPTER 3
BASIC ANALOG CMOS SUBCIRCUITS
In this chapter some of the basic subcircuits commonly utilized in analog MOS
integrated circuits are examined. These blocks include a variety of bias circuits,
current mirrors, single-stage amplifiers. source followers, and differential stages.
‘These subcircuits are typically combined to synthesize a more complex circuit func-
tion. The operational amplifier and comparator, covered in later chapters, are exam-
ples of how simple subcircuits are combined to form more complex functions.
‘The first part of this chapter covers the subject of the bias circuits in CMOS
technology and the current mirrors. Next, the CMOS gain stage is introduced, with
particular emphasis on the use of active devices as active loads. The current mirror
subcircuit covered as a biasing clement is utilized as a dynamic load to obtain very
high voltage gains from a single-stage amplifier. The differential amplifier, which
represents a broad class of circuits, is discussed next. The differential amplifier is
‘one of the most widely used gain stages, whose basic function is to amplify the
difference between two input signals. Finally, the last part of the chapter deals with
the small-signal analysis and frequency response of CMOS amplifier stages. A good
understanding of the topics presented in this chapter is essential for the analog CMOS
designer, as most designs start at the subcircuit level and progress upward to realize
a more complex function.
3.1. BIAS CIRCUITS IN MOS TECHNOLOGY
Op-amp and amplifier stages, described in detail later, need various dc bias voltages
and currents for their operation. An ideal voltage or current bias is independent of,
the de power supply voltages (Vpp > 0 and Vsg = 0) and of temperature.
To obtain the dc bias voltages Voi, Vers «... Von Where Vss < Voy < Vo2 <->
483.1. BIAS CIRCUITS INMOS TECHNOLOGY = 49.
drain 'b
Figure 3.1. (a) Diode-connected NMOS transistor; (b) current-voltage characteristic of
diode-connected transistor.
Vin < Vow, voltage division can be used. Pure resistive dividers are seldom used
in MOS technology because the resulting voltages cannot be used directly to establish
bias currents in MOS transistors. Instead, combinations of MOSFETs and resistors
are often used. A MOS transistor with its gate connected to the drain forms a two-
terminal device, as shown in Fig. 3.1a. Its current-voltage characteristics are shown
in Fig. 3.1b. Since Vps = Vos, the dynamic resistance rz, is characterized by
B.D
Therefore, a useful approximation valid for low frequencies, small signals, and negli-
gible substrate effects is that the device behaves like a resistor of value I/g,,. Any
number of n- and p-channel devices and resistors can be combined to form a voltage
divider, as shown in Fig. 3.2, where both n- and p-channel transistors are used
Figure 3.2, Voltage-divider-based bias circuit.5D BASIC ANALOG CuInS SUuRCIAC HTS,
and Vss = 0 is chosen. Since Vas = Vos here for both devices, the condition for
saturation,
Vs > Vos — Vrs G2)
is satisfied, Hence the common value of the drain currents is given approximately
by
| (Vor — Van”
1
mle
Teas = x,
w(t
[ez — Voo) ~ VrnP 3)
a =
Here /eias is usually specified: then V1 and Vjo can be selected and Eq. (3.3) used
to find the W/L ratios of the devices and the value R of the resistor.
An undesirable feature of this configuration is that the bias voltages and current
depend on the supply voltages Vin and Vcs. In fact, the bias current increases rapidly
with increasing power supply voltage. Since such bias strings are used to provide
bias for other devices in the circuit. the de power consumption of the overall circuit
becomes heavily dependent on the supply’ voltages.
‘A CMOS circuit with (theoretically) perfect supply independence is shown in
Fig. 3.3. If Q; and Q, are matched transistors so that (W/L); = (W/L).. they ideally
carry equal currents. Choosing (W/L), = (W/L), will then result in Ves, = Vose,
Figure 3.3. Vreferenced supply-independent CMOS bias
source.3.1. BIAS CIRCUITS IN MOS TECHNOLOGY 51
and the voltage across the resistor, /iiasR, will be equal to Voso. This equilibrium
condition leads to the equation
4 [Te
Rls = Voso = \l epi + Vr- (3.4)
Equation (3.4), which is independent of Von, can be solved to obtain Iyias. Note that
in this analysis we neglected the effects of channel-length modulation (i.e., we as-
sumed that Ip is independent of Vps).
An alternative version of the bias circuit shown in Fig. 3.3 that uses the base-emit-
ter voltage (Vpe) of a bipolar transistor as the reference voltage is shown in Fig.
3.4a. In a CMOS process, the substrate, well, and the source—drain junction inside
the well can be used to form a vertical bipolar transistor. For example, Fig. 3.4b
r-woil
Prsubetrate:
(b)
Figure 3.4. (a) Vge-based supply-independent bias circuit; (b) vertical pnp bipolar transistor
in an n-well CMOS process.52 BASIC ANALOG CMOS SUBCIRCUITS
shows a vertical pnp device that is formed in an n-well process. The collector of
this pnp device is the p~ substrate that is permanently connected to the most negative
voltage on the chip. For the bipolar transistor, the collector current is given by
I, = Le%ee\s, 3.5)
where Vpe is the base-emitter voltage. Vr = &T/g, and J, is a constant current,
which is proportional to the cross-sectional area of the emitter, which is used to
describe the transfer characteristic of the transistor in the forward active region.
In Fig, 3.4a, as in Fig. 3.3. if (W/L), = (WIL) and (W/L); = (W/L)s, equal
currents are forced through the two branches of the bias circuit and the voltage drop
across resistor R equals Ve. Thus the bias current is given by
Tas = 36)
Combining Eys. (3.5) and (3.6), we have
Vrs Ioias
Tics = ya. (3.7)
This equation can be solved iteratively for Iyias
Both supply-independent bias circuits have a second trivial steady -state condition,
cutoff when /pias = 0. To prevent the bias circuit from settling to the wrong steady-
state condition, a startup circuit is necessary in all practical applicanons. The circuit
to the right of the dashed line in Fig. 3.5 functions as a startup circuit. If nas = 0,
Qs is off and the voltage at node A is high, causing Qg to tun on and draw a current
through Qs, forcing the circuit to move to its other equilibrium state. Once the circuit
settles in the desired state, Qs tums on and node A goes low, turning off Qs. At
this state the startup circuit is essentially out of the picture.
Figure 3.5. Supply-independent bias circuit with startup.9.1, BIAS CIRCUITS INMOS TECHNOLOGY = 53
Another important performance aspect of the bias circuits is their temperature
dependence. Unfortunately, supply-independent bias circuits are not necessarily tem-
perature independent, because the base-emitter voltage (Vgr), and gate-to-source
voltage (Vcs) are both temperature dependent. If the temperature coefficient Ter is
defined as the relative change of the bias current per degree Celsius temperature
variation, we have [1]
— 1 Olbpias
Ter = ar
G.8)
Using the definition above and Eq. (3.6), the relative temperature coefficient of the
Vge-based bias generator can be calculated:
Vee L _ Vag OR) 1
Ter ot) log 39
1 aR
Ter = RF: (3.10)
Since the temperature coefficient of the base-emitter junction voltage is negative (— 2
mV/°C) while resistors typically have a positive temperature coefficient, the two terms
in Bq. (3.9) add, resulting in a net Te, that is quite high. The temperature behavior of
the threshold-based bias gencrator of Fig. 3.3 is similar to the Vpz-based circuit.
An alternative supply-independent bias generator is the A Vpg-based circuit shown
in Fig. 3.6a, The operation of this circuit is based on the difference between the
base-emitter voltages of two transistors operated at different current densitics. In
Fig. 3.6a, as in Figs. 3.3 and 3.4, (W/L), = (W/L), and (W/L); = (WIL). Therefore,
equal currents flow through the two branches of the circuit and Ves; = Ves. Also,
the pnp transistor M,, has an emitter area that is m times the emitter area of, Mo.
The voltage across the resistor R is AVpe = Varo — Veer. From Eq. (3.5),
Trias = Le Vee s,
Tring = mile Veer, .11)
and hence
AVee = Vee0 — Veer = Vrin(m). (3.12)
AVog appears across R and produces a current of value
= AYoe _ Vrinm
Tis R R (3.13)
Obviously, the resulting bias current is independent of the power supply Vpp. This
circuit also has two operating states: one at the desired operating current given by
Eq. (3.13) and the other at zero. To prevent the circuit from operating in the cutoff
state, a startup circuit similar to the one shown in Fig. 3.5 is required.54 BASIC ANALOG CMOS SUBCIRCUITS
Figure 3.6. (a) AVge-based supply independent bias
generator; (b) high-performance AVae-based supply-
independent bias generator.
‘The temperature coefficient of the bias current can be calculated from Eq. (3.13):
= (Vekm _ Vrinm aR) 1
ter = (GH R 7 RE at) ia @.14)
_ 1 av, _ 1 aR
“vp ar Ra’ G.15)
Ter3.2, MOS CURRENT MIRRORS AND CURRENT SOURCES 55,
Since V;/8T and AR/QT are both positive, the two terms in the temperature coeffi-
cients tend to cancel each other. Thus compared to Vge-or threshold-based bias
circuits, the A Vpg-based bias circuit can produce a much smaller temperature coeffi-
cient.
‘One drawback of the A Vpg-based bias generator is the strong dependence of Ibias
‘on the mismatches between Q, Q, and Q, Q, device pairs. The mismatch between
Q;-Qz will result in different currents to flow in the two branches of the circuit. If
1 + € represents the ratio of the two currents, A Vge will become
AVoe = Vr In (m(1 + €)], (3.16)
which is equivalent to modifying m, the ratio of the emitter areas, by | + €. The
mismatch between Q, and Q, and the current difference due to the mismatch of Qs
and Q, will make the Ves values of Qy and Qs different. This is equivalent to a de
offset voltage Vor = AVas. which modifies Eq. (3.13) to
(3.17)
Vr In [m(1_+ ©)] — AVas
Jog =
Assuming that m = 8, € = 0.01, and Vr = 26 mV at room temperature, A Vag =
26 X In[8(1 + 0.01)] = 54.3 mV results. For a AVcs = 5 mV offset voltage,
from Eq. (3.17), Ibies Will change by 10%. To reduce this variation special care
should be taken in the layout of Q:—Qs. For better geometrical matching, these
devices should use a common-centroid layout strategy [2].
‘The current-matching accuracy of the bias generator of Fig. 3.6a is further de-
graded due to the mismatch between the drain-to-source voltages of Q;-Q, and
Q1-Qz transistor pairs. The circuit can be made symmetrical, and the drain-to-source
voltage drops equalized, by adding transistors Qs to Qg to the two branches of the
circuit (Fig. 3.65). The improved configuration also uses the cascode current mirror
principle, described in Section 3.2, wo improve the power supply i¢jection. On the
other hand, the minimum power supply voltage is increased compared to the circuit
of Fig. 3.6a, due to the extra voltage drops required by the two cascode devices.
This becomes a major shortcoming in advanced submicron process technologies, ot
in low-power/low-voltage applications where the power supply voltage is limited
103.3 V.
3.2. MOS CURRENT MIRRORS AND CURRENT SOURCES
As will be seen in later sections, constant current sources and current mirrors are
important components in MOS amplifiers. The MOS current sources are quite similar
to the bipolar sources [1,3], where the current mirrors work on the principle that
identical devices with equal gate-to-source and drain-to-source voltages carry equal