Uploaded by linzhi xie

2007 EE214 Advanced Analog Inteqrated Circuits

advertisement
1
STANFORD UNIVERSITY
Department of Electrical Engineering
Prof. Boris Murmann
EE214: Analog Integrated Circuit Design
- Autumn 2007/08 http://eeclass.stanford.edu/ee214/
Table of Contents
Lecture 1
Lecture 2
Lecture 3
Lecture 4
Lecture 5
Lecture 6
Lecture 7
Lecture 8
Lecture 9
Lecture 10
Lecture 11
Lecture 12
Lecture 13
Lecture 14
Lecture 15
Lecture 16
Lecture 17
Lecture 18
Lecture 19
Lecture 20
Lecture 21
Lecture 22
Lecture 23
Lecture 24
Lecture 25
Lecture 26
Lecture 27
Lecture 28
Lecture 29
Introduction
CMOS Technology, Long Channel MOS Model
Common Source Amplifier
Technology Characterization: gm/ID
Technology Characterization: fT, gm/gds
gm/ID-based Design
Extrinsic Capacitance
Miller Approximation, ZV Time Constant Analysis
Electronic Noise
Electronic Noise (Continued)
Backgate Effect, Common Gate Stage
Common Drain Stage
Differential Pair
Current Mirrors, Offset Voltage
Process Variations, Feedback
Fully Differential Amplifiers, SC Circuits
Stability, Analysis of Feedback Circuits
Loop Gain Simulation
Two-Stage OTA
Compensation, Noise in Feedback OTAs
OTA Design Considerations
Step Response
Slewing
Feedback and Port Impedances, OTA Variants
Single Ended OTAs, Output Stage Examples
Supply Insensitive Biasing
Bandgap Reference
Bandgap Reference (Continued)
Technology Scaling
Class Summary
3
10
16
30
42
56
68
85
96
109
118
130
141
152
165
175
185
197
209
218
233
258
275
285
298
307
317
323
332
348
2
3
EE214
Analog Integrated Circuit Design
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Introduction
1
A Few Words About Your Instructor
•
Assistant Professor in EE since 2004
•
PhD, UC Berkeley 2003
– Digitally assisted A/D conversion
– Use "minimalistic" analog circuits (low power, fast)
– Correct errors using digital post-processor
•
~ 4 years work experience in IC industry
– Mixed signal IC design, low power, high voltage
•
Current research
– Digital correction techniques for data converters
– Sensor interfaces
– Circuit design in new technologies
• Post-CMOS devices, organic devices
B. Murmann
EE 214 Introduction
2
4
EE214 Basics (1)
•
Teaching assistants
– Mohammad Hekmat, Bob Wiser, Ross Walker
•
Administrative support
– Ann Guerra, CIS 207
•
Lectures are televised
– But please come to class to keep the discussion interactive!
•
Web page: http://eeclass.stanford.edu/ee214
– Check regularly, especially bulletin board
– Register for online access to grades and solutions
• Only enrolled students can register; we manually control the
access list based on Axess data
B. Murmann
EE 214 Introduction
3
EE214 Basics (2)
•
Required text
– Analysis and Design of Analog Integrated Circuits, 4th
Edition, Gray, Hurst, Lewis and Meyer, Wiley, 2001. (On
reserve in Engineering Library)
•
Course prerequisites
– EE101B or equivalent
– Basic device physics and models
• PN junctions, MOSFETs, BJTs
– Basic linear systems
• Frequency response, poles, zeros
– Some exposure to a circuit simulator, basic Unix commands
– May consider concurrent enrollment in EE114X to brush up
on the above (primarily for undergraduates)
B. Murmann
EE 214 Introduction
4
5
Assignments
•
Homework (20%)
– Handed out on Mondays, due following Monday in class
– Late policy
• Score drops 0.5 dB per hour after deadline
– Lowest HW score will be dropped
– Policy for off-campus students: Fax/email to SCPD before
deadline stated on handout
•
Midterm Exam (30%)
•
Project (20%)
– Design of an amplifier using HSpice (no layout)
– Work in teams of two
• OK to discuss with other teams, but no file exchange!
•
Final Exam (30%)
B. Murmann
EE 214 Introduction
5
Honor Code
•
Please remember you are bound by the honor code
– I will trust you not to cheat
– I will try not to tempt you
•
But if you are found cheating it is very serious
– There is a formal hearing
– You can be thrown out of Stanford
•
Save yourself and me a huge hassle and be honest
•
For more info
– http://www.stanford.edu/dept/vpsa/judicialaffairs/guiding/pdf/
honorcode.pdf
B. Murmann
EE 214 Introduction
6
6
Be Reasonable When Asking TAs
•
The TAs will not give you "the answer times two"…
•
They will also NOT debug your Spice deck
– Figuring out what's wrong with your circuit is an essential
component of this class
B. Murmann
EE 214 Introduction
7
Circuit Simulation
•
We will HSpice for circuit simulation
– You can use other tools at "own risk"
– "CAD Basics" document and example simulation files are provided
on course web site and in course directory
•
Plot HSpice results using Matlab ("HSpice Toolbox")
– Toolbox is installed in course directory
• See "CAD Basics" document for setup info
– Can download toolbox from Mike Perrott's homepage (MIT)
•
EE214 Technology
– 0.35μm CMOS
– BSIM3v3 models provided on web site and in course directory
•
First review session (this week) will focus on simulation basics
B. Murmann
EE 214 Introduction
8
7
The Spice Monkey Problem (1)
•
What most people know
– Even a very large number of
monkeys randomly
arranging characters will
never manage to write an
interesting book
•
What some people tend to
forget
– Even a very large number of
"Spice Monkeys" randomly
tweaking circuits will never
manage to design a robust,
optimized IC
B. Murmann
[Courtesy Isaac Martinez]
EE 214 Introduction
9
The Spice Monkey Problem (2)
•
Simply put
– Spice is nothing but a "calculator" that lets you evaluate and
test your ideas
– There is no need to simulate anything unless you already
know the (approximate) answer!
– Must always be aware of modeling limitations
•
Especially in the integrated circuits arena, uneducated, purely
simulator driven design can be costly
– Mask sets cost up to $2 Million (90 nm production)
– Turnaround time is on the order of months
– If your chip doesn't work, you cannot simply send the
customer a "patch"…
B. Murmann
EE 214 Introduction
10
8
Analysis versus Design
•
Unlike common perception, analog circuit analysis and design is
not "black magic"
•
Circuit analysis
– The art of decomposing a circuit into manageable pieces
– Based on the simple, but sufficiently accurate model
• "Just-in-time" modeling; do not use a complex model unless
you know why it's needed…
– One circuit ⇒ one solution
•
Circuit design
– The art of synthesizing circuits based on experience from
extensive analysis
– One set of specifications ⇒ Many solutions
– Design skills are best acquired through "learning by doing"
• This is why we'll have a design project…
B. Murmann
EE 214 Introduction
11
Learning Goals
•
Develop deeper understanding of MOS device behavior relevant
to analog design
•
Develop a feel for limits and tradeoffs in analog circuits (speed,
noise, power dissipation)
•
Learn to bridge the gap between complex device
models/behavior and basic hand calculations
– Design using look-up tables, "gm/ID methodology"
•
Develop a systematic, non-spice-monkey design style
•
Solidify the above aspects in a hands-on design project
– Design and optimization of a high performance feedback
amplifier used in many industrial circuits/applications
B. Murmann
EE 214 Introduction
12
9
Preview - Design Example of Lecture 20
Cf
CL
Cs
Vsd
+
+
Vod
-
Vid
Cs
CL
M4a,b
Cf
M1a,b
M2a,b
M3a,b
B. Murmann
Specs:
Loop bandwidth (fc) = 200MHz
Phase margin = 75 degrees
DR = 72dB
Closed-loop gain =2
Static gain error < 0.5%
EE 214 Introduction
13
Course Topics
•
CMOS technology and device models
•
Electronic noise
•
Single-stage amplifiers
•
Current mirrors, active loads
•
Differential pairs
•
Operational transconductance amplifiers (OTAs)
•
Feedback, stability and compensation
•
Temperature and supply independent biasing
B. Murmann
EE 214 Introduction
14
10
Lecture 1
CMOS Technology
Long Channel MOS Model
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 1
1
Overview
• Reading
– 2.8 (MOS fabrication), 2.9 (Active MOS devices)
– 2.10.1 (Resistors), 2.10.2 (Capacitors)
– 1.1, 1.5.0, 1.5.1, 1.5.2, 1.5.3 (Large signal MOS model)
• Introduction
– In this first lecture, we will cover some of the background
that positions EE214 as an introductory course on circuit
design using CMOS technology. In the lectures to come, we
will focus on the problem of amplifier design as a vehicle to
establish a set of considerations that apply to more complex
circuits and also other technologies. At first, we will review
the "long channel model" of a MOS transistor. Driven by
circuit examples, we will later augment this simple model to
include additional effects that are relevant in practice.
B. Murmann
EE 214 Lecture 1
2
11
The Big Picture
•
Most modern electronic information processing systems rely on
amplification of "small" physical signals
– E.g. signal from RF antenna, disk drive head, microphone, …
•
EE214 uses amplifiers as a vehicle to teach you the basics of
analog integrated circuit analysis and design
– Material forms basis for other and/or more complex circuits
B. Murmann
EE 214 Lecture 1
3
Technological Progress
Vacuum Tube
1906
Transistor
1947
Integrated Circuit
1958
B. Murmann
Modern Discrete
Transistors
Modern
CMOS
EE 214 Lecture 1
4
12
45nm CMOS (Intel)
Steve Cowden
THE ORGONIAN
July 2007
B. Murmann
EE 214 Lecture 1
5
Economics
[European
Nanotechnology
Roadmap]
B. Murmann
EE 214 Lecture 1
6
13
Future Applications
B. Murmann
EE 214 Lecture 1
7
Discrete vs. Integrated Circuits
Discrete Audio Amplifier
Integrated CMOS Audio Amplifier
•
Minimize transistor count
•
"Unlimited" number of transistors
•
Devices usually don't match
•
Devices match well
•
Arbitrary resistor values
•
Keep resistors < 10…100k
•
Capacitors 1pF…10mF
•
Keep capacitors < 10…50pF
B. Murmann
EE 214 Lecture 1
8
14
Modern Integrated Circuit Technologies
•
Parameter
CMOS
Si BJT
SiGe BJT
Device Speed
High
High
High
Noise
Poor
Good
Good
Transconductance
Poor
Good
Good
Intrinsic gain
Poor
Better
Best
Why use CMOS for analog integrated circuits?
– Low cost, driven by high volume digital ICs
– Integration with high density digital circuits
• BiCMOS tends to be expensive
B. Murmann
EE 214 Lecture 1
9
Basic MOS Operation (1)
0V
0V
VD (>0V)
0V
•
With zero voltage at the gate, device is "off"
– Back-to-back reverse biased pn junctions
B. Murmann
EE 214 Lecture 1
10
15
Basic MOS Operation (2)
>0
•
With a positive gate bias applied, electrons are pulled toward
the positive gate electrode
•
Given a large enough bias, the electrons start to "invert" the
surface (p→n); a conductive channel forms
– Magic "threshold voltage" Vt (more later)
B. Murmann
EE 214 Lecture 1
11
Basic Operation (3)
ID=?
>0
VDS>0
•
If we now apply a positive drain voltage, current will flow
•
How can we calculate this current as a function of VGS, VDS?
B. Murmann
EE 214 Lecture 1
12
16
Lecture 2
Common Source Amplifier
Small-Signal Model
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 2
1
Overview
•
Reading
– 3.0 (Amplifier basics), 3.1 (Model selection)
– 3.3.2 (Common source amplifer)
– 1.6.0 - 1.6.5 (Small signal MOS model)
•
Introduction
– Today we'll complete our derivation of the basic longchannel MOSFET I-V characteristics. As a next step, we'll
use this simple model to construct our first amplifier – a
common source stage. Looking at its transfer function, we'll
find that treating signals as "small" with respect to the bias
conditions allows us to linearize the circuit. Next, we
generalize this approach and develop a more universal
"plug-and-play" small-signal model for MOS devices that are
biased in the active region.
B. Murmann
EE 214 Lecture 2
2
17
Basic MOS Operation
ID=?
>0
VDS>0
•
How can we calculate ID as a function of VGS, VDS?
B. Murmann
EE 214 Lecture 2
3
Assumptions
>0
VDS>0
1) Current is controlled by the mobile charge in the channel. This is a very
good approximation.
2) "Gradual Channel Assumption" - The vertical field sets channel charge,
so we can approximate the available mobile charge through the voltage
difference between the gate and the channel
3) The last and worst assumption (we will fix it later) is that the carrier
velocity is proportional to lateral field (ν = μE). This is equivalent to Ohm's
law: velocity (current) is proportional to E-field (voltage)
B. Murmann
EE 214 Lecture 2
4
18
First Order IV Characteristics (1)
•
What we know:
Qn ( y ) = Cox [VGS − V ( y ) − Vt ]
I D = Qn ⋅ v ⋅ W
v = μ⋅E
∴ I D = Cox [VGS − V ( y ) − Vt ] ⋅ μ ⋅ E ⋅ W
B. Murmann
EE 214 Lecture 2
5
First Order IV Characteristics (2)
I D = Cox [VGS − V ( y ) − Vt ] ⋅ μ ⋅ E ⋅ W
E=
dV ( y )
dy
I D dy = WμCox [VGS − V ( y ) − Vt ] ⋅ dV
L
VDS
0
0
I D ∫ dy = WμCox ∫ [VGS − V ( y ) − Vt ] ⋅ dV
I D = μCox
W
L
⎡
VDS ⎤
(
)
V
V
−
−
⎢ GS
⎥ ⋅ VDS
t
2
⎢⎣
⎥⎦
•
For VDS/2 << VGS-Vt, this looks a lot like a linear resistor: I=1/R × V
•
Lets plot this IV relationship...
B. Murmann
EE 214 Lecture 2
6
19
ID
Plot of First Order IV Curves
VGS-Vt
VDS
•
Something is wrong here...
– Current should never decrease with increasing VDS
•
What happens when VDS>VGS-Vt?
– VGD = VGS-VDS becomes less than Vt, i.e. no more
channel or "pinch off"
B. Murmann
EE 214 Lecture 2
7
Pinch-Off
– VGS +
N
+ VDS –
Qn(y), V(y)
N
Voltage at the end of channel
Is fixed at VGS-Vt
y
y=0
y=L
•
Effective voltage across channel is VGS - Vt
– After channel charge goes to 0, there is a high lateral field
that ‘sweeps’ the carriers to the drain, and drops the extra
voltage (this is a depletion region of the drain junction)
•
To first order, current becomes independent of VDS
B. Murmann
EE 214 Lecture 2
8
20
Modified Plot and Equations
Triode
Region
ID
Active
Region
VGS-Vt
VDS
Triode Region:
I D = μCox
W
L
⎡
V ⎤
⎢(VGS − Vt ) − DS ⎥ ⋅ VDS
2 ⎥⎦
⎢⎣
Active Region:
I D = μCox
W
L
(VGS − Vt ) ⎤
1
W
⎡
⋅ (VGS − Vt ) = μCox (VGS − Vt ) 2
⎢⎣(VGS − Vt ) −
⎥
2
2
L
⎦
B. Murmann
EE 214 Lecture 2
9
First-Order MOS Model Summary
VDS
W
1
I D ≅ μCox (VGS − Vt )2
2
L
VGS-Vt
old
sh
e
hr
...)
-T
er
t
b
a
l
Su
re
o
(m
"VCCS"
ACTIVE
TRIODE
I D ≅ μCox
W
L
VDS ⎤
⎡
⎢⎣(VGS − Vt ) − 2 ⎥⎦ ⋅ VDS
Vt
B. Murmann
≅
≅
VGS
EE 214 Lecture 2
10
21
Model Accuracy
•
The above equations constitute the most basic MOS IV model
– "Long channel model", "quadratic model", "low field model"
•
Unfortunately this model doesn't describe modern CMOS
devices accurately
– Pushing towards extremely small geometries has resulted in
very high electric fields
• Some of the assumptions on slide 4 become invalid
• Other second order dependencies arise
•
Nevertheless, we will use this simple model in the first few
lectures to develop some basic circuit intuition
– Will fix and refine as we go…
– "Just-in-time" modeling
B. Murmann
EE 214 Lecture 2
11
Let's Build Our First Amplifier
•
One way to amplify
– Convert input voltage to current using voltage controlled
current source (VCCS)
– Convert back to voltage using a resistor (R)
•
"Voltage gain" = ΔVout/ΔVin
– Product of the V-I and I-V conversion factors
B. Murmann
EE 214 Lecture 2
12
22
Common Source Amplifier
•
MOS device acts as VCCS
W
1
I D = μCox (Vi − Vt )2
2
L
B. Murmann
W
1
Vo = VDD − μCox (Vi − Vt )2 ⋅ R
2
L
EE 214 Lecture 2
13
Biasing
•
Need some sort of "battery" that brings input voltage into useful
operating region
•
Define VOV=VI-Vt, "quiescent point gate overdrive"
– VOV=VGS-Vt with no input signal applied
VO
ΔVo
ΔVi
"Signal"
VOV
"Bias"
VI
B. Murmann
EE 214 Lecture 2
14
23
Relationship Between Incremental Voltages
•
What is ΔVo as a function of ΔVi?
1
W
VO + ΔVo = VDD − μCox (VOV + ΔVi )2 ⋅ R
2
L
1
W
ΔVo = − μCox R ⋅ (VOV + ΔVi )2 − VOV 2
L
2
1
W
= − μCox R ⋅ 2VOV ΔVi + ΔVi 2
2
L
⎡
2I
ΔVi ⎤
= − D ⋅ R ⋅ ΔVi ⎢1 +
⎥
VOV
⎣ 2VOV ⎦
[
[
]
]
•
As expected, this is a nonlinear relationship
•
Nobody likes nonlinear equations; we need a simpler model
– Fortunately, a linear approximation to the above expression
is sufficient for 90% of all analog circuit analysis
B. Murmann
EE 214 Lecture 2
15
Small Signal Approximation (1)
ΔVo = −
•
⎡
ΔVi ⎤
2I D
⋅ R ⋅ ΔVi ⎢1 +
⎥
VOV
⎣ 2VOV ⎦
Assuming ΔVi << 2VOV, we have
ΔVo ≅ −
•
2I D
⋅ R ⋅ ΔVi
VOV
If we further pretend that the input voltage increment is infinitely
small, we can find this result directly by taking the derivative of
the large signal transfer function at the "operating point" VI
dVo
2I
= − D ⋅R
dVi V =V
VOV
i
B. Murmann
I
EE 214 Lecture 2
16
24
Small Signal Approximation (2)
•
Graphical illustration:
dVo/dVi
VO
VOV
VI
•
The slope of the above tangent is the so called "small signal
gain" of our amplifier
B. Murmann
EE 214 Lecture 2
17
Small Signal MOS Model
•
Fortunately we don't have to repeat this analysis for every single
circuit we build
•
Instead, we derive a linearized circuit model for the MOS
transistor and plug it into arbitrary circuits
B. Murmann
EE 214 Lecture 2
18
25
Transconductance
•
The parameter that relates small signal gate voltage to drain
current is called transconductance (gm), or y21 in two-port
nomenclature
•
The transconductance is found by differentiating the large signal
I-V characteristic of the transistor in its operating point
W
1
I D = μCox (VGS − Vt )2
2
L
gm =
id
W
W
∂I
= D = μCox (VGS − Vt ) = μCox VOV
vgs ∂VGS
L
L
gm =
B. Murmann
2I D
VOV
EE 214 Lecture 2
19
Additional Model Components
•
Now that we've decided to move on using "small signal"
approximations, it also becomes easier to refine our model and
make it more realistic
•
Let's first take a look at "intrinsic gate capacitance"
– Intrinsic means that these capacitances are unavoidable and
required for the operation of the device
– Note that there are plenty of extrinsic, technology related
capacitances
• We'll talk about some of those later
•
When talking about gate capacitance, we must distinguish
several operating regions
– Transistor on
• Triode and active regions
– Transistor "off"
• Subthreshold operation
B. Murmann
EE 214 Lecture 2
20
26
Transistor in Triode Region
G
S
W
D
L
C GC
C CB
•
Gate terminal and conductive channel form a parallel plate
capacitor across gate oxide CGC= WLεox/tox= WLCox
– We can approximately model this using lumped capacitors of size
½ CGC each from gate-source and gate-drain
• Changing either voltage will change the channel charge
•
The depletion capacitance CCB adds extra capacitance from
drain and source to substrate
– Usually negligible
B. Murmann
EE 214 Lecture 2
21
Transistor in Active Region
•
Assuming a long channel model, if we change the the source
voltage in the forward active region
– The voltage difference between the gate and channel at the
drain end remains at Vt, but the voltage at the source end
changes
– This means that the "bottom plate" of the capacitor does not
change uniformly
•
Detailed analysis shows that in this case Cgs=2/3WLCox
– See text, section1.6.2
•
In the long channel model for forward active operation, the drain
voltage does not affect the channel charge
– This means Cgd=0 in the forward active region!
• Neglecting second order effects and extrinsic caps, of course
B. Murmann
EE 214 Lecture 2
22
27
Transistor Off
G
S
D
W
L
C GC
C CB
•
There is no conductive channel
– Gate sees a capacitor to substrate, equivalent to the series
combination of the gate oxide capacitor and the depletion
capacitance
•
If the gate voltage is taken negative, the depletion region
shrinks, and the gate-substrate capacitance grows
– With large negative bias, the capacitance approaches CGC
B. Murmann
EE 214 Lecture 2
23
Intrinsic MOS Capacitor Summary
Triode
Cgs
0
½ WLCox
Cgd
0
½ WLCox
0
Cgb
⎛ 1
1 ⎞
⎟⎟
⎜⎜
+
⎝ CCB WLCox ⎠
0
0
CCB =
B. Murmann
Forward
Active
Subthreshold
ε Si
xd
2/
3
WLCox
−1
WL
EE 214 Lecture 2
24
28
Finite dID/dVDS (1)
•
In the simple model considered so far, the drain current was
independent of VDS (active region)
•
In reality, the drain current has a weak dependence on VDS
Triode
Region
Active
Region
ID
Finite dID/dVDS
VGS-Vt
VDS
B. Murmann
EE 214 Lecture 2
25
Finite dID/dVDS (2)
•
"Channel length modulation" is outdated nomenclature for a
combination of several physical effects (DIBL, SCBE, …) that
cause finite dID/dVDS
•
The precise dependence of ID on VDS is very hard to model
– You can convince yourself by looking at the BSIM3 manual
•
The simplest and most popular model for hand analysis
assumes that the large signal current ID increases linearly with
VDS and lumps all dependencies into a single "fudge factor" λ
– λ is inversely proportional to channel length; i.e., longer
channels exhibit smaller dID/dVDS
ID =
B. Murmann
1
W
μCox ( VGS − Vt )2 ( 1 + λVDS )
2
L
EE 214 Lecture 2
26
29
Small Signal Output Conductance
•
From a small signal perspective, finite dID/dVDS translates into an
output conductance that depends on the operating point
ID
Slope = gds
Operating Point
VDS
dI D
d ⎡1
W
=
μCox ( VGS − Vt )2 ( 1 + λVDS
⎢
dVDS dVDS ⎣ 2
L
1
W
= μCox ( VGS − Vt )2 ⋅ λ
2
L
λI D
=
≅ λI D
1 + λVDS
g ds =
B. Murmann
⎤
)⎥
⎦
EE 214 Lecture 2
27
1st Order Small Signal Model (Active Region)
gm =
2I D
VOV
2
C gs = WLCox
3
g ds ≅ λ ⋅ I D
•
Sometimes ro=1/gds is used to denote finite output resistance
B. Murmann
EE 214 Lecture 2
28
30
Lecture 3
Common Source Amplifier Performance
Technology Characterization: gm/ID
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 3
1
Overview
•
Reading
– 1.6.8 (Transit Frequency)
– 1.8 (Weak Inversion)
•
Introduction
– Having established some basic modeling tools, we will now
begin to look at the performance of our common source stage:
bandwidth, power dissipation and maximum gain. We'll find that
these metrics are proportionally related to fundamental
performance measures of the MOS device: transit frequency
(gm/Cgs), current efficiency (gm/ID) and "intrinsic gain" (gm/gds) .
After looking at gm/ID in our EE214 0.35-μm technology, we find
that additional modeling is needed to explain the behavior of this
parameter as a function of the gate overdrive VOV. As a first
refinement, we discuss the behavior at subthreshold bias, i.e.
VOV<0.
B. Murmann
EE 214 Lecture 3
2
31
Common Source Amplifier Revisited
•
Consider a generic common source stage driven by a "transducer"
H( s ) =
B. Murmann
vo ( s )
1
= − gm R ⋅
vi ( s )
1 + sRi C gs
EE 214 Lecture 3
3
Performance Measures
•
DC voltage gain
¾ ADC specification and R set
required gm
•
Bandwidth
¾ Want small Cgs to maximize
bandwidth
•
Power dissipation
¾ Want small ID to minimize power
dissipation
B. Murmann
EE 214 Lecture 3
ADC = − g m R
f −3dB =
1
1
2π Ri C gs
P = VDD ⋅ I D
4
32
Device Perspective (1)
•
What we really want from our MOS transistor
– Some gm without investing much current (ID)
– Some gm without introducing large Cgs
•
To quantify how good of a job our transistor does, we can
therefore define the following "figures of merit"
gm
ID
•
gm
C gs
Using long channel MOS equations, we find
gm
2
=
I D VOV
•
and
and
g m 3 μVOV
=
C gs 2 L2
VOV is the "knob" that let's us trade power efficiency (gm/ID) for
speed (gm/Cgs)!
B. Murmann
EE 214 Lecture 3
5
Device Perspective (2)
gm/I D
gm/Cgs
VOV
•
Part of your job as a designer is to choose VOV such that
– You get sufficient bandwidth
– And use as little power as possible to accomplish this
•
Even though we've come to this graph using a very simple
example, the observed tradeoff tends to hold in general
– Of course, second order considerations will factor in as you
learn more about circuit design…
B. Murmann
EE 214 Lecture 3
6
33
Product
•
In cases where we want to get the "best of both worlds", it is
interesting to look at the product of our two figures of merit
g m g m 3μ
⋅
=
I D C gs L2
gm/I D
gm/I D*gm/Cgs
gm/Cgs
VOV
•
While this result looks boring, it shows that using smaller
channel lengths improves circuit performance
– Either or both speed and power efficiency
B. Murmann
EE 214 Lecture 3
7
Scaling Impact
•
Thanks to "Moore's Law" feature sizes and thus the available
minimum channel lengths have been shrinking continuously
– Lmin decreases roughly 2x every 5 years
– Lmin=10μm in 1970, Lmin=45nm in 2007
•
From the above discussion, it is clear that we can exploit
technology scaling in different ways
– Build faster circuits (gm/Cgs), while keeping power efficiency
constant (gm/ID)
• E.g. A/D converter for a disk drive - want to maximize
bandwidth/throughput
– Build more power efficient circuits (gm/ID), while keeping the
bandwidth constant (gm/Cgs)
• E.g. A/D converter for video signals - bandwidth fixed by a
certain standard
B. Murmann
EE 214 Lecture 3
8
34
Transit Frequency (ωT)
•
The transit frequency of a transistor has "historically" been
defined as the frequency where the magnitude of the common
source current gain (|io/ii|) falls to unity
•
Ignoring extrinsic capacitance, it follows that
ωT =
•
g m 3 μVOV
=
C gs 2 L2
Incidentally, this metric is identical to the figure of merit we
considered earlier in the context of a CS amplifier…
B. Murmann
EE 214 Lecture 3
9
Transit Frequency Interpretation
•
The transit frequency is only useful as a figure of merit in the
sense that it quantifies gm/Cgs
•
It does not accurately predict up to which frequency you can use
the device
– At high frequencies, many assumptions in our "lumped"
transistor model become invalid
– Rule of thumb: lumped model is good up to about ωT/5
•
At higher frequencies, device modeling becomes more
challenging and many effects depend on how exactly you layout
and connect the device
– These effects are covered in more detail in EE314
– In EE214, we will assume that we "care" only about
frequencies up to ωT/5
B. Murmann
EE 214 Lecture 3
10
35
+ ωmax
•
A step into the right direction for quantifying the high frequency
capability of a MOSFET is to look at its power gain with gate
sheet resistance effects included
– The quantity ωmax is defined as the frequency at which the
magnitude of the common source power gain falls to unity
– Also known as "maximum frequency of oscillation"
•
Can show that
ωmax =
ωT
1
2 rgateC gd
(more in EE314…)
B. Murmann
EE 214 Lecture 3
11
Intrinsic Gain
•
With RL→∞, the basic common source stage achieves its
maximum possible voltage gain or "intrinsic gain"
– This is yet another interesting figure of merit for a transistor
ADC = g m R = g m (RL || ro )
ADC ,max = g m ro =
≅
•
gm
g ds
2
1 gm
=
λ I D λVOV
Interestingly, it will turn out that the voltage gain of other, more
complicated circuits (e.g. op-amps) is fundamentally linked to
the intrinsic device gain gm/gds
B. Murmann
EE 214 Lecture 3
12
36
"Level 1" Figures of Merit for Transistors
Long Channel Model
2
•
Current Efficiency
gm
ID
•
Transit Frequency
gm
C gs
=
3 μ V OV
2 L2
•
Intrinsic Gain
gm
g ds
≅
2
λ V OV
=
V OV
•
Can characterize any technology (MOS, BJT, …) with respect to
these basic quantities
•
Big question
– Does the long channel model accurately describe these FOM?
B. Murmann
EE 214 Lecture 3
13
gm/ID Simulation
$ gm/id vs. gate overdrive
.param gs=1
vds
vgs
mn1
d 0
g 0
d g 0 0
dc 1.5V
dc 'gs'
nch214 L=0.35um W=10um
.op
.dc gs 0.4V 1.2V 10mV
.probe ov
= par('gs-vth(mn1)')
.probe gm_id = par('gmo(mn1)/i(mn1)')
.options post brief
.lib './ee214_hspice.txt' nominal
.end
B. Murmann
EE 214 Lecture 3
14
37
Result
40
35
EE214 technology
2/VOV
gm/I D [S/A]
30
BJT (q/kT)
25
20
15
10
5
0
-0.2
-0.1
0
0.1
0.2
V
B. Murmann
OV
0.3
0.4
0.5
[V]
EE 214 Lecture 3
15
Observations
•
Our long channel predication is fairly close for VOV > 150mV
•
Unfortunately gm/ID does not approach infinity for VOV → 0
•
It also seems that we cannot do better than a BJT, even though
the long channel equation would predict that for 0 < VOV < 2kT/q
≅ 52mV at room temperature
•
For further analysis, it helps to identify three distinct operating
regions
– Strong inversion: VOV > 150mV
• Deviations due to short channel effects
– Subthreshold: VOV < 0
• Behavior similar to a BJT, gm/ID nearly constant
– Moderate Inversion: 0 < VOV < 150mV
• Transition region, an interesting mix of the above
B. Murmann
EE 214 Lecture 3
16
38
Subthreshold Operation
A plot of the device current in our previous simulation:
1
10
0
0.8
10
-1
0.6
10
-2
10
-3
10
-4
I D [mA]
I D [mA]
•
0.4
0.2
0
-0.5
-5
0
0.5
1
10
-0.5
VOV [V]
•
0
0.5
1
VOV [V]
Questions:
– What determines the current when VOV< 0, i.e. VGS< Vt?
– What is the definition of Vt?
B. Murmann
EE 214 Lecture 3
17
Definition of Vt
•
Vt is defined as the VGS at which the number of electrons pulled
to the surface equals the number of doping atoms
•
Seems somewhat arbitrary, but makes sense in terms of surface
charge control
B. Murmann
EE 214 Lecture 3
18
39
Mobile Charge versus VOV
6.E-07
5.E-07
Fixed Charge
Mobile Charge
Total Charge
Charge [C]
4.E-07
3.E-07
2.E-07
1.E-07
0.E+00
-1.0
-0.5
0.0
0.5
1.0
VOV [V]
•
Around Vt (VOV=0), the relationship between mobile charge in the
channel and gate voltage becomes linear (Qn ~ Cox[VGS-Vt])
– Exactly what we assumed to derive the long channel model
B. Murmann
EE 214 Lecture 3
19
Mobile Charge on a Log Scale
•
On a log scale, we see that there are mobile charges before we
reach the threshold voltage
– Fundamental result of solid-state physics, not short channels
1.E-06
1.E-07
Mobile Charge [C]
1.E-08
1.E-09
1.E-10
1.E-11
1.E-12
1.E-13
1.E-14
1.E-15
1.E-16
-1.00
-0.50
0.00
0.50
1.00
VOV [V]
B. Murmann
EE 214 Lecture 3
20
40
BJT Similarity
•
We have
– An NPN sandwich, mobile minority carriers in the P region
•
This is a BJT!
– Except that the base potential is here controlled through a
capacitive divider, and not directly an electrode
B. Murmann
EE 214 Lecture 3
21
Subthreshold Current
•
We know that for a BJT
I C ≅ I S ⋅ eVBE /( kT / q )
•
In our case we have
I D ≅ I 0 ⋅ e(VGS −Vt ) /( nkT / q )
•
n is given by the capacitive divider
n=
C js + Cox
Cox
= 1+
C js
Cox
where Cjs is the depletion layer capacitance
•
In our technology n ≅ 1.5
B. Murmann
EE 214 Lecture 3
22
41
Subthreshold Transconductance
gm =
•
dI D 1 I D ⋅ q
=
dVGS n kT
g m dI D 1 q
=
=
I D dVGS n kT
Similar to BJT, but unfortunately n (≅1.5) times lower
40
35
gm/I D [S/A]
EE214 technology
2/VOV
~1.5x
30
BJT (q/kT)
25
20
15
10
5
0
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
VOV [V]
B. Murmann
EE 214 Lecture 3
23
42
Lecture 4
Short Channel Effects
Technology Characterization: fT, gm/gds
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 4
1
Overview
•
Reading
– 1.7 (Short Channel Effects)
•
Introduction
– Today, we continue our discussion on gm/ID modeling in a
MOS device. We explain the remaining discrepancies with
the long channel model and then move on to an examination
of gm/Cgs (fT) and gm/gds. In conclusion, we find that the long
channel model cannot accurately predict either performance
metric we care about (gm/ID, gm/Cgs and gm/gds). As a solution
to this problem, we will explore a chart-based design
methodology in the remainder of this course.
B. Murmann
EE 214 Lecture 4
2
43
Re-cap
40
35
EE214 technology
2/VOV
gm/I D [S/A]
30
25
BJT (q/kT)
Subthreshold
Operation
20
?
15
10
?
5
0
-0.2
-0.1
0
0.1
0.2
V
B. Murmann
OV
0.3
0.4
0.5
[V]
EE 214 Lecture 4
3
Moderate Inversion
•
In the transition region between subthreshold and strong
inversion, we have two different current mechanisms
Drift (MOS) - ν = μ E
Diffusion (BJT) - ν = D
dn kT dn
=
μ
dx
q dx
•
Both current components are always present
– Neither one clearly dominates around Vt
•
Can show that ratio of drift/diffusion current ~(VGS-Vt)/(kT/q)
– MOS equation becomes dominant at several kT/q
•
One way to close the gap between the two regimes is to work
with a mathematical fit
– Sometimes useful for computer optimization, not so great for
hand analysis…
B. Murmann
EE 214 Lecture 4
4
44
A Curve Fitting Attempt
⎧1 q
; VOV → 0
⎪
gm 2 q
1
⎪ m kT
=
≅⎨
2
I D m kT
⎪ 2
⎛ VOV ⋅ q ⎞
1+ 1+ ⎜
⎟
⎪V ; VOV → ∞
⎝ mkT ⎠
⎩ OV
40
EE214 technology
2/V
35
OV
Fitted Equation (m=1.7)
g /I
m D
[1/V]
30
25
20
15
10
5
0
0
0.1
0.2
0.3
V
OV
B. Murmann
0.4
0.5
[V]
EE 214 Lecture 4
5
A Closer Look at Strong Inversion
15
EE214 technology
2/V
g /I [1/V]
m D
OV
10
5
0
•
0.15
0.2
0.25
0.3
0.35
VOV [V]
0.4
0.45
0.5
Long channel model overestimates gm/ID by roughly 10…20%
– Something worth looking into…
B. Murmann
EE 214 Lecture 4
6
45
Short Channel Effects
•
Velocity saturation due to high lateral field
•
Mobility degradation due to high vertical field
•
Vt dependence on channel length and width
•
ro = f(VDS)
•
…
•
We will limit the discussion in EE214 to the first two aspects of
the above list
– Focus on qualitative understanding, since we will not factor
these effects into our hand calculations
B. Murmann
EE 214 Lecture 4
7
Velocity Saturation (1)
•
In lecture 2, we assumed that the carrier velocity is proportional
to the lateral E-field, v=μE
•
Unfortunately, the speed of carriers in silicon is limited
– At very high fields (high voltage drop across the conductive
channel), the carrier velocity saturates
ν=
Approximation:
μE
E
1+
Ec
E>>Ec ⇒ v=vscl=μEc
E=Ec ⇒ v=vscl/2
B. Murmann
EE 214 Lecture 4
8
46
Velocity Saturation (2)
•
It is important to distinguish various regions in the above plot
– Low field, the long channel equations still hold
– Moderate field, the long channel equations become
somewhat inaccurate
– Very high field across the conducting channel – the velocity
saturates completely and becomes essentially constant (vscl)
•
To get some feel for latter two cases, let's first estimate the E
field using simple long channel physics
•
In the forward active region, at pinch-off, the lateral field across
the channel is
E=
VOV
200mV
V
e.g.
= 0.57 ⋅106
m
L
0.35μm
B. Murmann
EE 214 Lecture 4
9
Field Estimates
•
In our 0.35μm technology, we have for an NMOS device
Ec =
vscl
μ
=
m
s = 6.2 ⋅106 V
m
m2
1.73 ⋅ 10 5
0.028
Vs
•
The above example shows that an 0.35μm NMOS device at
VOV=200mV does not operate anywhere near the critical field
(E=0.57·106 << Ec)
•
How about, e.g., 0.13μm?
E=
•
V
200mV
= 1.5 ⋅106
0.13μm
m
Still not too bad…
B. Murmann
EE 214 Lecture 4
10
47
Short Channel Equation
•
Bottom line is that most existing and future MOS analog circuits
are impaired, but not completely limited by velocity saturation
– The digital folks will tell you a different story
• Why?
•
A simple equation that captures the moderate deviation from the
long channel forward active drain current is (see text)
1
1
W 2
I D ≅ μCox VOV
⋅
2
L
⎛ VOV ⎞
⎜⎜1 +
⎟⎟
⎝ Ec L ⎠
E L ⋅ VOV
1
W
≅ μCox VOV ⋅ c
(Ec L + VOV )
2
L
"Parallel Combination"
B. Murmann
EE 214 Lecture 4
11
Typical Values for EcL
NMOS
PMOS
0.35μm
2V
8V
0.13μm
0.6V
2.4V
•
As long as VOV is "much less" than these voltages, the above
simplified equation holds with reasonable accuracy
•
We can use these numbers to check our earlier simulation data
for gm/ID. With the correction factor, we have
gm
2
1
≅
⋅
I D VOV ⎛ VOV ⎞
⎜⎜1 +
⎟⎟
⎝ Ec L ⎠
e.g .
2
1
2
⋅
≅
⋅ 0.9
VOV ⎛ 0.2 ⎞ VOV
⎜1 +
⎟
2 ⎠
⎝
(Note that this expression is found by using 1.223 and 1.233 in the text,
and not from the approximate ID expression on the previous slide)
•
Reasonable agreement with simulation data on slide 6
B. Murmann
EE 214 Lecture 4
12
48
Mobility Degradation due to Vertical Field
•
In short channel MOSFETS, the oxide thickness has been
continuously scaled down with feature sizes
– 6.5nm in 0.35μm, 2.2nm in 0.13μm technology
•
As a result, there is a large vertical electric field that tries to pull
the carriers closer to the "dirty" silicon surface
– Imperfections impede movement and thus mobility
•
This effect can be included by replacing the mobility term with
an "effective mobility"
μeff ≅
•
μ
(1 + θVOV )
θ = 0.1...0.4
1
V
Yet another "fudge factor"
– Possible to lump with EcL parameter
B. Murmann
EE 214 Lecture 4
13
Summary – gm/ID
•
The long channel model does not predict gm/ID with reasonable
accuracy in any operating regime
– Accuracy also tends to get worse in newer technology
•
Once again, we'll find a way to deal with this in practice
•
Simple trick: Change of design variables
– Instead of "thinking", in terms of VOV, we will use gm/ID as a
design variable, and not as an unknown that is determined
from our choice of VOV (or other long channel model
parameters)
• "gm/ID design methodology" - more later…
B. Murmann
EE 214 Lecture 4
14
49
fT Simulation
* ft versus gate overdrive
.param gs=1
vds
vgs
mn1
d 0
g 0
d g 0 0
dc 1.5V
dc 'gs'
nch214 L=0.35um W=10um
.op
.dc gs 0.4V 1.2V 10mV
.probe ov
.probe ft
= par('gs-vth(mn1)')
= par('1/2/3.142*gmo(mn1)/(-cgsbo(mn1))')
.options post brief dccap
* Note: "dccap" forces HSpice to recalculate caps in
* each simulation step (instead of using constant .op
* value). See HSpice manual for additional info.
.lib './ee214_hspice.txt' nominal
.end
B. Murmann
EE 214 Lecture 4
15
Result
NMOS W/L=10/0.35
30
EE214 technology
Long Channel Fit
25
f [GHz]
T
20
15
10
5
0
-0.2
-0.1
0
0.1
0.2
VOV [V]
Long channel model:
B. Murmann
fT =
EE 214 Lecture 4
0.3
0.4
0.5
1 3 μVOV
2π 2 L2
16
50
Observations - fT
•
Again, a simple long channel model doesn't do a very good job
– Large fT discrepancy in subthreshold operation and in strong
inversion (large VOV)
•
The reasons for these discrepancies are exactly the same as
the ones we came across when looking at gm/ID
– Bipolar action in subthreshold operation and moderate
inversion
– Short channel effects at large VOV
• Less gm, hence lower gm/Cgs
•
Same conclusion, we won't be able to make good predictions
with a simple long channel relationship
B. Murmann
EE 214 Lecture 4
17
gm/ID· fT
NMOS W/L=10/0.35
160
EE214 technology
Long Channel
gm/I D*f T [GHz/V]
140
120
100
80
Sweet spot (?)
60
40
20
0
-0.1
Short channel
effects
Long channel predicts
too much gm/ID
0
0.1
0.2
VOV [V]
Long Channel:
B. Murmann
0.3
0.4
0.5
gm
1 3μ
⋅ fT =
ID
2π L2
EE 214 Lecture 4
18
51
Intrinsic Gain Simulation (VOV)
* gm/gds versus gate overdrive
.param gs=1.5
mn1
d
g 0
0 nch214
vg
g
0
dc 'gs'
vd
d
0
dc 1.5
L=0.35um
W=10um
1.5V
.op
.dc gs 0 1.5
.probe ov1
.probe av1
10m
=
=
par('gs-vth(mn1)')
par('gmo(mn1)/gdso(mn1)')
.options post brief
.lib './ee214_hspice.txt' nominal
.end
B. Murmann
EE 214 Lecture 4
19
Result
NMOS, W/L=10/0.35, V
100
=1.5V
DS
gm/gds
80
60
40
20
0
-0.2
EE214 Technology
Long Channel Model, λ=0.1
0
0.2
0.4
V
•
OV
0.6
0.8
[V]
Impossible to approximate with long channel model equation!
B. Murmann
EE 214 Lecture 4
20
52
Using a Longer Channel
NMOS, W/L=10/0.7, VDS=1.5V
1500
EE214 Technology
Long Channel Model, λ=0.01
gm/gds
1000
500
0
-0.2
0
0.2
0.4
0.6
0.8
VOV [V]
•
Curve is also closer to long channel model, but still far off for small VOV…
•
Also, as expected, gm/gds is larger for a device with longer channel length
B. Murmann
EE 214 Lecture 4
21
Dependence on VDS
•
The long channel model predicts that gds and gm/gds are independent
of VDS
– As long as device is biased in active region
•
This is also no longer true in modern devices
– gds (and therefore gm/gds) shows a significant dependence on VDS
ID
Slope = gds1
OP2
Slope = gds2
OP1
VDS
B. Murmann
EE 214 Lecture 4
22
53
Intrinsic Gain Simulation (VDS)
* gm/gds versus vds
.param vt1=571.5m
mn1
d
g 0
vg
vd
g
d
0
0
.op
.dc vd 0 3
.probe gm1
.probe gds1
0 nch214
L=0.35um
W=10um
dc 'vt1+0.2'
dc 1.5
10m
=
=
par('gmo(mn1)')
par('gdso(mn1)')
.options post brief
.lib './ee214_hspice.txt' nominal
.end
B. Murmann
EE 214 Lecture 4
23
Result
NMOS, W/L=10/0.35, VOV=200mV
80
1.5
gm [S]
70
60
-3
1
0.5
50
0
0
1
30
6
20
4
10
0
0
B. Murmann
2
3
2
3
VDS [V]
40
1/gds [Ω]
gm/gds
x 10
0.5
1
1.5
VDS [V]
2
2.5
3
EE 214 Lecture 4
x 10
4
2
0
0
1
VDS [V]
24
54
Gradual Onset
NMOS, W/L=10/0.35, VOV=200mV
80
1.5
60
Triode
gm [S]
70
"Active"
-3
1
0.5
50
gm/gds
x 10
0
0
40
6
1/gds [Ω]
30
20
10
0
0
0.2
B. Murmann
0.4
0.6
VDS [V]
0.8
1
x 10
0.2
0.4
0.6
VDS [V]
0.8
1
0.2
0.4
0.6
VDS [V]
0.8
1
4
4
2
0
0
EE 214 Lecture 4
25
Observations – Intrinsic Gain
•
gm/gds shows a strong dependence on VDS bias
– Mostly due to varying gds
•
There is a gradual transition from triode to active
– Long channel model would have predicted an abrupt change
to large intrinsic gain at VDS = VOV
– Typically need VDS > VOV + 4kT/q to ensure at least
moderate intrinsic gain
•
At high VDS, gds increases due to SCBE (substrate current
induced body-effect); this causes a decrease in gm/gds
– Highly technology dependent, and usually not present in
PMOS devices
– If you are interested in more details, please refer to EE316
or a similar course
B. Murmann
EE 214 Lecture 4
26
55
+ gds nonlinearity
•
Is vds really a "small signal"?
•
The small signal approximation for gds becomes somewhat
inappropriate when the vds swing spans a large fraction of a
nonlinear ID-VDS characteristic
•
Luckily, in most practical situations, other (well understood)
sources of nonlinearity dominate (e.g. transconductance)
ID
OP
OP1
VDS
B. Murmann
EE 214 Lecture 4
27
Why care about the Long Channel Model?
•
By now, it should be clear that the long channel model does not
accurately predict the performance of a modern MOS device
– There is no simple expression that accurately links gm/ID, fT and
gm/gds to "long channel design parameters" such as VOV
– VOV also doesn't predict the onset of active operation ("Vdsat") all
that well
•
In EE214, we will use the long channel model only to understand trends
and proportionalities
– For design and optimization, we'll need a more accurate approach
•
Key idea
– The primary variables we care about from a performance
perspective are gm/ID, fT and gm/gds
– So why not work directly with these variables?
• Using Spice-generated design charts and/or look-up tables
•
We'll look at this idea using a few design examples
B. Murmann
EE 214 Lecture 4
28
56
Lecture 5
gm/ID-Based Design
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 5
1
Overview
•
Introduction
– In the past two lectures, we have learned that the long
channel model does not accurately predict the performance
of modern MOS devices. Hence, we switch toward a
strategy in which circuit-oriented performance metrics (such
as gm/ID) are used directly for design and optimization. For a
chosen operating point (gm/ID), other relevant parameters
(such as the device width) are determined using Spicegenerated design charts that serve as a replacement for
(inaccurate) model equations.
B. Murmann
EE 214 Lecture 5
2
57
Overview
•
References
– F. Silveira et. al. "A gm/ID based methodology for the design
of CMOS analog circuits and its application to the synthesis
of a silicon-on-insulator micropower OTA," IEEE Journal of
Solid-State Circuits, Sept. 1996, pp. 1314-1319.
– D. Foty, M. Bucher, D. Binkley, "Re-interpreting the MOS
transistor via the inversion coefficient and the continuum of
gms/Id," Proc. Int. Conf. on Electronics, Circuits and Systems,
pp. 1179-1182, Sept. 2002.
– Denis Flandre's Notes: "Méthodologie gm/ID: un chaînon
entre l'analyse symbolique et la synthèse de circuits
analogiques basse puissance," available at
http://www.comelec.enst.fr/taisa/Presentations/DenisFlandre.pps
– B. E. Boser, "Analog Circuit Design with Submicron
Transistors," IEEE SSCS Meeting, Santa Clara Valley, May
19, 2005, http://www.ewh.ieee.org/r6/scv/ssc/May1905.htm
B. Murmann
EE 214 Lecture 5
3
Design Example 1
3V
IB
Vo
RL
vi
1.5V
VB
•
CL
Given specifications
– DC gain=-2, IB ≤ 2mA, f-3dB=100MHz, CL=10pF
– Minimize transistor area (L=Lmin, W as small as possible)
B. Murmann
EE 214 Lecture 5
4
58
Does ro Matter?
ADC = g m (RL || ro )
1
ADC
⎛ 1
1⎞
= g m ⎜⎜
+ ⎟⎟
⎝ RL ro ⎠
1
1
=
+
g m RL g m ro
−1
1
1
1
=
+
2 g m RL g m ro
•
Even at L=Lmin= 0.35μm, we have gmro > 50 (see slide 23 of
lecture 4 )
•
ro will be negligible in this design problem
B. Murmann
EE 214 Lecture 5
5
Hand Calculations (1)
f −3dB =
1
1
2π RLC L
ADC ≅ − g m RL = −2
⇒ RL =
1
1
= 159Ω
2π 100 MHz ⋅ 10 pF
⇒ gm =
2
= 12.6 mS
159Ω
•
Using all the available current, we have
•
How about using less current?
– Using less current means that we'll
need a device with larger W
– But specifications asked to minimize W
B. Murmann
EE 214 Lecture 5
1
g m 12.6mS
=
= 6.3
2mA
ID
V
g m ~ 2 I D μCox
fixed
↓↓
W
L
↑↑
6
59
Hand Calculations (2)
•
To complete the design, we need to find the actual device width
•
As we know, using long channel equations will be very
inaccurate
•
This is where the idea of chart-based design comes in
•
Current density chart
– Plot of current density ID/W as a function of gm/ID
– Can generate this chart once and use it throughout the
design process
B. Murmann
EE 214 Lecture 5
7
Current Density Chart
NMOS L=0.35um (VDS=1.5V)
90
80
ID /W [μA/μm]
70
60
50
40
30
20
10
0
0
5
10
15
20
25
gm /ID [S/A]
B. Murmann
EE 214 Lecture 5
8
60
A Better Current Density Chart
=1.5V)
=1.5V)
NMOS L=0.35um(V(V
DSDS
I /W [μA/μm]
D
23μA/μm
1
10
0
10
3
4
5
6
7
8
6.3 1/V
B. Murmann
9
10 11 12 13 14 15 16 17 18 19 20
gm /ID [S/A]
EE 214 Lecture 5
9
Spice Verification
•
So, the device width is
•
How to set VB?
– E.g. using a replica
device (MN2)
I
W= D
ID
W
=
2000
μm = 87 μm
23
3V
IB
IB
VB
MN2
Vo
vi
MN1
RL
CL
1.5V
B. Murmann
EE 214 Lecture 5
10
61
DC Operating Point
*** mosfets
element
model
beta
0:mn1
0:mn2
0:nch214
0:nch214
region
Saturati
gam eff
gm
Saturati
60.5903m
60.6023m
894.1238m
894.1238m
13.0740m
12.5695m
236.3211u
280.4581u
id
2.1548m
2.0000m
gds
ibs
0.
0.
gmb
2.8628m
2.7961m
ibd
0.
0.
cdtot
116.2472f
130.8667f
859.4584m
cgtot
154.7093f
154.7008f
859.4584m
cstot
295.0430f
295.3025f
cbtot
304.1593f
318.7492f
cgs
107.8492f
108.1404f
cgd
19.7902f
19.7903f
vgs
859.4584m
vds
1.4754
vbs
0.
0.
vth
573.1479m
583.4011m
vdsat
206.8747m
201.7080m
vod
286.3105m
276.0573m
g m 13.074 mS
1
=
= 6.1
I D 2.1548 mA
V
B. Murmann
EE 214 Lecture 5
11
AC Response
7
6
6.03dB = 2.003
|vo/vi| [dB]
5
4
3
2
1
0 0
10
10
1
10
2
f [MHz]
B. Murmann
EE 214 Lecture 5
12
62
Does VDS matter?
NMOS L=0.35um
VDS=0.5V
VDS=1.5V
I D/W [uA/um]
VDS=2.5V
10
At gm/ID = 6.3 1/V:
VDS=2.5V
VDS=0.5V
1
ID/W = 21.8 A/m (VDS=0.5V)
ID/W = 23.3 A/m (VDS=1.5V)
ID/W = 23.6 A/m (VDS=2.5V)
∴ Insignificant dependence
on VDS; OK to use a single
chart for design (e.g.
VDS=1.5V)
10
0
3
4
B. Murmann
5
6
7
8
9
10
11 12 13 14
gm/I D [1/V]
15
16 17
18
19
20
EE 214 Lecture 5
13
Observations and Remarks (1)
•
The design is essentially "right on" target!
– No need for any Spice tweaking
•
We accomplished this by focusing on a performance related
parameter (gm/ID) in the design process
– Note that strictly speaking, device width is not a design
parameter, since it does not directly relate to any of the
electrical specs (gain, bandwidth, current) that we were
given
B. Murmann
EE 214 Lecture 5
14
63
Observations and Remarks (2)
•
The key advantage of gm/ID based design is that it allows you to
transition from hand analysis to Spice without much of the
"usual" modeling uncertainties
– Simply because we are incorporating relevant simulation
data into the design process
– Enables you to optimize your circuit without even running a
Spice simulation
•
To see why this is good, let's compare with some popular
alternatives, as seen in many labs and cubicles around the
country…
B. Murmann
EE 214 Lecture 5
15
Design Methodology 1 (Worst)
•
Have an existing design that somehow works for a different
process, different specs, …
•
Port this design over and tweak all 137 transistor geometries
until I meet the specs…
•
500 Spice runs later, I am approaching the deadline with a
design that somehow works, god knows how/why
B. Murmann
EE 214 Lecture 5
16
64
Design Methodology 2 (Better)
•
Remember the square law transistor model from class
•
Go through the pain of estimating μCox, and do some hand
calculations
•
Plug my design into Spice and realize that everything is about
20…80% off
•
Throw away my hand analysis and revert back to the "Spice
Monkey" design flow from here
B. Murmann
EE 214 Lecture 5
17
Design Methodology 3 (Much Better)
•
Spend some time to characterize your technology using Spice
– E.g. intrinsic gain and current density as a function of gm/ID
•
Do hand calculations using the generated technology data
– Use Matlab, MathCAD or Excel script
– Quickly iterate through tens of different designs, if necessary
•
Implement and verify in Spice
– Only minor tweaking necessary (if any)
– Done!
B. Murmann
EE 214 Lecture 5
18
65
Intrinsic Gain, NMOS
EE214 technology, NMOS, 0.35...0.7um
10
2
gm/gds
10
3
5
10
15
20
gm/I D [S/A]
B. Murmann
EE 214 Lecture 5
19
Intrinsic Gain, PMOS
EE214 technology, PMOS, 0.35...0.7um
2
10
1
gm/gds
10
5
10
15
20
gm/I D [S/A]
B. Murmann
EE 214 Lecture 5
20
66
Current Density, NMOS
EE214 technology, NMOS, 0.35...0.7um
1
10
0
I D/W [A/m]
10
5
10
15
20
gm/I D [S/A]
B. Murmann
EE 214 Lecture 5
21
Current Density, PMOS
I D/W [A/m]
EE214 technology, PMOS, 0.35...0.7um
10
1
10
0
10
-1
5
10
15
20
gm/I D [S/A]
B. Murmann
EE 214 Lecture 5
22
67
+ A Note on Current Density Charts
•
Designing with current density charts in a normalized, widthindependent space works because
– Current density and gm/ID are independent of W
• ID/W ~ W/W
• gm/ID ~ W/W
– There is a one-to-one mapping from gm/ID to current density
Long channel:
gm
2
=
I D VOV
General case:
gm
= f (VOV )
ID
B. Murmann
ID 1
1 2
1⎛ I ⎞
= μCox VOV
= μCox ⎜⎜ 2 D ⎟⎟
W 2
L
L ⎝ gm ⎠
⎛
⎛ g ⎞⎞
ID
= g (VOV ) = g ⎜⎜ f −1 ⎜⎜ m ⎟⎟ ⎟⎟
W
⎝ ID ⎠⎠
⎝
EE 214 Lecture 5
−2
23
68
Lecture 6
Extrinsic Capacitance
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 6
1
Overview
•
Reading
– 1.6.7 (Parasitic Elements)
•
Introduction
– In today's lecture, we'll look at another CS amplifier design
example – this time with an input source that has a relatively
large resistance. Through this example, we find that we need
more modeling to accurately predict the resulting pole at the
gate node. Our discussion leads to a discussion of parasitic
extrinsic capacitors around the MOSFET - overlap and
junction capacitance.
B. Murmann
EE 214 Lecture 6
2
69
Design Example 2
VDD
IB
Transducer
Vo
vi
Ri
RL
Ri
vi
+
vgs
-
Cgs gmvgs
ro
+
RL vo
-
1.5V
v (s)
1
H( s ) = o
= − g m ( ro || RL ) ⋅
vi ( s )
1 + sRi C gs
VB
•
Given specifications
– DC gain=-4, IB ≤ 0.5mA
– RL=1k, Ri=10k
– Maximize and estimate bandwidth
B. Murmann
DC gain
Frequency
Dependence
EE 214 Lecture 6
3
Hand Calculation
•
Just as in the previous design example, we know that
gm/gds >> |ADC|. Hence we simply find
ADC ≅ − g m RL = −4
gm =
•
4
= 4 mS
1kΩ
In order to maximize bandwidth, we need to make Cgs (and
hence W) as small as possible. Again, this is the case for using
up all the available current
4mS
1
gm
=
=8
I D 0.5mA
V
•
In order to estimate the circuit's bandwidth, we need to know Cgs
– Solution: transit frequency chart
B. Murmann
EE 214 Lecture 6
4
70
Transit Frequency Chart
NMOS L=0.35um
30
25
f T [GHz]
20
16GHz
15
10
5
0
0
5
10
15
20
25
gm/I D [1/V]
B. Murmann
EE 214 Lecture 6
5
Bandwidth
•
Using the transit frequency chart, we find
C gs =
f −3dB =
•
1 gm
1 4mS
=
= 40 fF
2π fT 2π 16GHz
1 1
1
1
=
= 398MHz
2π Ri C gs 2π 10k ⋅ 40 fF
As a last step, we use the current density chart to find the device
width
B. Murmann
EE 214 Lecture 6
6
71
Current Density Chart
NMOS L=0.35um
I D/W [uA/um]
15.5μA/μm
10
1
10
0
3
4
5
6
7
B. Murmann
8
9
10 11 12 13 14 15 16 17 18 19 20
gm/I D [1/V]
EE 214 Lecture 6
7
Spice Verification
•
Device width
•
Simulation circuit
B. Murmann
I
W= D
EE 214 Lecture 6
ID
W
=
500
= 32μm
15.5
8
72
DC Operating Point
**** mosfets
element
model
region
id
ibs
0:mn2
gam eff
Saturati
500.0000u
549.5104u
0.
vgs
806.0164m
vds
806.0164m
vbs
0:nch214
Saturati
0.
ibd
beta
0:mn1
0:nch214
0.
0.
0.
806.0164m
22.1486m
22.1424m
894.1238m
894.1238m
3.9889m
4.2006m
gm
gds
84.1846u
73.4138u
gmb
894.3095u
927.0356u
49.1459f
43.2542f
cdtot
1.4505
cgtot
56.5600f
56.5745f
0.
cstot
108.9160f
108.8336f
cbtot
vth
584.0239m
573.2955m
vdsat
172.3376m
178.0416m
vod
221.9925m
232.7209m
g m 4.2mS
S
=
= 7.65
I D 549 μA
A
118.8806f
113.0007f
cgs
39.4669f
39.3691f
cgd
7.2418f
7.2416f
C gs = 39.4 fF
B. Murmann
Good agreement.
EE 214 Lecture 6
9
Frequency Response
15
11.85dB = 3.91
3dB
|vo/vi| [dB]
10
5
0 0
10
10
1
10
2
10
3
f [MHz]
•
Simulation result: f-3dB=184MHz; hand analysis: f-3dB=398MHz (!)
B. Murmann
EE 214 Lecture 6
10
73
Interpretation
•
Even though the estimated value of Cgs matches simulation data
very well, there is a large error in the estimated bandwidth
– Means that our simple bandwidth expression is quite
inaccurate
•
The reason for this discrepancy is that we have neglected
"extrinsic" capacitances that affect the frequency response of
our circuit in a significant way
•
This example motivates the need for a more accurate
capacitance model
B. Murmann
EE 214 Lecture 6
11
Extrinsic Capacitance
Cjsb
Cjdb
•
Overlap capacitance
– Gate to source and gate to drain
•
Junction capacitance
– Source to bulk and drain to bulk
B. Murmann
EE 214 Lecture 6
12
74
Overlap Capacitance
•
Two components
– Direct overlap ~ CoxWLoverlap
– Additional component due to fringing field
• Non-negligible in modern technology (gate thickness is large
compared to other feature sizes)
•
EE214 technology parameters (capacitance per width)
– NMOS: Col= 0.23fF/μm
– PMOS: Col= 0.48fF/μm
B. Murmann
EE 214 Lecture 6
13
Junction Capacitance
•
Two components
– Area (AS, AD) and Perimeter (PS, PD)
C jdb =
•
AD ⋅ C j
⎛ VDB ⎞
⎜1 +
⎟
PB ⎠
⎝
mj
+
PD ⋅ C jsw
⎛ VDB ⎞
⎜1 +
⎟
PB ⎠
⎝
mjsw
EE214
Technology
Cj
Cjsw
mj, mjsw
PB
NMOS
0.85 fF/μm2
0.49 fF/μm
0.39
0.51V
PMOS
fF/μm2
0.48 fF/μm
0.48
0.93V
1.11
HSpice automatically calculates junction capacitance based on
W and a geometry factor ("hdif")
– May need specify AS, AD, PS, PD in other simulators
B. Murmann
EE 214 Lecture 6
14
75
Layout Dependence (hdif=0.5μm)
AS = AD = 1μm ⋅ W
PS = PD = 2 μm + 2W
(1μm=2*hdif)
AS = 1μm ⋅W
W
AD = 1μm ⋅
2
PS = 4 μm + 2W
PD = 2 μm + W
B. Murmann
EE 214 Lecture 6
15
MOS Capacitor Summary
Subthreshold
Triode
Cgs
Col
½ WLCox+ Col
Cgd
Col
½ WLCox+Col
Col
"small"
"small"
Cgb
B. Murmann
⎛ 1
1 ⎞
⎟⎟
⎜⎜
+
⎝ CCB WLCox ⎠
Active
2/
3
WLCox + Col
−1
Csb
Cjsb
Cjsb+ CCB/2
Cjsb+ 2/3CCB
Cdb
Cjdb
Cjdb+ CCB/2
Cjdb
EE 214 Lecture 6
16
76
A Closer Look at Gate Capacitance (Simulation)
NMOS W/L=10/0.35, VDS
=0.5V
14
Capacitance [fF]
12
Cgs
10
8
Cgd
6
4
Cgb
2
0
0
0.5
1
1.5
2
V
GS
B. Murmann
2.5
3
3.5
[V]
EE 214 Lecture 6
17
Observations
•
Cgs, Cgd show gradual transitions
– From subthreshold to active
– From active to triode
•
Cgb is non-zero and significant across all operating regions
– Cgb is significant even in active and triode regions since
channel is not a perfect shield; it allows gate field lines to pass
through and terminate at the bulk
•
In our 0.35μm technology, Cgd is approximately constant across
active region and ~10-20% of Cgs
– In ≤ 90nm technologies Cgd can be as large as Cgs (!)
B. Murmann
EE 214 Lecture 6
18
77
Improved Definition of fT
•
So far, we had
fT =
•
Taking extrinsic capacitances into account, we redefine
fT =
•
1 gm
2π C gs
1
gm
1 gm
=
2π C gs + C gb + C gd 2π C gg
The curve on the following slide was generated with the HSpice
deck shown on slide 15, lecture 4, using the following command
– .probe ft
= par('1/2/3.142*gmo(mn1)/cggbo(mn1)')
B. Murmann
EE 214 Lecture 6
19
Improved Transit Frequency Chart
NMOS L=0.35um
30
Improved estimate using C
gg
Previous estimate
25
fT [GHz]
20
15
11.25
10
5
0
B. Murmann
0
5
10
15
gm /ID [S/A]
EE 214 Lecture 6
20
25
20
78
Transit Frequency, NMOS
EE214 technology, NMOS, 0.35...0.7um
14
12
f T [GHz]
10
8
6
4
2
5
10
15
20
gm/I D [S/A]
B. Murmann
EE 214 Lecture 6
21
Transit Frequency, PMOS
EE214 technology, PMOS, 0.35...0.7um
7
6
f T [GHz]
5
4
3
2
1
5
10
15
20
gm/I D [S/A]
B. Murmann
EE 214 Lecture 6
22
79
(gm/ID)*fT, NMOS
EE214 technology, NMOS, 0.35...0.7um
90
gm/I D*f T [GHz*S/A]
80
70
60
50
40
30
20
5
10
15
20
gm/I D [S/A]
B. Murmann
EE 214 Lecture 6
23
(gm/ID)*fT, PMOS
EE214 technology, PMOS, 0.35...0.7um
40
gm/I D*f T [GHz*/A]
35
30
25
20
15
10
5
5
10
15
20
gm/I D [S/A]
B. Murmann
EE 214 Lecture 6
24
80
PMOS Well Capacitance
•
In the EE214 (N-well) technology, the PMOS transistor is a 5
terminal device
– G, D, S, B, Substrate
•
N-well forms a PN junction with the substrate
– Often "AC shorted" when N-well=VDD, Substrate=GND
– Not shorted when we connect N-well to source!
• Resulting capacitance ~ 0.05 fF/μm2
• Not modeled in Spice! Must add extra diode manually in this case
B. Murmann
EE 214 Lecture 6
25
Model for PMOS Well Capacitance
•
Model available in ee214_hspice.txt:
* well-to-substrate diode
* example instantiation (area = 10um*10um = 100pm^2)
*
(anode)
(cathode) (model) (area)
* d1 sub_node well_node
dwell
100p
.model dwell d cj0=1e-4 is=1e-5 m=0.5 bv=40
B. Murmann
EE 214 Lecture 6
26
81
Complete Small Signal Model (Active Region)
B. Murmann
EE 214 Lecture 6
27
+ A Note on Transcapacitance
•
It turns out that this complicated model still doesn't describe the
parasitics with 100% accuracy
– Since MOSFETS are 4-terminal devices, we are really
dealing with a "4-terminal capacitor" and not with a network
of simple two-terminal capacitors
•
In practice, such "transcapacitance" effects are rarely relevant
for design & hand analysis
– Model based on two-terminal capacitors is usually good to
within a few percent
– We'll simply ignore transcapacitance and use Spice as a
final check to see if this was OK…
•
In case you are curious about more details, please refer to
section "Introducing Transcapacitance" in the HSpice manual
B. Murmann
EE 214 Lecture 6
28
82
HSpice .OP Output Variables
HSpice (.OP)
cdtot
cgtot
cstot
cbtot
cgs
cgd
B. Murmann
43.2542f
56.5745f
108.8336f
113.0007f
39.3691f
7.2416f
Corresponding Small Signal
Model Elements
≡ Cgd + Cdb
cgtot ≡ Cgs + Cgd + Cgb
cstot ≡ Cgs + Csb
cbtot ≡ Cgb + Csb+ Cdb
cgs ≡ Cgs
cgd ≡ Cgd
cdtot
EE 214 Lecture 6
29
Cgd and Cdb Calculations (1)
•
Cgd and Cdb can be calculated accurately using the data on
slides 13 and 14
– But we need to know the device width (and VDS)
•
For design in a "normalized" space, it is often desirable to have
width-independent estimates for these caps
– E.g. Cdg/Cgg and Cdb/Cgg
•
The next slide shows simulated values for these ratios at
L=0.35μm and VDS=1.5V
B. Murmann
EE 214 Lecture 6
30
83
Cgd and Cdb Calculations (2)
NMOS, L=0.35um, VDS=1.5V
1
PMOS, L=0.35um, VDS=1.5V
1
C /C
db
C /C
0.8
gd
gg
0.8
gg
0.6
0.6
0.4
0.4
0.2
0.2
0
5
10
15
g /I
m D
20
Cdb/Cgg
Cgd/Cgg
0
5
10
20
PMOS:
kgdp=Cgd/Cgg ≅0.26
kdbp=Cdb/Cgg ≅0.80
NMOS:
kgdn=Cgd/Cgg ≅0.13
kdbn=Cdb/Cgg ≅0.65
B. Murmann
15
gm/I D [S/A]
[S/A]
EE 214 Lecture 6
31
Cgd and Cdb Calculations (3)
•
For lengths other than 0.35μm, we can use
C gd
C gg
•
≅ k gd ⋅
L = Lx
0.35 μm
Lx
Cdb
C gg
≅ k db ⋅
L = Lx
0.35 μm
Lx
For a drain bias other than 1.5V, we would in principle need
another adjustment factor for Cdb
– But, since the dependence on VDS is weak (square root), it is
often not worth the effort
• Cdb can be significant, but it is often not the dominant
capacitance
B. Murmann
EE 214 Lecture 6
32
84
"Level 2" Figures of Merit
gm
ID
gm
C gg
gm
g ds
B. Murmann
EE 214 Lecture 6
C gd
C gg
Cdb
C gg
33
85
Lecture 7
Miller Approximation
Zero-Value Time Constant Analysis
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 7
1
Overview
•
Reading
– 7.1, 7.2.0, 7.2.1 (Miller Effect in CS Stage, only pp. 488-493)
– 7.3.0, 7.3.1 7.3.2 (Zero-Value Time Constant Analysis)
– 7.3.3 (Cascade Amplifier Frequency Response)
– Supplementary document "Bandwidth estimation techniques," by
Tom Lee (optional, see website).
•
Introduction
– Last lecture, we found that using a simple circuit model based on
intrinsic capacitance only is not sufficient for accurate bandwidth
prediction in our CS stage. Having learned about the involved
extrinsic capacitances, we are now in a position to improve our
hand analysis and match the Spice result with good precision. To
simplify the analysis, we will utilize the so-called "Miller
Approximation." Next, we will take at look at the the "Zero-Value
Time Constant Analysis" as an alternative method, which is useful
for a much broader class of circuits.
B. Murmann
EE 214 Lecture 7
2
86
Design Example 2 Revisited
H( s ) = ?
≅ RL
B. Murmann
EE 214 Lecture 7
3
Bandwidth Estimation (1)
•
To simplify the problem, let's first neglect Cdb
– We'll need to check later if this assumption was OK
•
Next, we cut the circuit as shown below and calculate the
equivalent admittance seen looking into the right side of the cut
Y( s ) =
B. Murmann
i( s )
vgs ( s )
EE 214 Lecture 7
4
87
Bandwidth Estimation (2)
i = (vgs − vo )⋅ sC gd
∴Y ( s ) =
B. Murmann
g m vgs +
i( s )
= (1 − Av ( s )) ⋅ sC gd
vgs ( s )
vo
+ (vo − vgs )⋅ sC gd = 0
RL
with
C
⎛
⎜ 1 − s gd
gm
Av ( s ) = − g m RL ⎜⎜
1 + sRL C gd
⎜
⎝
EE 214 Lecture 7
⎞
⎟
⎟
⎟
⎟
⎠
5
Miller Approximation
•
Assuming that the poles and zeros in Av(s) occur at much higher
frequencies than the bandwidth we are trying to estimate, it is
OK to replace Av(s) with its DC value
– This is known as the "Miller approximation"
– It is always a good idea to check (later) if the approximation
was indeed valid!
•
With the Miller approximation, we have
Y ( s ) ≅ (1 + g m RL ) ⋅ sC gd
•
This is the same as a capacitor to GND with a value of (1+gmRL)
times Cgd
– "Amplified capacitance"
B. Murmann
EE 214 Lecture 7
6
88
Generalization
vtest
vtest
Z
=
=
v
−
A
v
itest
1 − Av
test
v test
Z
Yin = Y (1 − Av )
Z in =
•
Interesting cases
– Av=0 ⇒ Zin=Z (no surprise…)
– Av=1 ⇒ Zin=∞
• "Bootstrapping"
– Av>1, e.g. Av=2 ⇒ Zin=-Z (negative!)
– Av<0, ⇒ Zin=Z/(1+|Av|)
• Impedance reduction
B. Murmann
EE 214 Lecture 7
7
Modified Input Network
f −3dB ≅
•
1
1
2π Ri C gs + C gb + (1 + g m RL ) ⋅ C gd
[
]
Very simple!
– At least much simpler than using exact expressions
• See e.g. equation 7.19 in the text
•
Next, we'll verify if the involved assumptions hold in our example
circuit, and also see how accurately we can match Spice
B. Murmann
EE 214 Lecture 7
8
89
Improved Bandwidth Estimate
•
Using the transit frequency chart, we find
C gg =
f −3 dB ≅
•
1 gm
1
4mS
=
= 57 fF
2π fT 2π 11.25GHz
1
1
2π Ri C gs + C gb + (1 + g m RL )C gd
[
]
=
1
1
1
1
=
2π Ri C gg + g m RL C gd
2π Ri C gg 1 + g m RL k gdn
=
1
1
= 184 MHz
2π 10 kΩ [57 fF (1 + 4 ⋅ 0.13)]
[
]
[ (
)]
Our simulation result from last lecture was
f −3 dB ,Spice = 184 MHz
B. Murmann
EE 214 Lecture 7
9
Assumption Check (1)
•
It is interesting (and necessary in general) to check how good
the Miller assumption was in this analysis
•
We assumed that
C
⎛
⎜ 1 − s gd
gm
Av ( s ) = − g m RL ⎜⎜
1 + sRLC gd
⎜
⎝
⎞
⎟
⎟ ≅ −g R
m L
⎟
⎟
⎠
up to the frequency of interest (~184MHz)
•
Let's check this by calculating the magnitudes of the pole and
zero in Av(s)
1
1
1
1
=
= 21.5GHz
2π RL C gd 2π 1kΩ ⋅ 7.4 fF
1 gm
1 4mS
=
= 86GHz
2π C gd 2π 7.4 fF
B. Murmann
EE 214 Lecture 7
10
90
Assumption Check (2)
•
We had also assumed that the impact of Cdb is negligible
– How can we make sure this is OK?
•
One possible solution: Re-derive Av(s) with Cdb present
Cgd
Ri
vi
+
vgs
-
i(s)
Cgs+Cgb Y(s)
g m vgs +
RL
+
Cdb vo
-
vo
+ (vo − vgs )⋅ sC gd + v0 ⋅ sCdb = 0
RL
Av ( s ) =
B. Murmann
gmvgs
vo ( s )
vgs ( s )
EE 214 Lecture 7
11
Assumption Check (3)
C
⎞
⎛
⎟
⎜
1 − s gd
gm
v (s)
⎟
∴ Av ( s ) = o
= − g m RL ⎜⎜
1 + sRL ( C gd + Cdb ) ⎟
vgs ( s )
⎟
⎜
⎠
⎝
•
Zero is unchanged, but Cdb lowers pole frequency
•
Using Cdb ≅ Cgg·0.65, we find Cdb ≅ 37fF, and hence
1
1
1
1
=
= 3.6 GHz
2π RL ( C gd + Cdb ) 2π 1kΩ ⋅ ( 7.4 fF + 37 fF )
•
Still not too bad, since 3.6GHz >> 180MHz
– But what if we now add some load capacitance at the output
of the amplifier?
• Will appear in parallel with Cdb
B. Murmann
EE 214 Lecture 7
12
91
Effect of Output Loading (1)
C
⎞
⎛
⎟
⎜
1 − s gd
gm
vo ( s )
⎟
⎜
= − g m RL ⎜
Av ( s ) =
1 + sRL ( C gd + Cdb + CL ) ⎟
vgs ( s )
⎟
⎜
⎠
⎝
•
Suppose CL=10pF, then
1
1
1
1
≅
= 16MHz << 180MHz !
2π RL ( C gd + Cdb + C L ) 2π 1kΩ ⋅10 pF
B. Murmann
EE 214 Lecture 7
13
Effect of Output Loading (2)
•
Obviously, with the added CL=10pF, the Miller approximation
does not hold any more
•
So what is going on in this modified circuit, with large CL?
– Cgd is "amplified" only at low frequencies, before CL
"destroys" the gain from vgs to vo
– The pole caused by Cgd no longer impairs the 3-dB
bandwidth of the circuit
– Intuitively, the bandwidth must now be somehow set by CL
• After all, it destroys the gain from vgs to vo, which implies that
the gain from vi to vo must also roll off
•
With very large CL, it is intuitively clear that the approximate
bandwidth of the circuit should be ~1/(2πRLCL)
– But how about other cases and circuits that don't have a
straightforward Miller approximation?
B. Murmann
EE 214 Lecture 7
14
92
Zero-Value Time Constant Analysis
•
Fortunately, there is a more general method that allows us to
estimate the bandwidth of arbitrary circuits (within limits)
– Without going though the pain of deriving the complete sdomain transfer function
•
"ZVTC Analysis," or "Open Circuit Time Constant Analysis"
•
Here's how it works
– Remove all but one capacitor. Short independent voltage
sources, remove independent current sources
– Calculate resistance seen by capacitor and compute τj=RjoCj
– Repeat for all capacitors in the circuit
– Sum all time constants and calculate bandwidth estimate
ω−3dB ≅
B. Murmann
1
∑τ
j
EE 214 Lecture 7
15
Example (1)
•
Step 1:
τ 1 = Ri C1
B. Murmann
EE 214 Lecture 7
16
93
Example (2)
•
Step 2:
R2o =
vtest vgs + RL ( g m vgs + itest ) itest Ri + RL ( g mitest Ri + itest )
=
=
= Ri + RL + g m RL Ri
itest
itest
itest
τ 2 = ( Ri + RL + g m RL Ri )C2
•
Step 3:
R3o = RL
τ 3 = RLC3
B. Murmann
EE 214 Lecture 7
17
Bandwidth Estimate Based on ZVTC
∴ f −3dB ≅
•
1
1
1
1
=
2π τ 1 + τ 2 + τ 3 2π Ri C1 + ( Ri + RL + g m RL Ri )C2 + RL C3
Reality check using numbers from design example 2
C1 = C gs + C gb = 49.6 fF
C2 = C gd = 7.4 fF
C3 = Cdb = 35.8 fF
∴ f −3dB ≅
•
τ 1 = 10kΩ ⋅ 49.6 fF = 496 ps
τ 2 = ( 5 ⋅10kΩ + 1kΩ ) ⋅ 7.4 fF = 377 ps
τ 3 = 1kΩ ⋅ 35.8 fF = 36 ps
1
1
= 175MHz
2π τ 1 + τ 2 + τ 3
Not bad!
– Spice simulation gave 184MHz
– Miller approximation result was 184MHz
B. Murmann
EE 214 Lecture 7
18
94
Inclusion of CL
•
What happens if again, we consider adding a large load
capacitance (CL)?
C1 = C gs + C gb = 49.6 fF
τ 1 = 10kΩ ⋅ 49.6 fF = 496 ps
τ 2 = ( 5 ⋅10kΩ + 1kΩ ) ⋅ 7.4 fF = 377 ps
τ 3 = 1kΩ ⋅10 pF = 10,000 ps
C2 = C gd = 7.4 fF
C3 = Cdb + C L ≅ 10 pF
∴ f −3dB ≅
1
1
= 14.6MHz
2π τ 1 + τ 2 + τ 3
•
Now the third time constant dominates and significantly reduces
our bandwidth estimate
•
Looks like this is a powerful method
– Miller effect is taken care of, output loading effect is included
•
Most importantly though, the method provides us with insight
about the limiting elements in our circuit!
B. Murmann
EE 214 Lecture 7
19
How ZVTCs Work
•
Intuition
– Each time constant relates to the bandwidth that we would
get if no other capacitors were present
• "Local bandwidth bottleneck"
– For simplicity, the ZVTC method linearly combines the local
bottlenecks to estimate the overall bandwidth
•
Mathematically, the ZVTC method is based on the
approximation
H( s ) =
•
K
K
≅
bn s n + bn −1s n −1 + ... + b1s + 1 b1s + 1
Can show that
– b1 corresponds to the sum of all time constants in the circuit
– This approximation is OK unless there are "undamped"
complex poles, or several limiting poles with comparable
magnitude
B. Murmann
EE 214 Lecture 7
20
95
A Simple Example
Exact bandwidth:
ω−3dB =
1
⋅
RC
ZVTC Method:
2 −1 =
0.644
RC
ω−3dB =
0.5
RC
•
Roughly -22% error
•
ZVTC estimates tend to be conservative
– Actual bandwidth will almost always be at least as high as
estimate
B. Murmann
EE 214 Lecture 7
21
ZVTC Accuracy and Other Caveats
•
Accuracy tends to be OK when there is a single dominant pole
– Not surprising, since the "approximation" shown on slide 20
makes no error for a single pole system
– Fortunately, many practical circuits indeed have a somewhat
dominant pole
•
Some elements, like AC coupling caps, must be eliminated
before applying the ZVTC method
– These caps are meant to be shorts at high frequencies, and
do not degrade the signal bandwidth
– Can use method of "short circuit time constants" to
determine coupling cap sizes
• See supplementary handout "Bandwidth Estimation
Techniques"
•
The ZVTC method tells us nothing about zeros in the transfer
function!
B. Murmann
EE 214 Lecture 7
22
96
Lecture 8
Electronic Noise
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 8
1
Overview
•
Reading
– 11.1 (Noise Introduction)
– 11.2.2 (Thermal Noise)
– 11.3.3 (MOS Transistor Noise)
– Supplementary Handout: "Introduction to Noise" by Daniel Cooley
•
Introduction
– Electronic noise is a significant and fundamental issue in the design
of high performance analog circuits. The noise level of a circuit
affects the "fidelity" or accuracy of the signals that are being
processed. As we shall see, minimizing electronic noise is costly;
there exists a steep tradeoff with power dissipation and bandwidth.
Today's lecture will provide an introduction to electronic noise at the
component level (resistors and MOSFETs). We will use these
results in the remainder of the course to analyze the impact of
noise in various circuits.
B. Murmann
EE 214 Lecture 8
2
97
Types of Noise
•
"Man made noise", interference noise
– Signal coupling
– Substrate coupling
– Finite power supply rejection
– Solutions
• Fully differential circuits
• Layout techniques
•
"Electronic noise" or "device noise" (focus of this lecture)
– Fundamental
• E.g. "thermal noise" caused by random motion of carriers
– Technology related
• "Flicker noise" caused by material defects and "roughness"
B. Murmann
EE 214 Lecture 8
3
Significance of Electronic Noise (1)
Signal-to-Noise Ratio
SNR =
"Signal"
B. Murmann
Psignal
Pnoise
∝
2
Vsignal
Pnoise
"Noise"
EE 214 Lecture 8
4
98
Significance of Electronic Noise (2)
•
Example: Noisy image
http://www.soe.ucsc.edu/~htakeda/kernelreg/kernelreg.htm
B. Murmann
EE 214 Lecture 8
5
Significance of Electronic Noise (3)
•
The "fidelity" of electronic systems is often determined by their SNR
– Examples
• Audio systems
• Imagers, cameras
• Wireless and wireline transceivers
•
Electronic noise directly trades with power dissipation and speed
– In most circuits, low noise dictates large capacitors (and/or small R,
large gm), which means high power dissipation
•
Noise has become increasingly important in modern technologies with
reduced supply voltages
– SNR ~ Vsignal2/Pnoise ~ VDD2/Pnoise
•
Designing a low power, high performance circuit requires good
understanding of electronic noise!
B. Murmann
EE 214 Lecture 8
6
99
Topics/Questions
•
How to model noise of circuit components
•
How to calculate/simulate noise performance of a complete
circuit (e.g. SNR)
•
Do we need to worry about electronic noise in all circuits?
– The answer must be no; digital folks never talk about this…
– Need to get a feel for situations where noise matters
B. Murmann
EE 214 Lecture 8
7
Ideal Resistor
i(t)
1V/1kΩ
•
Constant current, independent of time
•
Non-physical
– In a physical resistor, carriers "randomly" collide with lattice
atoms, giving rise to small current variations over time
B. Murmann
EE 214 Lecture 8
8
100
Physical Resistor
i(t)
1V/1kΩ
•
"Thermal Noise" or "Johnson Noise"
– J.B. Johnson, "Thermal Agitation of Electricity in Conductors,"
Phys. Rev., pp. 97-109, July 1928.
•
Can model random current component e.g. using a noise
current source in(t)
B. Murmann
EE 214 Lecture 8
9
Properties of Thermal Noise
•
Present in any conductor
•
Independent of DC current flow
•
Instantaneous noise value is unpredictable since it is a result of
a large number of random, superimposed collisions with
relaxation time constants of τ ≅ 0.17ps
– Consequences:
• Gaussian amplitude distribution
• Knowing in(t) does not help predict in(t+Δt), unless Δt is on the
order of 0.17ps (cannot sample signals this fast)
• The power generated by thermal noise is spread up to very
high frequencies (1/τ ≅ 6,000Grad/s)
•
The only predictable property of thermal noise is its average
power!
B. Murmann
EE 214 Lecture 8
10
101
Average Power
•
For a deterministic current signal with period T, the average
power is given by
T/2
1
Pav =
i 2 (t ) ⋅ R ⋅ dt
∫
T −T / 2
•
This definition can be extended to capture non-deterministic
random signals
– Assuming a real, stationary and ergodic random process
T/2
1
in2 (t ) ⋅ R ⋅ dt
T →∞ T ∫
−T / 2
Pn = lim
•
For notational convenience, we typically drop R in the above
expression and work with "mean square" current (or voltages)
T/2
1
i = lim
in2 (t ) ⋅ dt
T →∞ T ∫
−T / 2
2
n
B. Murmann
EE 214 Lecture 8
11
Thermal Noise Spectrum
•
The so-called power spectral density (PSD) shows how much
power a signal caries at a particular frequency
•
In the case of thermal noise, the power is spread uniformly up to
very high frequencies (about 10% drop at 2,000GHz)
PSD(f)
n0
f
•
The total average noise power Pn in a particular frequency band
can be found by integrating the PSD
f2
Pn = ∫ PSD( f ) ⋅ df
f1
B. Murmann
EE 214 Lecture 8
12
102
Thermal Noise Power
•
Nyquist showed that the noise PSD of a resistor is
PSD( f ) = n0 = 4 ⋅ kT
•
k is the Boltzmann constant and T is the absolute temperature
– 4kT = 1.66·10-20 Joules at room temperature
•
The total average noise power of a resistor in a certain
frequency band is therefore
f2
Pn = ∫ 4 kT ⋅ df = 4 kT ⋅ ( f 2 − f 1 ) = 4 kT ⋅ Δf
f1
B. Murmann
EE 214 Lecture 8
13
Equivalent Noise Generators
•
Can model the noise using either an equivalent voltage or
current generator
in2 =
vn2 = Pn ⋅ R = 4 kT ⋅ R ⋅ Δf
Pn
1
= 4 kT ⋅ ⋅ Δf
R
R
For R = 1kΩ :
For R = 1kΩ :
vn2
V2
= 16 ⋅ 10 −18
Δf
Hz
in2
A2
= 16 ⋅ 10 − 24
Δf
Hz
vn2
= 4 nV / Hz
Δf
in2
= 4 pA / Hz
Δf
Δf = 1MHz ⇒ vn2 = 4 μV
B. Murmann
Δf = 1MHz ⇒ in2 = 4 nA
EE 214 Lecture 8
14
103
Two Resistors in Series
(
vn2 = vn1 − vn 2
•
)
2
= vn21 + vn22 − 2 ⋅ vn1 ⋅ vn 2
Since vn1(t) and vn2(t) are statistically independent, we have
vn2 = vn21 + vn22 = 4 ⋅ kT ⋅ (R1 + R2 ) ⋅ Δf
•
Always remember to add independent noise sources using
mean squared quantities
– Never add RMS values!
B. Murmann
EE 214 Lecture 8
15
MOSFET Thermal Noise (1)
•
As one would expect, the noise of a MOSFET operating in the
triode region is equal to that of a resistor
•
In the forward active region, the thermal noise of a MOSFET can
be modeled using a drain current source with spectral density
id2 = 4 kT ⋅ γ ⋅ g m ⋅ Δf
•
For a long channel MOSFET γ=2/3
•
For the past ten years, researchers have been debating over the
value of γ in short channels
– Preliminary (wrong) results had suggested that in short
channels γ can be as high as 2…5 due to hot carrier effects
B. Murmann
EE 214 Lecture 8
16
104
MOSFET Thermal Noise (2)
[Scholten]
•
Fortunately, these discussions have come to an end with the
conclusion that short channels have γ ≅ 1
– A. J. Scholten et al., "Noise modeling for RF CMOS circuit simulation," IEEE
Trans. Electron Devices, pp. 618-632, Mar. 2003.
– R. P. Jindal, "Compact Noise Models for MOSFETs," IEEE Trans. Electron
Devices, pp. 2051-2061, Sep. 2006.
B. Murmann
EE 214 Lecture 8
17
Spice Simulation (1)
* EE214 MOS device noise simulation
vd
vm
vg
mn1
h1
dd
dd
g
d
c
0
d
0
g
0
1.5
0
dc 0.8 ac 1
0 0 nch214
ccvs vm 1
L=0.35u W=10u
.op
.ac dec 100 10k 1gig
.noise v(c) vg
1.5V
dd
0V
d
c
CCVS
1V/A
.options post brief
.lib './ee214_hspice.txt' nominal
.end
B. Murmann
EE 214 Lecture 8
18
105
Spice Simulation (2)
avg(id2)/df [A 2/Hz]
HSpice ("outnoise")
4kT*2/3*gm
10
-22
10
-23
10
-24
10
(gm=1.28mS)
-2
10
-1
10
0
10
1
10
2
10
3
f [MHz]
B. Murmann
EE 214 Lecture 8
19
1/f Noise
•
Also called "flicker noise" or "pink noise"
•
Caused by traps near Si/SiO2 interface that randomly capture and
release carriers
•
Occurs in virtually any device, but is most pronounced in MOSFETS
•
One (empirical) way to model flicker noise:
– Known as "NLEV=2" HSpice model
•
•
i12/
f
=
Kf
2
gm
Δf
W ⋅L f
Cox
For other models, see HSpice manual or
– D. Xie et al, "SPICE Models for Flicker Noise in n-MOSFETs from
Subthreshold to Strong Inversion," IEEE Trans. CAD, pp. 12931303, Nov. 2000
Kf is strongly dependent on technology; numbers for EE214 0.35μm
CMOS technology:
– Kf,NMOS = 0.5·10-25 V2F
– Kf,PMOS = 0.25·10-25 V2F
B. Murmann
EE 214 Lecture 8
20
106
1/f Noise Corner
•
By definition, the frequency at which the flicker noise density
equals the thermal noise density
Kf
Cox
2
gm
Δf
= 4 kTγ ⋅ g m ⋅ Δf
W ⋅ L f co
⇒ f co =
=
Kf
1 gm
4 kTγ Cox W ⋅ L
1 1 ⎛ gm
⎜
4 kTγ Cox L ⎜⎝ I D
Kf
⎞⎛ I D ⎞
⎟⎟⎜
⎟
⎠⎝ W ⎠
•
Example: NMOS, L=0.35μm, gm/ID=10V-1 → ID/W=10A/m
⇒ fco = 244kHz
•
In more recent technologies, 1/f corner frequencies can be on
the order of 10MHz
B. Murmann
EE 214 Lecture 8
21
1/f Noise Contribution (1)
•
Just as with white noise, the total 1/f noise contribution is found
by integrating its spectral density
i12/ f ,tot
f2
=
2
gm
Δf
C W ⋅L f
f1 ox
∫
Kf
2
2
⎛ f2 ⎞ K f gm
⎛ f ⎞
gm
⎜
⎟
ln⎜ ⎟ =
2.3 log ⎜⎜ 2 ⎟⎟
=
Cox W ⋅ L ⎝ f 1 ⎠ Cox W ⋅ L
⎝ f1 ⎠
Kf
•
Total flicker noise depends on number of frequency decades
– Same flicker noise power in 1…10 Hz as in 1…10 GHz
– RMS noise proportional to sqrt(# of frequency decades)
•
So, does flicker noise matter?
– Look at noise integral to see its relative contribution to total
drain current noise
B. Murmann
EE 214 Lecture 8
22
107
Spectrum [A2/Hz]
1/f Noise Contribution (2)
10
10
10
10
-20
-21
•
-22
-23
10
2
10
4
10
6
10
8
Sqrt(Integral) [nA]
f [Hz]
For circuits with
"high bandwidth",
flicker noise is often
insignificant
– Beware of
exceptions…
100
50
0 2
10
10
4
10
6
10
8
f [Hz]
B. Murmann
EE 214 Lecture 8
23
Lower Integration Limit
•
Does the flicker noise PSD go to infinity for f→0? Fun
discussions in
– E. Milotti, "1/f noise: a pedagogical review," available at
http://arxiv.org/abs/physics/0204033
•
Even if the PSD goes to infinity, do we care?
– Let's say we are sensing a signal for a very long time (down
to a very low frequency), e.g.
• 1 year ≅ 32 Msec, 1/year ≅ 0.03 μHz
– Number of frequency decades in 1/year to 100Hz ≅ 10
• Means instead of 7 decades, we'd now have 17 in the previous
example
• sqrt(17/7) = 1.56 → Only 56% more flicker noise!
•
More fun reading on flicker noise
– H. Schmid, "Aaargh! I just loooove flicker noise," IEEE Circuits and
Systems Magazine, Vol.7, Issue 1, pp. 32-35, 2007.
B. Murmann
EE 214 Lecture 8
24
108
MOS Model with Noise Generator
Noiseless!
id2
B. Murmann
= 4 kT ⋅ γ ⋅ g m ⋅ Δf +
Kf
Cox
2
gm
Δf
W ⋅L f
EE 214 Lecture 8
25
Other MOSFET Noise Sources
•
Gate noise
– "Shot noise" from gate leakage current
– Noise from to finite resistance of gate material
– Noise due to randomly changing potential/capacitance
between channel and bulk
• Relevant only at very high frequencies
•
Bulk noise
•
Source barrier noise in very short channels (Navid & Dutton)
– Shot noise from carriers injected across source barrier
•
More in EE314…
B. Murmann
EE 214 Lecture 8
26
109
Lecture 9
Electronic Noise (Continued)
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 9
1
Overview
•
Reading
– 11.4 (Circuit Noise Calculations)
– 11.5 (Equivalent Input Noise Generators)
– 11.9 (Noise Bandwidth)
•
Introduction
– Having established the basic noise mechanisms in
MOSFETS, today's lecture looks at noise in circuits. We will
learn how to calculate the signal-to-noise ratio in a basic RC
circuit and a common source amplifier. These examples are
useful prerequisites for analyzing more complicated circuits
(e.g. OTAs, later in this class). Furthermore, we will use
these simple examples to develop a basic feel for the
relevance of noise and associated tradeoffs.
B. Murmann
EE 214 Lecture 9
2
110
Noise in Circuits (1)
•
Most interesting circuits have more than one relevant noise
source
•
In order to quantify the net effect of all noise sources, we must
refer the noise sources to a single "interesting" node pair of the
circuit
– Usually output or input
B. Murmann
EE 214 Lecture 9
3
Noise in Circuits (2)
•
Output referred noise
– Refer noise to output via
individual noise transfer
functions
– Physical concept, exactly
what one would measure in
the lab
•
Input referred noise
– Represent total noise via a
fictitious input source that
captures all circuit-internal
noise sources
– Useful primarily for "fair"
comparisons
• E.g. independent of circuit
gain
B. Murmann
EE 214 Lecture 9
4
111
Datasheet Example
B. Murmann
EE 214 Lecture 9
5
Circuit Example 1
•
Let's calculate
– Output referred noise
– Input referred noise
– Signal-to-noise ratio at output
B. Murmann
EE 214 Lecture 9
6
112
Output Referred Noise
•
Short input voltage, add noise source
•
Calculate transfer function from noise source to output
•
Multiply noise PSD with squared transfer function to get output
PSD
vn2,out
Δf
1
= 4 kT ⋅ R ⋅
1 + sRC
B. Murmann
2
= (4 nV )
2
1
1 + j 2πf ⋅ 1ns
EE 214 Lecture 9
2
7
Input Referred Noise
•
Input referred noise is simply the PSD of resistor (in this example)
v2
1k
R
C
1pF
vn2,in
Δf
B. Murmann
Vout
= 4 kT ⋅ R = (4 nV )2
EE 214 Lecture 9
8
113
+ A Note on Equivalent Input Noise Generators
•
In general, input referred noise must be modeled by an
equivalent voltage and current source
– See section 11.5 in the text (optional)
•
Modeling the noise with a voltage alone is sufficient when the
circuit is indeed driven by an ideal voltage source (always an
approximation)
– Or, when the input impedance of the circuit is large
• This is the case in MOS circuits at low to moderate frequencies
• Different story at RF frequencies
– You'll learn more about this in EE314
B. Murmann
EE 214 Lecture 9
9
Spice Simulation
* EE214 RC circuit noise
-16
outnoise
innoise
ac 1
1k
1pF
.ac dec 100 100 1gig
.noise v(out) vin
.options post brief
.lib './ee214_hspice.txt' nominal
.end
Noise [V2/Hz]
vin in 0
r1 in out
c1 out 0
10
10
-17
10
-18
10
4
10
6
10
8
f [Hz]
B. Murmann
EE 214 Lecture 9
10
114
Signal-to-Noise Ratio
1 2
Vout , peak
2
SNR =
=
f2 2
Pnoise
vn ,out
⋅ df
Δf
Psignal
∫
f1
•
Over which bandwidth should we integrate the noise?
•
Two interesting cases
– The output is measured or observed by a system with finite
bandwidth (e.g. human ear)
• Use frequency range of that system as integration limits
• Applies on a case by case basis
– "Total integrated noise"
B. Murmann
EE 214 Lecture 9
11
Total Integrated Noise (1)
•
Simply the integral over all frequencies (zero to infinity)
•
Relevant metric when output is observed without (significant)
band limiting or when the output is sampled (e.g. switched
capacitor circuits)
•
The total integrated noise in our circuit is
vn2,out ,tot
∞
2
1
kT
= 4 kT ⋅ R ⋅
df =
1 + j 2πf ⋅ RC
C
∫
0
•
Interesting result
– Total integrated noise at the output depends only on C (even
though R is generating the noise)
B. Murmann
EE 214 Lecture 9
12
115
Increasing R increases
the noise power
spectral density, but
also decreases the
bandwidth
– R drops out in the
end result
10
-14
10
-16
10
-18
10
-20
10
-22
10
Sqrt(Integral) [μV]
•
Spectrum [V2/Hz]
Total Integrated Noise (2)
R=1k
R=100k
2
10
60
6
10
f [Hz]
8
10
8
10
10
10
R=1k
R=100k
40
20
0 2
10
B. Murmann
4
10
4
6
10
f [Hz]
10
EE 214 Lecture 9
10
13
SNR (1)
•
Back to our example; plugging in numbers
SNR =
•
Psignal
Pnoise
SNR in dB
1 2
Vout , peak
0.5V 2
0.5V 2
= 2
=
=
= 122 ⋅ 10 6
2
kT
kT
(64 μV )
C
1 pF
(
)
10 log 122 ⋅ 10 6 = 81dB
•
Is this "good"?
•
Typical system requirements
– Audio: SNR ≅ 100dB
– Video: SNR ≅ 60dB
– Gigabit Ethernet Transceiver: SNR ≅ 35dB
B. Murmann
EE 214 Lecture 9
14
116
SNR (2)
•
Assuming Vout,peak = 1V
SNR [dB]
20
40
60
80
100
120
140
•
C [pF]
0.00000083
0.000083
0.0083
0.83
83
8300
830000
Hard to make such small capacitors…
Designer will definitely be concerned
about thermal noise, capacitor sizes
set by SNR
"Hardcore" thermal noise battle…
General rules of thumb
– Up to SNR ~ 40dB, integrated circuits are usually not limited
by thermal noise
– Achieving SNR >100dB is extremely difficult
• Must usually rely on external components or "tricks" (such as
oversampling, see EE315)
B. Murmann
EE 214 Lecture 9
15
MDS and DR
•
Minimum detectable signal (MDS)
– A somewhat arbitrary definition
– Quantifies the signal level in a circuit that yields SNR=1; i.e.
noise power = signal power
•
Dynamic range (DR)
DR =
•
Psignal ,max
MDS
If the noise level in the circuit is independent of the signal level
(not always the case), it follows that the DR is equal to the "peak
SNR", i.e. the SNR with the maximum signal applied
B. Murmann
EE 214 Lecture 9
16
117
Circuit Example 2: Common Source Amplifier
[Ignoring 1/f noise for simplicity]
vo2 ( f )
R
⎛1
⎞
= 4 kT ⎜ + γg m ⎟ ⋅
df
⎝R
⎠ 1 + j 2πf ⋅ RC
vo2,tot
2
∞
2
R
⎛1
⎞
= 4 kT ⎜ + γg m ⎟ ⋅
df
⎝R
⎠ 1 + j 2πf ⋅ RC
∫
0
kT
(1 + γg m R )
C
kT
(1 + γ Av )
=
C
kT
=
⋅α
C
=
B. Murmann
EE 214 Lecture 9
17
CS Stage Noise/Power Tradeoff
•
Assuming that we're already using the maximum available
signal swing, improving the SNR by 6dB means
– Increase C by 4x
– Decrease R by 4x to maintain bandwidth
– Increase gm by 4x to preserve gain
•
Assuming that we can keep gm/ID constant, this means that we
must increase ID by 4x
•
Bottom line
– Improving the SNR in a noise limited circuit by 6dB
("1bit") QUADRUPLES power dissipation !
B. Murmann
EE 214 Lecture 9
18
118
Lecture 10
Backgate Effect
Common Gate Stage
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 10
1
Overview
•
•
Reading
– 1.6.6 (Body Transconductance)
– 3.3.4, 7.2.4.2 (Common Gate Stage)
– 3.4.2.2, 7.3.4 (up to p. 526) (Cascode Stage)
Somewhat tedious
to read…
Introduction
– Having completed our discussion of simple common source
amplifiers, we now continue by exploring alternative ways of
building single transistor stages. First, we will look at the
common gate (CG) stage, followed by a discussion of the
common drain (CD) stage in the next lecture. While CS
stages can usually be configured with source and bulk nodes
tied together, this is often not the case in the CG and CD
configurations. Hence, we'll first need to take a look at the so
called backgate effect (sometimes called "body effect")
which becomes relevant in this context.
B. Murmann
EE 214 Lecture 10
2
119
The "Atoms" of Analog Circuit Design
•
As we've seen from the discussion so far, a common source
stage is sufficient for building a simple amplifier
– How about the other two possible configurations?
•
We'll find that common gate and drain stages can be
incorporated as valuable add-ons, for building "better" amplifiers
•
Interestingly, many analog circuits can be decomposed into a
combination of the above three fundamental building blocks
B. Murmann
EE 214 Lecture 10
3
Bulk Connection
•
In the EE214 (N-well) technology, only the PMOS device has an
isolated bulk connection
•
Newer technologies (e.g. 0.13μm CMOS) also tend to have
NMOS devices with isolated bulk ("twin-well" process)
B. Murmann
EE 214 Lecture 10
4
120
Bulk Connection Scenarios
DD
DD
B. Murmann
EE 214 Lecture 10
5
Backgate Effect (1)
VSB>0
•
With positive VSB, depletion region around source grows
•
Increasing amount of negative fixed charge in depletion region
tends to "repel" electrons coming from source
– Need larger VGS to compensate for this effect
B. Murmann
EE 214 Lecture 10
6
121
Backgate Effect (2)
•
This effect is usually factored in as an effective increase in Vt
•
Detailed analysis shows
Vt = Vt 0 + γ
•
(
2φ f + VSB − 2φ f
)
A change in Vt also means a change in drain current
– Define small signal backgate transconductance
g mb =
∂I D
∂I
=− D
∂VSB
∂VBS
g mb
∂Vt
γ
∂V ∂I D ∂VGS
=
=− t
=
gm
∂VSB 2 VSB + 2φ f
∂VSB ∂Vt ∂I D
-1
B. Murmann
EE 214 Lecture 10
7
gmb Simulation
NMOS W/L=10/0.35um, V
24
-V =0.2V
GS
t
Spice
Fit using 2φf=0.6V, γ=0.36V-1
22
gmb/gm [%]
20
18
16
14
12
10
0
B. Murmann
0.5
1
VBS [V]
EE 214 Lecture 10
1.5
2
8
122
Modified Small Signal Model
B. Murmann
EE 214 Lecture 10
9
Common Gate Stage
Cgd+Cdb
-(gm+gmb)vi
ii
Define:
B. Murmann
CS = C gs + Csb
EE 214 Lecture 10
+
RS v i
-
RL
Vo
ro
Cgs+Csb
g' m = g m + g mb
10
123
CG Current Transfer
io
g' m
≅
ii g' + sC + 1
m
S
RS
≅
g' m RS
1
⋅
1 + g' m RS 1 + s RS CS
1 + g' m RS
io
1
≅
ii 1 + s CS
g' m
B. Murmann
for
g' m RS >> 1
EE 214 Lecture 10
11
CG Input Impedance (1)
KCL @ vo : 0 =
vo vo vtest
+ −
− g' m vtest
RL ro
ro
⇒ vo ≅ g' m (RL || ro )vtest
KCL @ vtest : itest = g' m vtest +
⇒ Yin =
vtest vo
−
ro
ro
itest
g' r
≅ m o + sCS
vtest RL + ro
v gs = −vtest
B. Murmann
EE 214 Lecture 10
12
124
CG Input Impedance (2)
Yin ≅
•
(R + r ) ⎞
g' m ro ⎛
⎜⎜1 + sCS L o ⎟⎟
RL + ro ⎝
g' m ro ⎠
At low frequencies
Rin =
•
1
1 ⎛ RL ⎞
⎜1 +
⎟
≅
Yin g' m ⎜⎝
ro ⎟⎠
Two interesting cases
– RL<<ro:
Rin ≅
1
g' m
(well known)
– RL>>ro:
Rin ≅
RL
g' m ro
(not so well known…)
B. Murmann
EE 214 Lecture 10
13
CG Output Impedance
itest =
vtest v gs
+
+ g' m vgs
ro
ro
v gs = −itest RS
Rout =
vtest
≅ ro (1 + g' m RS )
itest
(Very high if g'mRS>>1 !)
B. Murmann
EE 214 Lecture 10
14
125
CG Summary
•
Current gain is unity up to very high frequencies
– Our "simple" device model predicts up to roughly fT
•
Input impedance is very low
– At least when the output is also terminated with some
reasonable impedance
•
Can achieve very high output resistance
•
In summary, a common gate stage is ideal for turning a decent
current source into a much better one
– Seems like this is something we can use to improve our
common source stage
• Which is indeed nothing but a decent (voltage controlled)
current source
B. Murmann
EE 214 Lecture 10
15
Cascode Stage
•
Invented ~1920, in the context of vacuum tube circuits
– http://web.mit.edu/klund/www/cascode.html
Gm = g m1 ⋅
i0
≅ g m1
ii
Ro ≅ ro 2 (1 + g' m 2 ro1 )
Gm Ro = g m1ro 2 (1 + g' m 2 ro1 ) ≅ g m1ro1 ⋅ g' m 2 ro 2 ~ ( g m ro )
2
B. Murmann
EE 214 Lecture 10
16
126
High Frequency Benefits
vx
g ⎛ R ⎞
= g m1Z x ≅ m1 ⎜⎜1 + L ⎟⎟
vi
g' m 2 ⎝ ro 2 ⎠
B. Murmann
•
Very close to 1, for moderate values of RL
– Mitigates Miller effect
– Even if RL is large, there is often a load
capacitance that provides a low impedance
termination to help maintain this feature
•
Additional benefit
– Cascode mitigates direct forward coupling
from Vi to Vo at high frequencies
EE 214 Lecture 10
17
High Frequency Issues
io
≅
ii
•
1+ s
1
C gs + Csb
g' m
Cascode causes pole around fT
– Usually non-dominant
– Can be a headache for
stability/phase margin in
circuits with feedback
• More later…
B. Murmann
EE 214 Lecture 10
18
127
Design Example 2 (Lecture 6) Revisited
•
B. Murmann
What we expect to see in Spice
after adding the cascode device
– Bandwidth should increase
(reduction of Miller effect)
– Non-dominant pole around
some fraction of fT of cascode
device
EE 214 Lecture 10
19
Frequency Response (1)
15
without cascode
with cascode
|vo/vi| [dB]
10
5
0 0
10
10
1
10
2
10
3
f [MHz]
•
Bandwidth increased from 184MHz to 248MHz
B. Murmann
EE 214 Lecture 10
20
128
Frequency Response (2)
without cascode
with cascode
10
0
|vo/vi| [dB]
-10
-20
-30
-40
-50
-60 0
10
10
1
10
2
10
3
10
4
10
5
f [MHz]
B. Murmann
EE 214 Lecture 10
21
Non-Dominant Pole Estimate
**** mosfets
element 0:mn1
0:mnc
model
0:nch214
0:nch214
region
Saturati
Saturati
id
502.6426u 502.6426u
vgs
806.0164m 962.3326m
vds
837.6674m 659.6900m
vbs
0.
-837.6674m
vth
583.4970m 753.1734m
gam eff 894.1238m 938.7402m
gm
4.0020m
4.0280m
gds
82.8309u 114.8096u
gmb
896.5590u 604.0200u
cdtot
48.7686f
42.9326f
cgtot
56.5608f
55.8554f
cstot
108.9121f
87.3547f
cbtot
118.5039f
89.2450f
cgs
39.4622f
40.6097f
cgd
7.2418f
7.0879f
B. Murmann
f p2 ≅
≅
g' m
1
2π C gsc + Csbc + Cdb1
1
4 mS + 0.6 mS
2π 40.6 fF + (87.4 fF − 40.6 fF ) + (48.8 fF − 7.2 fF )
f p 2 ≅ 5.7 GHz
For comparison:
1 4 mS
2π 56 fF
≅ 11.4GHz
fT 2 ≅
EE 214 Lecture 10
22
129
Supply Headroom Issue
B. Murmann
•
Even if we adjust VB such
that VDS1 is small, adding
a cascode reduces the
available signal swing
•
This can be a big issue
when designing circuits
with VDD≅1V
– Typically need each
VDS>~0.2V
– A severe dynamic
range penalty
EE 214 Lecture 10
23
Cascode Noise
RL
RL
•
A closer look reveals that
cascodes can contribute
significant noise at high
frequencies
– Noise current A no
longer compensates B
at high frequencies
•
We'll take a more
quantitative look at this
later in this course
Vo
Vo
VB
VB
i2n2
i2n1
B. Murmann
It is typically argued that
cascodes do not add a
significant amount of noise
VDD
VDD
Vi
•
B
A
Vi
i2n1
EE 214 Lecture 10
24
130
Lecture 11
Common Drain Stage
(Source Follower)
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 11
1
Overview
•
Reading
– 3.3.7 (Common Drain Stage)
– 7.2.2 (Frequency Response of Voltage Buffers)
– 5.3 (Source Follower as an Output Stage, optional)
•
Introduction
– Last lecture, we have seen that a common gate stage has a
fairly low input impedance, and high output impedance. The
common drain stage that we'll analyze today exhibits the
exact opposite features: High input impedance and low
output impedance. After an analysis of relevant port
characteristics, we will discuss some potential applications
and also drawbacks of this circuit.
B. Murmann
EE 214 Lecture 11
2
131
Common Drain Stage
B. Murmann
EE 214 Lecture 11
3
CD Voltage Transfer (1)
⎛
1
vo ⎜⎜ sC Ltot + sC gs +
RLtot
⎝
⎞
⎟⎟ − vi sC gs − g m (vi − vo ) = 0
⎠
g m + sC gs
vo
=
vi g + sC + sC + 1
m
gs
Ltot
RLtot
1+
C Ltot = C L + Csb
B. Murmann
RLtot = RL ||
1
|| ro
g mb
sC gs
vo
gm
gm
=
⋅
(
1
s
C
vi g +
gs + C Ltot )
1+
m
1
RLtot
gm +
RLtot
EE 214 Lecture 11
4
132
Low Frequency Gain
av 0 =
•
gm
RLtot = RL ||
1
gm +
RLtot
1
|| ro
g mb
Interesting cases
– RL→∞, ro→∞, gmb=0
av 0 = 1
• PMOS, source tied to body, ideal current source
av 0 =
– RL→∞, ro→∞, gmb≠0
• NMOS, ideal current source)
(typically ≅ 0.8)
– ro→∞, gmb=0, RL finite
av 0 =
• PMOS, source tied to body, load resistor
B. Murmann
gm
g m + g mb
gm
gm +
EE 214 Lecture 11
1
RL
5
High Frequency Gain
s
v
z
av (s ) = o = av 0 ⋅
s
vi
1−
p
1−
•
g
z=− m
C gs
1
RLtot
p=−
C gs + C Ltot
gm +
Three scenarios
|z|<|p|
|z|>|p|
|z|=|p|
(infinite bandwidth !?)
B. Murmann
EE 214 Lecture 11
6
133
CD Input Impedance
•
By inspection
Yin = s (C gd + C gb ) + sC gs (1 − av ( s ))
B. Murmann
•
Gain term av(s) is real and close to
unity up to fairly high frequencies
•
Hence, up to moderate
frequencies, we see a capacitor
looking into the input
– A fairly small one, Cgd + Cgb,
plus a fraction of Cgs
EE 214 Lecture 11
7
PMOS Stage with Body-Source Tie
•
Gate-body capacitance is in
parallel with Cgs
•
gmb generator inactive
– Low frequency gain very
close to unity
•
Very small input capacitance
Yin = sC gd + s (C gs + C gb )(1 − av ( s ))
Yin ≅ sC gd
B. Murmann
EE 214 Lecture 11
8
134
Bootstrapped PMOS Stage
•
"Extremely" small input
capacitance
Yin ≅ sC gd (1 − avP ( s )avN ( s ))
B. Murmann
EE 214 Lecture 11
9
CD Output Impedance (1)
•
Let's first look at an analytically
simple case
– Input driven by ideal voltage
source
•
By inspection
Z out =
•
B. Murmann
1
1
g m + g mb s (C gs + Csb )
Low output impedance
– Resistive up to very high
frequencies
EE 214 Lecture 11
10
135
CD Output Impedance (2)
•
Now include finite source resistance
Zx =
vo
ix
ix = (vo − v g )(g m + sC gs )
⎛ vg
= vo ⎜⎜1 −
⎝ vo
B. Murmann
⎞
⎟⎟(g m + sC gs )
⎠
vg
vo
=
Ri
1
+ Ri
sC gs
EE 214 Lecture 11
11
CD Output Impedance (3)
Zx ≅
•
Two interesting cases
1 (1 + sR i C gs )
sC gs ⎞
gm ⎛
⎜⎜ 1 +
⎟
g m ⎟⎠
⎝
Ri < 1/gm
Ri > 1/gm
Inductive behavior!
B. Murmann
EE 214 Lecture 11
12
136
Equivalent Circuit for Ri > 1/gm
R1 || R 2 =
1
gm
R 2 = Ri
L=
•
Ri2 C gs
g m Ri − 1
This circuit is prone to ringing!
– L forms an LC tank with any capacitance at the output
B. Murmann
EE 214 Lecture 11
13
Inclusion of Parasitic Input Capacitance
•
What happens to this result if we don’t neglect Ci=Cgd+Cgb?
Zx =
1
Ri {C gs + C i }
B. Murmann
<
1 (1 + sR i {C gs + C i })
sC gs ⎞
gm ⎛
⎜⎜ 1 +
⎟ (1 + sR i C i )
g m ⎟⎠
⎝
gm
1
<
C gs Ri C i
EE 214 Lecture 11
1
Ri {C gs + C i }
<
g
1
< m
Ri C i C gs
14
137
Application 1: Level Shifter
•
Output quiescent point is roughly Vt+Vov lower than input
quiescent point
B. Murmann
EE 214 Lecture 11
15
Application 2: Buffer
•
Low frequency voltage gain of the above circuit is ~gmRbig
– Would be ~gm(Rsmall||Rbig) without CD buffer stage
B. Murmann
EE 214 Lecture 11
16
138
Issues
•
Several sources of nonlinearity
– Vt is a function of Vo (NMOS, without S to B connection)
– ID and thus Vov changes with Vo
• Gets worse with small RL
•
Reduced input and output voltage swing
– Consider e.g. VDD=1V, Vt=0.3V, VOV=0.2V
• CD buffer stage consumes 50% of supply headroom!
– In low VDD applications that require large output swing, using
a CD buffer is often not possible
– CD buffers are more frequently used when the required
swing is small
• E.g. pre-amplifiers or LNAs that turn μV into mV at the output
B. Murmann
EE 214 Lecture 11
17
Application 3: Load Device
•
Advantages compared to resistor load
– "Ratiometric"
• Gain depends on ratio of similar
parameters
• Reduced process and
temperature variations
– First order cancellation of
nonlinearities
•
av 0 =
B. Murmann
g m1
g m 2 + g mb 2
Disadvantage
– Reduced swing
EE 214 Lecture 11
18
139
Summary – Elementary Transistor Stages
•
Common source
– VCCS, makes a good voltage amplifier when terminated with
a high impedance
•
Common gate
– Typically low input impedance, high output impedance
– Can be used to improve the intrinsic voltage gain of a
common source stage
• "Cascode" stage
•
Common drain
– Typically high input impedance, low output impedance
– Great for shifting the DC operating point of signals
– Useful as a voltage buffer when swing and nonlinearity are
not an issue
B. Murmann
EE 214 Lecture 11
19
Practical Biasing Issues (1)
B. Murmann
EE 214 Lecture 11
•
Typically cannot
afford to have
sufficiently large
decoupling cap for VB
•
Also, may want to
work with ground
referenced input
voltage
20
140
Practical Biasing Issues (2)
•
The modified circuit on the left
solves allows ground referenced
inputs
– But now there is a high pass
filter in the signal path
– Need large area to achieve
low corner frequency
– Typically, this approach is
practical only in RF circuits
• Why?
B. Murmann
EE 214 Lecture 11
21
Better Solution: Differential Amplifier
B. Murmann
•
Bias point of Vop and Vom set
by IB
– Independent of quiescent
value of Vip and Vim
•
Differential output (Vop-Vom)
depends only on differential
input (Vip-Vim)
•
More next lecture…
EE 214 Lecture 11
22
141
Lecture 12
Differential Pair
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 12
1
Overview
•
Reading
– 3.5.0, 3.5.3, 3.5.5 (Differential Pair)
•
Introduction
– The differential pair is the most widely used two-transistor
sub-circuit in analog ICs. Using a differential pair as a
replacement for a simple common source stage eliminates
the need for cumbersome gate biasing. In addition, its
differential input and output voltages are more immune to
parasitic signal coupling. Today we will analyze some
properties of differential pairs and introduce the notion of
common- and differential-mode signal components.
B. Murmann
EE 214 Lecture 12
2
142
Differential Pair
Id1 Id2
Vip
Vim
ITAIL
•
When Vip=Vim, and both
transistors are identical, we
must have Id1=Id2=ITAIL/2
•
How about Vip=Vim=1V versus
Vip=Vim=2V?
– Makes no difference!
•
From a signal perspective, we
care only about the difference
of the applied voltages
– Makes sense to introduce
a new variable
• Vid=(Vip-Vim)
B. Murmann
EE 214 Lecture 12
3
Differential and Common Mode (1)
Id1 Id2
ITAIL
+
Vip
-
B. Murmann
Vid
+
Vim
-
•
We now still need a second
variable that describes the
potential of nodes Vip and Vim
with respect to GND
– Could choose either Vip or Vim
•
More elegant solution
– Cut Vid in half and define a
new independent variable
– "Common mode" voltage Vic
EE 214 Lecture 12
4
143
Differential and Common Mode (2)
Id1 Id2
Vid = Vip − Vim
Vid
2
V
Vim = Vic − id
2
Vip = Vic +
ITAIL
+
Vip
-
Vid/2
Vid/2
Vic
B. Murmann
⇒ Vic =
+
Vim
-
Vip + Vim
2
EE 214 Lecture 12
5
Coupling Noise Immunity
•
Using differential pairs/signals not only solves the biasing issue,
but also mitigates coupling noise issues
Single Ended Signaling
B. Murmann
Differential Signaling
EE 214 Lecture 12
6
144
Large Signal Transfer Function (1)
•
Id1 Id2
Let's first do a simple analysis using
long channel equations
Vip − Vgs1 = Vim − Vgs 2
I d 1 + I d 2 = I TAIL
M1 M2
ITAIL
Vgs1 = Vt +
+
Vip
-
Vid/2
Vid/2
Vic
B. Murmann
+
Vim
-
2I d1
W
μCox
L
⇒ I od = I d 1 − I d 2 =
Vgs 2 = Vt +
1
W
μCox Vid
2
L
2I d 2
W
μCox
L
4 ITAIL
− Vid 2
W
μCox
L
EE 214 Lecture 12
7
Large Signal Transfer Function (2)
•
We can turn this into a more elegant expression by using
I TAIL 1
W
2
= μCox VOV
L
2
2
where VOV is the quiescent point gate overdrive with Vid=0
⎛ V
I
I −I
V
⇒ od = d 1 d 2 = id 1 − ⎜⎜ id
I TAIL
I TAIL
VOV
⎝ 2VOV
•
⎞
⎟⎟
⎠
2
This equation predicts
– Iod/ITAIL = 0 when Vid=0, as expected
– Complete current steering (Iod/ITAIL=±1) takes place when
Vid= ±VOV√2
B. Murmann
EE 214 Lecture 12
8
145
Large Signal Plot
•
Note that the equation on previous slide is only valid in the
center of this transfer function (between saturation
⎝ points)
– Why?
1
Slope = 1
Iod/ITAIL 0
1
2
1
0
Vid/VOV
-√2
B. Murmann
1
2
√2
EE 214 Lecture 12
9
Observations
•
Looks like something we have seen before
– A transfer function that is somewhat linear as long as
Vid<<VOV
•
For small signal analysis, we can find an equivalent
transconductance by differentiation at the operating point
Gm =
•
dI od
dVid
=
Vid =0
I TAIL
VOV
Note that the transconductance of M1 and M2 is given by
g m1,2
B. Murmann
I
2 TAIL
2I D
2 = I TAIL
=
=
VOV
VOV
VOV
EE 214 Lecture 12
10
146
Does the Tail Node Move?
Id1 Id2
•
Can show that
Vx
Vx = Vic − Vt − VOV
ITAIL
•
+
Vip
-
Vid/2
Vid/2
Vic
+
Vim
-
B. Murmann
1⎛ V
1 − ⎜⎜ id
4 ⎝ VOV
⎞
⎟⎟
⎠
2
From this expression, we see that
from a small signal perspective the tail
node is pinned at Vic-Vt-VOV
– "AC ground"
EE 214 Lecture 12
11
Small Signal Equivalent
id1
gmvid/2
id2
iod = id 1 − id 2 = g m vid
iod/2
B. Murmann
Sufficient to work with half
circuit!
– Can directly apply
everything we've
learned about single
transistor stage
•
Half circuit caveats
– Can not analyze
nonlinearity using half
circuits
– Assumes that M1 and
M2 are identical
-gmvid/2
Vx
gmvid/2
•
iod
gmvid
EE 214 Lecture 12
12
147
Differential Voltage Amplifier Example
B. Murmann
EE 214 Lecture 12
13
Short Channel Effects
Id1 Id2
•
Can model short channel effects using
long channels with source degeneration
Vip − Vgs1 − I d 1 Rsx = Vim − Vgs 2 − I d 2 Rsx
Rsx
ITAIL
Rsx =
B. Murmann
I d 1 + I d 2 = I TAIL
Rsx
Vgs1 = Vt +
1 VOV
gm ε c L
2I d1
W
μCox
L
Vgs 2 = Vt +
2I d 2
W
μCox
L
⇒ big mess...
EE 214 Lecture 12
14
148
Nonlinearity Comparison (1)
•
Can look at this using Taylor expansions
•
Without Rsx (long channel), we can show that
⎛V
I od
≅ ⎜⎜ id
I TAIL ⎝ VOV
•
⎞ 1 ⎛ Vid
⎟⎟ − ⎜⎜
⎠ 8 ⎝ VOV
3
⎞
⎟⎟ − ...
⎠
With Rsx (short channel) this becomes
⎛
I od
Vid
≅ ⎜⎜
I TAIL ⎝ (1 + g m Rsx ) ⋅VOV
⎞ 1⎛
⎞⎛
Vid
1
⎟⎟ − ⎜⎜
⎟⎟⎜⎜
⎠ 8 ⎝ 1 + g m Rsx ⎠⎝ (1 + g m Rsx ) ⋅ VOV
3
⎞
⎟⎟ − ...
⎠
•
Short channel differential pair is "more linear"
– But also has less gm at a given current
•
For fairness, let's compare nonlinearity at same gm/ID
B. Murmann
EE 214 Lecture 12
15
Nonlinearity Comparison (2)
•
Recall that
•
Plugging this into the equations on the previous slide yields
gm
1
2
≅
I D 1 + g m Rsx VOV
3
⎞
⎞ 1 ⎛ 1 gm
⎛ 1 gm
I od
Vid ⎟⎟ − ⎜⎜
Vid ⎟⎟ − ...
≅ ⎜⎜
I TAIL ⎝ 2 I D
⎠
⎠ 8 ⎝ 2 ID
Long Channel
3
⎞⎛ 1 g m
⎞
⎞ 1⎛
⎛ 1 gm
I od
1
⎟⎟⎜⎜
Vid ⎟⎟ − ⎜⎜
Vid ⎟⎟ − ...
≅ ⎜⎜
I TAIL ⎝ 2 I D
⎠
⎠ 8 ⎝ 1 + g m Rsx ⎠⎝ 2 I D
•
Short Channel
Minor win for short channel in terms of linearity
– In practice, this may be overshadowed by additional effects
that are not captured in the simple degeneration model
B. Murmann
EE 214 Lecture 12
16
149
"Linear Region" of Transfer Function
•
In the long channel case, we found that for good linearity we
need Vid<<VOV
•
As we have seen from the derivation above, this requirement
can be generalized to capture both long and short channel
cases simultaneously
2
Vid <<
•
⎛ gm ⎞
⎜⎜ ⎟⎟
⎝ ID ⎠
A few approximate rules of thumb
For ~1.5% nonlinearity error:
Vid < 0.5 ⋅
2
⎛ gm ⎞
⎜⎜ ⎟⎟
⎝ ID ⎠
B. Murmann
For ~0.1% nonlinearity error:
Vid < 0.2 ⋅
2
⎛ gm ⎞
⎜⎜ ⎟⎟
⎝ ID ⎠
EE 214 Lecture 12
17
Voltage Amplifier Transfer Functions
•
In a differential amplifier, we primarily
want to have large gain that links only
the two differential variables
Adm =
•
vod
vid
Unfortunately, circuit nonidealities will
also cause nonzero "parasitic" gain
terms
Acm =
voc
vic
Adm −cm =
voc
vid
Acm− dm =
vod
vic
See text 3.5.6.9
B. Murmann
EE 214 Lecture 12
18
150
Common Mode Gain
Acm =
•
Ideally zero (RTAIL=∞)
•
With finite RTAIL:
Acm = −
B. Murmann
voc
vic
gm
⋅ (R ro [1 + g m ⋅ 2 RTAIL ])
1 + g m ⋅ 2 RTAIL
EE 214 Lecture 12
19
Common Mode Rejection Ratio
•
Figure of merit that quantifies ratio of desired/undesired gain
– Ideally infinite
CMRR =
•
For our simple resistively loaded differential pair, this becomes
(assuming R<<ro and ignoring body effect)
CMRR ≅
•
Adm
Acm
gm ⋅ R
= 1 + g m ⋅ 2 RTAIL
gm
⋅R
1 + g m ⋅ 2 RTAIL
Other important figures of merit
Adm
Acm− dm
B. Murmann
Adm
Adm −cm
EE 214 Lecture 12
20
151
Power Supply Rejection Ratio
•
In practice, "noise" on the supplies
will also propagate to the output
– In a differential system usually
due to (half-) circuit imbalance
•
Define
A+ =
vod
vdd
PSRR+ =
B. Murmann
A− =
Adm
A+
vod
vss
PSRR− =
Adm
A−
EE 214 Lecture 12
21
Input Referred Interpretation
vdd
vss
PSRR+ PSRR−
vi or vid
VDD
Amplifier
vo or vod
•
E.g. 1mV input signal, 100mV supply noise
– Need PSRR >> 100 (40dB)
•
PSRR can be a very critical issue in highly integrated, complex
integrated circuits
– Lots of potential supply noise sources
• E.g. cross-talk between analog and digital sections
B. Murmann
EE 214 Lecture 12
22
152
Lecture 13
Current Mirrors
Offset Voltage
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 13
1
Overview
•
Reading
– 4.1, 4.2 (Current Mirrors)
– 3.5.6.6, 3.5.6.7 (Input Offset Voltage)
•
References
– M. Pelgrom et al., "Matching properties of MOS transistors," IEEE
J. Solid-State Circuits, Oct. 1989.
– P.G. Drennan et al., "Understanding MOSFET mismatch for analog
design," IEEE J. Solid-State Circuits, March 2003.
•
Introduction
– In this lecture, we will take a closer look at practical
implementations of current mirrors with emphasis on high
swing biasing techniques. In addition, we will discuss
nonidealities such as threshold voltage mismatch and IR
drop.
B. Murmann
EE 214 Lecture 13
2
153
Current Mirror Bias
•
Objectives
– Want "accurate" mirror ratio ITAIL/IREF
– Want large RTAIL (and small CTAIL) for good CMRR
– Want small Vmin to maximize common mode input range
B. Murmann
EE 214 Lecture 13
3
Basic Sizing Considerations
•
Always use L1=L2
•
Typically make W1/W2 or W2/W1 integer
– Use unit devices connected in parallel
– "m-factor" in Spice
• E.g. M1 d g s b W=10u L=0.35u m=5
B. Murmann
EE 214 Lecture 13
4
154
Inaccuracy due to ΔVDS
ID
=1/r o
slope
ΔI
VDS
ΔV
ΔI = I1 − I 2 ≅
•
V1 − V2 ΔV
=
ro
ro
Two options
– Use device with large ro (Large L)
– Make V1 as close as possible to V2
B. Murmann
EE 214 Lecture 13
5
Output Resistance (M2)
2500
NMOS W/L=10, VGS=800mV, L=0.35μm...0.7μm
2000
ro [kΩ]
1500
1000
500
0
0
B. Murmann
0.5
1
1.5
2
VDS [V]
EE 214 Lecture 13
2.5
3
3.5
6
155
Output Resistance Zoom
700
NMOS W/L=10, VGS=800mV, L=0.35μm...0.7μm
600
ro [kΩ]
500
400
300
200
100
0
0
B. Murmann
0.1
0.2
0.3
0.4
VDS [V]
0.5
0.6
0.7
EE 214 Lecture 13
7
Higher Rout
Rout ≅ g m ro2
•
A cascode can help create a higher output resistance, e.g. to
improve CMRR in a differential pair
•
Even though the impedance is now high at the current mirror
output, we still need V1=V2 to minimize systematic errors in the
current ratio I2/I1
B. Murmann
EE 214 Lecture 13
8
156
Solution 1
•
Works, but VOUTmin is very large
– Using long channel algebra,
we have
VOUT min ≅ 2(Vt + VOV ) − Vt = Vt + 2VOV
•
B. Murmann
Note that using long channel
equations tends to be OK for
current mirrors
– Usually operated in strong
inversion
– Channel length often not
minimum
EE 214 Lecture 13
9
Solution 2
•
Use some kind of "magic battery" that sets the cascode gate
potential such that VOUTmin = 2VOV (minimum possible)
•
"High swing" bias
B. Murmann
EE 214 Lecture 13
10
157
Magic Battery 1
•
Gate overdrive of ¼ device is twice as large as VOV of all other
devices
– Vcas=Vt+2VOV
– VOUTmin = 2VOV
•
Sensitive to body effect
B. Murmann
EE 214 Lecture 13
11
Magic Battery 2
•
Less sensitive to body effect
•
Stacked device mimics transistor with length 4L
– May not match characteristics of devices with length L all
that well
B. Murmann
EE 214 Lecture 13
12
158
Magic Battery 3
I1 ≅
V ⎞
W 2
1
1W ⎛
= μCox
μCox VOV
⎜ [Vx + VOV + Vt ] − Vt − x ⎟ ⋅ Vx
L
2
3 L⎝
2⎠
⇒ Vx = VOV
•
Insensitive to body effect
•
(1/3) triode device will not exactly produce VOV
– Typically make device ~1/5 of W, to allow for safety margins
and improved Rout
B. Murmann
EE 214 Lecture 13
13
Magic Battery 4
B. Murmann
EE 214 Lecture 13
•
No extra current branch
•
Needs lots of headroom on
input side
•
Sensitive to body effect
14
159
Magic Battery 5
I1
I1
W1/L1
W2/L2
I1
I2
Vcas
W2/L2
W2/L2
Vx
(2/3)*W1/L1
W1/L1
•
The previous circuits assumed that Lmirror=Lcascode
– Sometimes want Lmirror≠Lcascode
•
The above circuit makes Vcas =Vt2 + VOV2 + VOV1
– VOUTmin becomes VOV2 + VOV1, as desired
B. Murmann
W1/L1
EE 214 Lecture 13
15
Design Considerations
•
In all of the above-discussed circuits, care must be taken to keep
the devices in their active region with sufficient safety margin, and
not to "waste" too much ro
– Typically use ~1/6 in 1/4 size device approach (slide 10)
– Typically use ~1/5 in 1/3 triode device approach (slides 12-14)
•
Work with integer ratios and unit devices as much as possible
– Using a 1/5 device really means that we work with a unit
device of size 1, and use 5 of them elsewhere
•
Sometimes desirable to keep mirror ratio (I2/I1) reasonably small
– Say, no larger than 10…20
– Ensures reasonable bandwidth at bias nodes for fast recovery
from noise coupling
B. Murmann
EE 214 Lecture 13
16
160
Capacitive Coupling
Vx
Low cap:
fast recovery
big bounce
t
Vx
High cap:
slow recovery
small bounce
t
•
Can use decoupling capacitors to reduce the amplitude of noise
coupling into bias nodes
•
If noise is "deterministic" and occurs at the right point in time,
you might be better off not decoupling, but making the bias node
"fast" (small mirror ratio) so it can recover quickly
B. Murmann
EE 214 Lecture 13
17
Offset Voltage
•
A current mirror relies on accurate matching of device
parameters (e.g. Vt)
•
Unfortunately, there will always be some mismatch between two
nominally identical devices
•
A common way of modeling mismatch is based on a
decomposition into two components
– A difference in threshold voltage
– A difference in μCoxW/L = β
•
The differences are typically modeled as zero-mean Gaussian
random variables with a standard deviation that primarily
depends on the device area (W·L)
B. Murmann
EE 214 Lecture 13
18
161
Pelgrom Coefficients
σ ΔV =
t
AVt
σ Δβ =
WL
β
Aβ
WL
•
In 0.35μm technology: AVt ≅ 7mV-μm, Aβ ≅ 1%-μm
– AVt tends to scale down with technology, roughly proportional
to gate oxide thickness
– Aβ has remained roughly constant with scaling
•
Example: W=10μm, L=0.35μm
σ ΔV =
t
7 mV
= 3.7 mV
3.5
B. Murmann
σ Δβ =
β
1%
= 0.53%
3.5
EE 214 Lecture 13
19
Inaccuracy Due to Mismatch
ΔI = I 1 − I 2 ≅ g m ΔVt + I 1Δβ
ΔI
I1
•
gm
ΔVt + Δβ
I1
Example: W=10μm, L=0.35μm, gm/ID=10S/A
⎛
⎝
σ ΔI = ⎜ 10
I1
•
≅
2
S
⎞
⋅ 3.7 mV ⎟ + (0.53% )2 =
A
⎠
(3.7% )2 + (0.53% )2
= 3.74%
Threshold mismatch usually dominates
B. Murmann
EE 214 Lecture 13
20
162
Well Proximity Effects
•
P. G. Drennan et al., "Implications of Proximity Effects for
Analog Design," Proc. CICC, pp.169-176, Sep. 2006.
B. Murmann
EE 214 Lecture 13
21
Inaccuracy Due to IR Drop
ΔI = I 1 − I 2 ≅ g mVwire
ΔI
I1
•
≅
gm
Vwire
I1
Want small gm/ID ("large VOV") to mitigate errors due to wire IR
drop
– Unfortunately this means large Vmin
B. Murmann
EE 214 Lecture 13
22
163
A Note on Current Mirror Accuracy
•
As we have seen, it is very hard to build "highly" accurate
current mirrors
– Systematic errors: ΔVDS, IR Drop, …
– Random errors: ΔVt, …
•
This is often not an issue
– E.g. gm ~ sqrt(ID); 20% error in ID causes ~10% error in gm
•
For mirrors using short channels, it is sometimes OK to tweak
the mirror ratio to compensate for ΔVDS-induced systematic
errors
– Keep in mind that you should not overdo this; it makes no
sense to tweak a mirror in Spice to unrealistic accuracies
– Tweaking is only going to work in practice if you know that
gds is properly modeled in your technology file
• Not always the case
B. Murmann
EE 214 Lecture 13
23
Current Distribution (1)
•
Typically, we'll only have one single reference current generator
on a chip
•
Can generate/distribute currents across chip in two different ways
– Distribute gate voltage
• Can cause big problems due to IR drop and process gradients
• Usually limited to local distribution
– Distribute currents
• Have one global bias cell close to reference that sends currents
into local biasing sub-circuits
• Disadvantage: consumes additional current
B. Murmann
EE 214 Lecture 13
24
164
Current Distribution (2)
Iref
B. Murmann
EE 214 Lecture 13
25
165
Lecture 14
Process Variations
Feedback
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 14
1
Overview
•
Reading
– 8.0, 8.1, 8.2, 8.3 (Feedback)
– 9.2 (Relation Between Gain and Bandwidth)
– Supplementary Handout "Feedback Systems" by Tom Lee
(see web, optional)
•
Introduction
– In today's lecture we will first take a look at typical
component variations in CMOS technology. In light of these
numbers, we then consider negative feedback as a tool for
building amplifiers with precisely defined gain. As we shall
see, using a forward amplifier with "high", but arbitrary gain,
combined with a ratiometric, passive feedback network
allows us to build drift and process insensitive precision gain
stages.
B. Murmann
EE 214 Lecture 14
2
166
PVT
•
So far, we've assumed that our Spice model accurately predicts
the performance of every single chip we make
•
We have also assumed that all our circuits run at room
temperature, and VDD is precisely fixed
•
In practice, it is the circuit designer's job to ensure that the
circuit works in presence of large variations
– PROCESS: Variations among production lots
• "Slow, Nominal and Fast" corners
• Sometimes there are even significant variations across wafers
and individual chips
– VOLTAGE: VDD is usually specified only within ±10%
• E.g. VDD= 2.7…3.3V in our technology
– TEMPERATURE: Ambient temperature variations
• 0…70°C (or -40…125°C)
B. Murmann
EE 214 Lecture 14
3
Graphical Interpretation
[Razavi, p. 599]
B. Murmann
EE 214 Lecture 14
4
167
Typical Variations
Parameter
Lot-to-Lot
Temperature
Coefficient
Device Matching
(Area Dependent)
Vt
±100mV
-2mV/°C
~1…10mV
μCox
±20%
-0.33%/°C
±0.2%
Rpoly2 (50Ω/square)
±20%
0.2%/°C
±0.1%
Rnwell (1k Ω /square)
±40%
1%/°C
±1%
Cpoly-poly2
±15%
-30ppm/°C
±0.1%
B. Murmann
EE 214 Lecture 14
5
Consequences
•
Performance metrics that depend on absolute component
values will show large variations
•
Sometimes this is OK
– E.g. bandwidth of a circuit
• Can overdesign, and make sure to have minimum required
bandwidth in presence of worst case variations
•
Sometimes this is not OK
– E.g. need a precise gain of two in an A/D converter
• gmRL or gmro will never be accurate enough
•
Solution: Use negative feedback to desensitize the circuit to
gmRL or gmro variations
B. Murmann
EE 214 Lecture 14
6
168
Negative Feedback
•
Harold S. Black, 1927
vout = a(vin − fvout )
vout
a
=
vin 1 + af
•
Interesting case:
af >> 1
B. Murmann
⇒
vout 1
≅
vin
f
EE 214 Lecture 14
7
Interpretation
•
As long as we have "large" gain in the forward path a, the
overall gain will depend only on f
•
Since vout/vin ~1/f, we often make f ≤ 1
– E.g. for a "closed loop" gain of two, we need f=0.5
•
f ≤ 1 is easy to implement, and ratiometric!
– A "wire" (f=1) or resistive or capacitive voltage divider
vout R1 + R2
≅
vin
R1
vout
≅1
vin
B. Murmann
EE 214 Lecture 14
8
169
Example
R1 = R2
•
⇒ f = 0.5
Case 1: a=100
vout
1
1
=
=
= 1.96078...
1
1
vin
+f
+ 0.5
a
100
•
Case 2: a=1000
vout
1
1
=
=
= 1.996007...
1
1
vin
+f
+ 0.5
a
1000
•
10x variation in forward gain, and only about 1.8% change in
closed loop gain!
B. Murmann
EE 214 Lecture 14
9
Gain Sensitivity
A=
a
1 + af
•
Define
•
Can show that
•
Fractional change in gain is reduced by product of a and f,
which can be made arbitrarily large (conceptually)
– Loop gain T
•
We will find that loop gain is a very meaningful parameter that
will also appear in bandwidth and impedance calculations
– More later…
B. Murmann
Δa
Δa
ΔA
= a = a
A 1 + af 1 + T
EE 214 Lecture 14
10
170
Effect of Negative Feedback on Nonlinearity
•
(1)
vout = a1 (vin − fvout ) + a3 (vin − fvout )
(2)
vout = b1vin + b3vin3
3
!
Substitute (2) into (1), then compare coefficients to get
b1 =
a1
1 + a1 f
b3 =
a3
(1 + a1 f )4
•
Linear term as expected, reduced by (1+T)
•
Cubic term reduced by (1+T)4!
B. Murmann
EE 214 Lecture 14
11
Negative Feedback and Bandwidth
E .g .
a( s ) =
a0
1−
•
s
p1
Closed loop transfer function:
A( s ) =
a0
1
a( s )
=
⋅
s
1
1 + a( s ) f 1 + a0 f 1 −
p1 1 + a0 f
•
Bandwidth increases by (1+T)!
– But gain is reduced by (1+T)
•
Product of gain and bandwidth remains constant
B. Murmann
EE 214 Lecture 14
12
171
Bode Plot Illustration
B. Murmann
EE 214 Lecture 14
13
Early Obstacles
•
Today we are taking the concept of negative feedback for
granted
•
At the time of his "invention," Harold Black had a hard time
convincing his colleagues that negative feedback was indeed
something useful
– It was very hard to make a "high gain" forward amplifier
using vacuum tubes
– Why have large gain and then throw it away by applying
negative feedback?
•
Long before negative feedback had been deemed useful,
positive feedback was routinely applied to increase the gain of
amplifiers
B. Murmann
EE 214 Lecture 14
14
172
Positive Feedback
•
Edwin H. Armstrong, 1915
vout = a(vin + fvout )
vout
a
=
vin 1 − af
•
Example
•
Tenfold increase in gain!
af = 0.9
B. Murmann
⇒
vout
a
=
= 10a
vin 1 − 0.9
EE 214 Lecture 14
15
Feedback Using Ideal OpAmp
(
vout = g ⋅ v + − v −
)
f =
R1
R1 + R2
a=g
vout
a
g
R + R2
=
=
≅ 1
vin 1 + af 1 + g R1
R1
R1 + R2
B. Murmann
EE 214 Lecture 14
16
173
Inverting Configuration
(
vout = g ⋅ v + − v −
)
f =?
•
a =?
Circuit does not map directly into generic block diagram
– Cannot directly identify a and f
– a≠g
B. Murmann
EE 214 Lecture 14
17
Superposition
⇒a
⇒ − af
B. Murmann
EE 214 Lecture 14
18
174
Inverting Configuration
a=−
R2
⋅g
R1 + R2
− af = −
B. Murmann
R1
⋅g
R1 + R2
EE 214 Lecture 14
19
Result
R2
a=−
⋅g
R1 + R2
R1
⋅g
af
R
R1 + R2
f =
=
=− 1
R
a −
R2
2
⋅g
R1 + R2
vout
a
R
=
≅− 2
vin 1 + af
R1
•
It can be quite tedious to try and morph arbitrary circuits into a
generic "af" block diagram
– Especially when impedances come into play
•
Elegant alternative: "Return Ratio Analysis"
– More later…
B. Murmann
EE 214 Lecture 14
20
175
Lecture 15
Fully Differential Amplifiers
Switched-Capacitor Circuits
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 15
1
Overview
•
Reading
– 6.0, 6.1.1, 6.1.2, 6.1.2 (Basic Feedback Concepts)
– 12.1, 12.2, 12.3, 12.4 (Fully Differential OpAmps)
– 6.1.7 (Internal Amplifiers)
•
Introduction
– Having discussed the basic properties of ideal feedback loops, we
will now take steps toward a practical implementation of feedback
amplifiers in CMOS technology. The first amplifier we consider uses
a very crude gain stage, composed of a simple differential pair and
an active current mirror load. For the time being, we will use this
simple topology as a vehicle to gain basic insight into terminology,
analysis and design tradeoffs.
– As we will argue, it is generally difficult to drive resistive loads with
an amplifier, as the load tends to reduce the circuit's gain. As a
result, many "internal amplifiers" in integrated circuits are based on
purely capacitive feedback, using a "switched-capacitor" approach.
B. Murmann
EE 214 Lecture 15
2
176
A Crude "High-Gain" Amplifier
•
Output common mode not well defined
– Depends on mismatch in currents between top and bottom
current mirror
B. Murmann
EE 214 Lecture 15
3
Common Mode Feedback (1)
MPa,b
MNT
•
Ictrl
Feedback loop forces Voc=Voc,desired by adjusting tail current
B. Murmann
EE 214 Lecture 15
4
177
Common Mode Feedback (2)
ID(VDS) of MP1a,b
Itail
= ID,MNT+Ictrl
Ictrl
0
VOC,des
VOC
VDD
•
Adjusting Ictrl changes VDS across PMOS loads (MP1a,b) and
hence alters VOC
•
Without feedback, VOC would be "randomly" set by mismatch
between load and tail current, and strongly depend on load gds
B. Murmann
EE 214 Lecture 15
5
Fully Differential vs. Single Ended (1)
•
•
Symmetrical
– Helps mitigate parasitic
coupling
– Good PSRR
– Easy to analyze
•
Lower complexity
– No CMFB
– Fewer feedback components
•
Can build non-inverting unity
gain buffer without using any
feedback components
Twice as much output swing
compared to single ended
B. Murmann
EE 214 Lecture 15
6
178
Fully Differential vs. Single Ended (2)
•
Most precision analog integrated circuits are based on fully
differential stages
– Data converters, filters, etc.
•
In contrast, printed circuit board circuits tend to be single ended
– Want minimum complexity and component count
•
In EE214, we will study mostly fully differential circuits
– More to come in EE315
•
In all of the upcoming assignments, we will use a very simple,
idealized common mode feedback circuit to avoid distraction
from the main design task
– Practical CMFB implementation examples will follow later in
this course
B. Murmann
EE 214 Lecture 15
7
Differential Mode Small Signal Half Circuit
B. Murmann
EE 214 Lecture 15
8
179
Load Considerations
•
Low load resistance will "destroy" the gain of our amplifier
– RL may may be an explicit load or due to loading from
feedback network
•
But, we want large (loop) gain for good precision
B. Murmann
EE 214 Lecture 15
9
Solution 1: Buffer
•
Can be very difficult to build
•
Can cost lots of headroom (e.g. CD stage)
•
Additional area, power
B. Murmann
EE 214 Lecture 15
10
180
Solution 2: Multi-Stage Amplifier
vod
vid
RL
CL
•
Resistive load "destroys" gain of second stage only
– First stage sees capacitive load
•
Costs additional area, power and must sacrifice stage 2 gain
B. Murmann
EE 214 Lecture 15
11
Solution 3: Don’t Use Resistors!
•
Can emulate resistors with "switched capacitors"
i=
V1 − V2
R
B. Murmann
Δq = C (V1 − V2 )
Δq Δq
iavg =
=
= f ⋅ C (V1 − V2 )
Δt T
Ravg =
1
f ⋅C
EE 214 Lecture 15
12
181
Switched Capacitor Circuits (1)
SC low-pass filter (passive)
SC integrator
SC gain stage
(actual implementations are fully differential)
B. Murmann
EE 214 Lecture 15
13
Switched Capacitor Circuits (2)
•
One of the most significant inventions in the history of ICs
•
Predominant approach for precision signal processing in CMOS
– CMOS technology provides good switches & capacitors
•
SC circuits have many advantages over RC implementations
– Transfer function set by ratio of capacitors
• RC product suffers from large process variations
– Corner frequencies (of filters) can be adjusted by changing
clock frequency
– Can make large time constants without using large resistors
• RC lowpass, 100Hz: R=16MΩ, C=100pF
• SC lowpass, 100Hz: f=10kHz, C1=6.25pF, C2=100pF
•
Reference
– R. Gregorian et al., "Switched-Capacitor Circuit Design,"
Proceedings of the IEEE, Vol. 71, No. 8, August 1983.
B. Murmann
EE 214 Lecture 15
14
182
SC Circuit During φ2
•
SC circuits can be very complicated,
but always boil down to same design
problem
– Build a (high-performance) amplifier
with capacitive feedback network
•
In EE214 we'll just design the circuit
during φ2 without worrying about the
switches
– More details on SC configurations
and switches are discussed in
EE315
– Particular focus in EE214 is placed
on gain stage configuration
• Gain set by capacitor ratio
B. Murmann
EE 214 Lecture 15
15
SC Gain Stage During φ2
•
Assuming ideal amplifier:
i = vi ⋅ jωCs
vo = −i ⋅
∴
B. Murmann
EE 214 Lecture 15
1
jωC f
vo
C
=− s
vi
Cf
16
183
Noise in SC Gain Stage (1)
•
Output is usually sampled by another switched capacitor stage
•
What is the relevant noise bandwidth?
B. Murmann
EE 214 Lecture 15
17
Noise in SC Gain Stage (2)
•
•
From Parseval's theorem, it follows that the noise
power of the output samples is equal to the power
spectral density of Vo (during φ2) integrated over all
frequencies
– Total integrated noise as discussed in lecture 9
The dynamic range of the circuit is thus given by:
– We'll derive expressions for different types of
amplifiers in lecture 19
B. Murmann
EE 214 Lecture 15
1 2
Vo , peak ,max
DR = 2
∞ 2
vn ,o
∫ Δf ⋅ df
0
18
184
OpAmps versus OTAs (1)
Operational Amplifier
B. Murmann
Operational Transconductance Amplifier
EE 214 Lecture 15
19
OpAmps versus OTAs (2)
OpAmp
OTA
•
"General Purpose"
•
•
Ideally a voltage controlled
voltage source
Most on-chip amplifiers are
OTAs
•
Ideally a voltage controlled
current source
•
High output impedance
•
Difficult to drive resistive
loads
•
Use capacitive (switched
capacitor) feedback network
•
Low output impedance
•
Can drive resistive and
capacitive loads
•
Essentially OTA + buffer
•
Buffer increases complexity
and power dissipation
B. Murmann
EE 214 Lecture 15
20
185
Lecture 16
Stability
Analysis of Feedback Circuits
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 16
1
Overview
•
Reading
– 9.3 (Stability)
– 8.8 (Return Ratio Analysis)
– Supplementary Handout "Feedback Systems" by Tom Lee
(see web, optional)
•
Introduction
– This lecture covers basics on the analysis of feedback
amplifiers. Using a simple OTA with capacitive feedback as
an example, we will study the so called "Return Ratio"
method as a tool to assess stability and calculate the
bandwidth of feedback amplifiers. Interestingly, most
relevant performance metrics follow directly from the "return
ratio" or loop gain of the circuit, which highlights the
significance of this parameter in feedback system design.
B. Murmann
EE 214 Lecture 16
2
186
Stability
A( s ) =
vo
a( s )
a( s )
=
=
vi 1 + a( s ) f ( s ) 1 + T ( s )
•
Most general criterion: BIBO
– Bounded input – bounded output
– Applies to any system
•
A continuous time linear system is BIBO stable if all its poles are
in the left half of the s-plane
– Can calculate roots of 1+T(s) to check stability
• Tedious and hard to do in general
B. Murmann
EE 214 Lecture 16
3
Methods for Checking Stability
•
Nyquist Criterion
– Based on evaluating T(s) in a polar plot
– Works for arbitrary T(s)
• Even if T(s) itself is unstable
– See books on control theory for details, or e.g.
• N. M. Nguyen and R. G. Meyer, "Start-up and Frequency Stability in
High-Frequency Oscillators," IEEE JSSC, pp. 810-820, May 1992.
•
Bode Criterion
– A subset of the general Nyquist criterion that can be applied when
T(s) itself is stable
• Safe to use in most electronic circuits
• Beware of exceptions
– System is unstable when |T(jω)| > 1 at the frequency where
Phase(T(jω)) = -180°
– Can use simple bode plot to check for stability
B. Murmann
EE 214 Lecture 16
4
187
Stability Measures
|T(jω)|
GM =
ω180
1
T ( jω ) ω =ω180
Typically want GM ≥ 3…5
ωc
Phase[T(jω)]
ωc
PM = 180° − Phase[T ( jω )]ω =ωc
ω180
Typically want PM ≥ 60…70°
B. Murmann
EE 214 Lecture 16
5
Closed Loop Peaking
Closed-loop gain,
normalized to 1/f
ω/ωc
[Text, p.632]
B. Murmann
EE 214 Lecture 16
6
188
OTA with Capacitive Feedback
•
Important questions
– Is this circuit stable? What is its phase margin?
– What is the low frequency closed loop gain?
– What is the closed loop bandwidth?
B. Murmann
EE 214 Lecture 16
7
Example: Simple Single Stage OTA
?
•
Issues
– Feedback network loads amplifier
– Amplifier loads feedback network
– At high frequencies, there exists a (capacitive) feedforward
path that overrides the amplifier
• Cannot be modeled using the generic "af" feedback block
diagram
B. Murmann
EE 214 Lecture 16
8
189
Solutions
•
If all we needed was the closed loop transfer function, we could
simply do a KCL/KVL based analysis
– Can be quite tedious, especially for more complex circuits
– Hard to assess stability and stability margin
•
Two port feedback analysis
– "Shunt-series, shunt-shunt, series-shunt, series-series"
feedback configurations
• See text, sections 8.4, 8.5, 8.6
– Attempts to identify amplifier (a) and feedback network (f)
with loading effects included
– Some feedback circuits cannot be modeled using two-ports
• E.g. bias circuits with feedback loops tend to have only one port
B. Murmann
EE 214 Lecture 16
9
Return Ratio Analysis (1)
•
Does not attempt to identify forward gain and feedback network
transfer functions separately
•
Analysis aims to identify gain around feedback loop
– Loop gain, loop transmission, a·f
• Different terminology for the same thing
•
From the loop gain of a circuit, we can determine
– Stability, closed loop gain characteristics, node impedances
•
Analysis can be applied to arbitrary feedback circuits,
independent of topology, port structure, etc.
•
We will first look at the complete framework of this technique
– Then identify a way to partition the analysis for our needs
and reduce its algebraic overhead
B. Murmann
EE 214 Lecture 16
10
190
Return Ratio Analysis (2)
1. Set all independent sources to zero
2. Identify a a controlled source in the feedback loop that you want
to analyze and break the loop by disconnecting the source
• E.g. VCCS, VCVS, …
3. Inject a test signal st at the breakpoint
• Current or voltage, depending on type of removed source
4. Find the return signal sr generated by the controlled source that
was disconnected from the circuit in step 2.
5. The return ratio of the controlled source is given by RR = –sr/st
• The text uses the symbol R , we will use RR for simplicity
• Provided that we have chosen a controlled source that
breaks the loop under consideration (and no other loop), the
return ratio of the source is equal to the loop gain of the
circuit, i.e. RR = T
B. Murmann
EE 214 Lecture 16
11
Example
v x = β ⋅ vo
β=
Cf
C f + Cs + C x
⎛
1 ⎞⎟
v0 = −it ⋅ ⎜⎜ R0
sCtot ⎟⎠
⎝
T (s ) = −
B. Murmann
C Ltot = C L + (1 − β )C f
⎛
ir
1
= β ⋅ Gm ⋅ ⎜ R0
⎜
it
sC L tot
⎝
EE 214 Lecture 16
⎞
⎟ = β ⋅ Gm R0
⎟ 1 + sRoC L
tot
⎠
12
191
Loop Gain Expression
T (s ) = β ⋅ Gm R0 ⋅
1
1 + sRo C L + (1 − β )C f
[
Capacitive feedback divider,
includes amplifier input
capacitance Cx
β=
•
]
Feedback loading
Cf
C f + Cs + C x
In literature, β is often called "feedback factor"
– We will call it "return factor" to avoid confusion with f of the
generic feedback block diagram
B. Murmann
EE 214 Lecture 16
13
Phase Margin (1)
T ( jω )
Ro→∞
T0
Important to note: Ro is irrelevant for
high-frequency analysis
ωp
T (s ) =
β ⋅ Gm R0
1 + sRoC L tot
≅
β ⋅ Gm
sC L tot
ωc
ω
at high frequencies, as long as
ωc
≅ βGm Ro >> 1
ωp
βGm
βGm
= 1 ⇒ ωc ≅
jωcC L tot
C L tot
B. Murmann
EE 214 Lecture 16
14
192
Phase Margin (2)
T ( jω )
Ro→∞
T0
ωp
ωc
ω
Phase[T ( jω )]
ωp
0°
ω
− 45°
⎛ω ⎞
Phase[T ( jω )] ω =ωc = − tan −1 ⎜ c ⎟ ≅ −90°
⎜ω ⎟
⎝ p⎠
PM ≅ 180° − 90° ≅ 90°
− 90°
B. Murmann
EE 214 Lecture 16
15
Closed Loop Transfer Function
A(s ) =
vo
T( s )
d
= A∞
+
vi
1+ T( s ) 1+ T( s )
⎛
1 af
⎜⎜ Note similarit y : A =
f 1 + af
⎝
⎞
⎟⎟
⎠
•
A∞ is the closed loop gain with ideal feedback (Gm→∞)
•
d is the direct signal feedthrough with the controlled source
removed (Gm→0)
– Can often be ignored; we'll look at this term later…
B. Murmann
EE 214 Lecture 16
16
193
Finding A∞
•
With infinite Gm, vx must be zero, and there is no current through Cx
0 = vi sCs + vo sC f
A∞ = −
B. Murmann
Cs
Cf
EE 214 Lecture 16
17
Closed Loop Bandwidth
T (s ) =
A(s ) ≅ A∞
β ⋅ Gm R0
1 + sRoC L tot
≅
β ⋅ Gm
sC L tot
ωc
T( s )
≅ A∞
1 + T( s )
s + ωc
=
ωc
s
(at high frequencies)
ωc
jω−3 dB + ωc
=
1
⇒ ω−3dB = ωc
2
•
Closed loop bandwidth is equal to unity gain (crossover)
frequency of |T(s)|
•
Consistent with what we've learned before; closed loop pole
should be ~T0 times larger than open-loop pole
ω−3dB ≅ ω pT0 =
B. Murmann
1
sC Ltot Ro
⋅ Gm Ro = ωc
EE 214 Lecture 16
18
194
Finding d
•
Writing a general expression for d will be quite messy
•
Better to distinguish the two cases of interest from the beginning
– Low frequency (Ro<<|1/jωC|)
– High Frequency (Ro>>|1/jωC|)
B. Murmann
EE 214 Lecture 16
19
High Frequency Result
A( s ) = A∞
Detailed analysis shows:
T( s )
d
+
1+ T( s ) 1+ T( s )
A(s ) = −
1−
sC f
Cs
Gm
C f 1 + sC Ltot
βGm
•
Pole as expected
•
Zero at Gm/Cf, typically a very high frequency
– Far beyond ω-3dB if CL is large
•
In most cases, calculating feedforward zeros is easier using
simple KCL analysis, with large resistors removed
B. Murmann
EE 214 Lecture 16
20
195
Low Frequency Result
A0 = A∞
d0 = 0
⇒
T0
d
+ 0
1 + T0 1 + T0
A0 = −
Cs T0
C f 1 + T0
•
No feedforward, capacitors are open circuits
•
Define static gain error
ε=
A∞ − A0
A∞
ε = 1−
⎛
A0
T
1
1⎞
= 1− 0 = 1−
≅ 1 − ⎜⎜1 − ⎟⎟
1
A∞
1 + T0
⎝ T0 ⎠
1+
T0
ε≅
B. Murmann
1
T0
EE 214 Lecture 16
21
Summary – OTA with Capacitive Feedback
•
To find the static gain error
– Write an expression for the low frequency loop gain, neglect
all impedances due to capacitors
– Note that capacitive voltage dividers still work at low
frequency!
•
To find the closed loop bandwidth
– Write an expression for the high frequency loop gain, neglect
finite output resistance of all devices
– The 3-dB bandwidth of the closed loop circuit is
(approximately) equal to the unity gain frequency of T(jω)
•
To assess stability and phase margin
– Determine phase of (high-frequency) T(jω) at its unity gain
frequency
• Boring for single pole systems, more interesting if two or more
poles are involved…
B. Murmann
EE 214 Lecture 16
22
196
Comparison
•
Two-Port Analysis
•
Return Ratio Analysis
•
a, f
•
T
•
1/f
•
A∞
A=
1 af
f 1 + af
A = A∞
T
d
+
1+T 1+T
(feedforward effects are not modeled)
B. Murmann
EE 214 Lecture 16
23
197
Lecture 17
Loop Gain Simulation
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 17
1
Overview
•
Introduction
– Last lecture, we've seen that return ratio analysis is a useful
tool for characterizing feedback amplifiers in terms of
stability, bandwidth and static precision. Since we are always
interested in verifying our hand analysis results using
simulations, it is desirable to have an equivalent method
available in Spice. In this lecture, we will discuss the socalled Middlebrook method for loop gain simulation. This
particular approach helps overcome the issue that ideal loop
breakpoints, e.g. at controlled sources, are not directly
accessible in Spice.
B. Murmann
EE 214 Lecture 17
2
198
References
•
H.W. Bode, Network Analysis and Feedback Amplifier Design, Van
Nostrand, New York, 1945.
•
R.D. Middlebrook, "Measurement of Loop Gain in Feedback Systems,"
Int. J. Electronics, Vol. 38, No.4, .pp. 485-512, 1975.
•
S. Rosenstark, "Loop Gain Measurement in Feedback Amplifiers," Int.
J. Electronics, Vol. 57, No.3., pp. 415-421, 1984.
•
P.J. Hurst, "Exact Simulation of Feedback Circuit Parameters," Trans.
on Circuits and Systems, pp.1382-1389, Nov. 1991.
•
P.J. Hurst, S.H. Lewis, "Simulation of Return Ratio in Fully Differential
Feedback Circuits," Proc. CICC 1994, pp.29-32.
•
Ken Kundert, "A Test Bench for Differential Circuits," Online:
http://www.designers-guide.com/Analysis/diff.pdf
•
M. Tian, V. Visvanathan, J. Hantgan, K. Kundert, "Striving for smallsignal stability," IEEE Circuits and Devices Magazine, pp. 31-41,
January 2001.
B. Murmann
EE 214 Lecture 17
3
Circuit Example 1
•
What is the loop gain in this circuit?
B. Murmann
EE 214 Lecture 17
4
199
Return Ratio Analysis
T (s ) = −
ir
it
•
As we've seen last lecture, such a hand calculation is pretty
straightforward
•
How can we simulate T(jω) in Spice?
– Using "real" transistor models
B. Murmann
EE 214 Lecture 17
5
Spice MOSFET AC Simulation Model
•
Nodes of ideal controlled source are not accesible!
– Cannot break loop at gm generator
B. Murmann
EE 214 Lecture 17
6
200
Popular Simulation Approach
T ( jω ) ≅ −
B. Murmann
vr
vt
•
Inaccurate
•
Hard to estimate mock load
•
May get different results for
different breakpoints
•
Ideally, we'd like to avoid all
of the above issues
•
Solution: Middlebrook
method
EE 214 Lecture 17
7
Problem Generalization
•
Middlebrook argued that any single single loop feedback circuit
can be partitioned as shown below
•
Hence, there is always some "nonideal" breakpoint between
impedances
– How can we use this breakpoint to find the loop gain?
available breakpoint
T ( s ) = gm
B. Murmann
EE 214 Lecture 17
Z1 ⋅ Z 2
Z1 + Z 2
8
201
Double Injection Trick
−
vy
vx
≡ Tv = g m ⋅ Z 2 +
Z2
Z1
True Loop T = g ⋅ Z1Z 2
m
Gain:
Z1 + Z 2
iy
ix
≡ Ti = g m ⋅ Z1 +
Solving yields:
1
1
1
=
+
1 + T 1 + Tv 1 + Ti
T =
Z1
Z2
•
No “DC“ break in the loop, all loading effects included!
•
Measure Tv and Ti separately, then calculate actual T
B. Murmann
EE 214 Lecture 17
TvTi − 1
Tv + Ti + 2
9
Simulation Example
Ti =
•
iy
ix
Tv = −
vy
vx
Two options
– Run two copies of the same circuit simultaneously, or
– Run two simulations with different stimuli
B. Murmann
EE 214 Lecture 17
10
202
Spice Code
* utils.sp
.param ai=1 av=0
m1
vo x
0
i1
0
100u
vx middle x
dc 0
cf
vo y
200f
vy middle y
ac 'av'
rf
vo y
100gig
it 0 middle
ac 'ai'
cs
y
0
400f
xt
x
y
looptest ai='ai' av='av'
.ends looptest
vo
0
nch214
L=0.35u W=10u
.subckt looptest x y ai=0 av=0
.op
.ac dec 10 1e3 10e9
.options post brief
.lib './ee214_hspice.txt' nominal
.include utils.sp
.alter
.param ai=0 av=1
.end
B. Murmann
EE 214 Lecture 17
11
Matlab Postprocessing
% load signals from first run and calculate Ti
m = loadsig('middlebrook1.ac0');
ti = evalsig(m, 'I_xt_vy')./evalsig(m, 'I_xt_vx')
% load signals from second run and calculate Tv
m = loadsig('middlebrook1.ac1');
tv = -evalsig(m, 'y')./evalsig(m, 'x')
f = 1e-6*evalsig(m, 'HERTZ');
% calculate loop gain
t
= (tv.*ti-1)./(2+tv+ti)
% calculate magnitude, phase and plot
% ...
B. Murmann
EE 214 Lecture 17
12
203
Magnitude [dB]
Simulation Result
40
20
Tv
Ti
T
0
-20
Phase [degrees]
-40 -2
10
10
-1
10
0
10
f [MHz]
1
10
2
10
3
10
4
10
0
10
f [MHz]
1
10
2
10
3
10
4
0
Tv
Ti
T
-50
-100 -2
10
B. Murmann
10
-1
EE 214 Lecture 17
13
How to do this in Awaves?
•
Generally much more tedious, but some shortcuts make it
possible to do this
– Remember that 1/(1+T) = 1/(1+Tv) + 1/(1+Ti)
– Plugging in voltages & currents for Tv and Ti, this becomes
1/(1+T) = vx/(vx-vy) + ix/(ix+iy)
– With unity test voltage and current, this simplifies to
1/(1+T) = vx/1V + ix/1A
– Probe real and imaginary parts of vx and ix in HSpice using
•
•
•
•
•
.probe ac vxreal = par('vr(x)')
.probe ac vximag = par('vi(x)')
.probe ac ixreal = par('ir(xt.vx)')
.probe ac iximag = par('ii(xt.vx)')
Can now create expressions for phase and magnitude of T and
plot using real and imaginary components of vx and ix
B. Murmann
EE 214 Lecture 17
14
204
Circuit Example 2
Cf
CL
Cs
+
vid
-
+
vod
Cs
CL
Cf
B. Murmann
EE 214 Lecture 17
15
Fully Differential Testbench
Tv = −
v yp − v ym
Ti =
v xp − v xm
B. Murmann
EE 214 Lecture 17
i yp − i ym
ixp − ixm
16
205
Differential Looptest Circuit
.subckt difflooptest xp xm yp ym ai=0 av=0
vxp middlep xp
dc 0
vyp middlep yp
ac 'av'
vxm middlem xm
ac 'av'
vym middlem ym
dc 0
it
middlem middlep
ac 'ai'
.ends difflooptest
B. Murmann
EE 214 Lecture 17
17
Phase [degrees]
Magnitude [dB]
Simulation Result
40
20
-20
Tv
Ti
T
-40 -2
10
10
0
-1
10
0
10
f [MHz]
1
10
2
10
3
10
4
10
0
10
f [MHz]
1
10
2
10
3
10
4
0
-50
-100 -2
10
B. Murmann
Tv
Ti
T
10
-1
EE 214 Lecture 17
18
206
Alternative Approach
•
Use baluns to convert differential signals into CM/DM
components and use simple single ended loop test circuit
•
The above setup is particularly for use with Spectre simulator
– Spectre has built-in Middlebrook analysis, called "stb"
– Simply replace looptest block with "iprobe" element
– Spectre automatically finds and plots T, no post-processing
needed
B. Murmann
EE 214 Lecture 17
19
Ideal Balun
xfmr
xfmr
.subckt balun vdm vcm vp vm
=
e1 vp vcm transformer vdm 0 2
e2 vcm vm transformer vdm 0 2
.ends balun
•
Useful for separating CM and DM signal components
•
Bi-directional, preserves port impedance
•
Uses ideal, inductorless transformers that work down to DC
•
Not available in all simulators (but often possible to emulate, see
[Kundert])
B. Murmann
EE 214 Lecture 17
20
207
+ New Middlebrook Method
•
Recently, Middlebrook came up with an alternative way of looking at
feedback circuits
– A more "design oriented analysis"
•
He titled this approach "The General Feedback Theorem – A final
solution for feedback systems"
•
If you are curious about this new method, please refer to
– http://rdmiddlebrook.com
– http://ardem.com/free_downloads.asp
– http://groups.yahoo.com/group/Design-Oriented_Analysis_D-OA/
B. Murmann
EE 214 Lecture 17
21
Multi Loop Considerations
•
Any practical feedback circuit has multiple feedback loops
– Fully differential circuits have CM/DM loops
– Local device feedback through Cgd, Rsource
– ...
•
Solutions
– Decompose fully differential circuit into CM/DM loops
– If a local feedback loop can be modeled as a combination of
a stable controlled source and passive impedances, the
multi-loop circuit reduces to a single loop [Hurst 94]
– If there is a common breakpoint that breaks all feedback
loops simultaneosly, stability can be checked by finding the
return ratio at the single breakpoint [Hurst 94]
B. Murmann
EE 214 Lecture 17
22
208
Last Resort: General Nyquist Criterion
[Bode 45]:
“If a circuit is stable when all its tubes have their
nominal gains, the total number of clockwise and
counterclockwise encirclements of the critical
point must be equal to each other in the series of
Nyquist diagrams for the individual tubes
obtained by beginning with all tubes dead and
restoring the tubes successively in any order to
their nominal gains“
[You may want to take a controls class if you are
interested in this...]
B. Murmann
EE 214 Lecture 17
23
Another Useful Quote
[Bode 45]:
“... thus the circuit may sing when the
tubes begin to lose their gain because of
age, and it may also sing, instead of
behaving as it should, when the gain
increases from zero as power is supplied
to the circuit...“
Always run one or more transient analyses for a
"true" stability check!
B. Murmann
EE 214 Lecture 17
24
209
Lecture 18
Differential Mode Voltage Range
Two-Stage OTA
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 18
1
Overview
•
Reading
– 6.3 (Basic Two-Stage MOS Amplifiers)
– 12.6.1 (Fully Differential Two-Stage-Amplifier)
– 12.6.5 (Neutralization)
•
Introduction
– As we will see in today's lecture, the available output swing
and open-loop gain of our simple single stage OTA are fairly
limited. Hence, we will begin to consider a two-stage OTA
architecture as an alternative. Unfortunately, with the
addition of a second stage, we also introduce a second pole,
which makes it difficult to achieve reasonable phase margin.
The compensation techniques discussed in the following
lecture will provide solutions to this problem.
B. Murmann
EE 214 Lecture 18
2
210
Output Swing of Simple OTA
•
Available output swing depends on input and output common
mode levels
•
May be limited by headroom for differential pair device (Vminn) or
active load (Vminp)
B. Murmann
EE 214 Lecture 18
3
Maximum Available Swing
•
Input and output common
mode adjusted such that all
devices operate at "edge" of
forward active region
– Well defined using long
channel model, very
gradual transition in short
channels
•
Unfortunately, the choice of
Vic and Voc are often dictated
by the circuits that interface
with the amplifier
– E.g. Vic=Voc=1.5V
Vodpp ,max = 2(VDD − Vmin p − Vmin n − Vmin t )
B. Murmann
EE 214 Lecture 18
4
211
Example Vic=Voc=VDD/2
•
Assuming that we are limited by Vminn, and Vminn~VOV, the
available differential peak-to-peak swing is ~4Vt
•
Since the transition to triode is smooth, which criterion should
we use find the "exact" output range of an amplifier?
B. Murmann
EE 214 Lecture 18
5
Gain vs. Output Swing DC Simulation
90
Vod/Vid [V/V]
80
-30%
70
60
Vodpp,max
50
40
•
-1.5
-1
-0.5
0
Vod [V]
0.5
1
1.5
In EE214, we arbitrarily define output range as the peak-to-peak
swing that causes no more than 30% drop in Vod/Vid
B. Murmann
EE 214 Lecture 18
6
212
How Much Gain Can We Get?
•
Small signal gain (around Vid=Vod=0):
avo = g mn ⋅
ron ⋅ rop
ron + rop
= avon
avo = avon
1
1
= avon
r
g a
1 + on
1 + mp von
rop
g mn avop
1
(g m / I D ) p
1+
(g m / I D )n
avo = avon || avon
avon
avop
for
(g m / I D ) p = (g m / I D )n
•
E.g. avon=avop=50, (gm/ID)n= (gm/ID)p ⇒ avo=25
•
Static gain error ~1/To ~1/avo ~1/25=4%
– Not precise enough for many applications
B. Murmann
EE 214 Lecture 18
7
Two-Stage Amplifier
•
More output swing, more gain ~(gmro/2)2
•
Output range no longer depends on input common mode
B. Murmann
EE 214 Lecture 18
8
213
Common Mode Feedback
•
Same as for single stage OTA!
•
Common mode of first stage output is set via VGS of common
source device in second stage
B. Murmann
EE 214 Lecture 18
9
Simplified AC Half Circuit with Feedback
T (s ) = β
B. Murmann
g m1 R1 ⋅ g m 2 R2
= β ⋅ a1 (s )a2 (s ) = β ⋅ a(s )
⎛
s ⎞ ⎛
s ⎞
⎜⎜1 − ⎟⎟ ⋅ ⎜⎜1 − ⎟⎟
p1 ⎠ ⎝
p2 ⎠
⎝
EE 214 Lecture 18
p1 = −
1
R1C1
p2 = −
1
R2C2
10
214
How About Miller Effect?
•
Two ways to deal with Miller amplification of Cgd in first stage
– Use cascodes
• Often needed for high gain, anyways
– Use "neutralization caps"
• See text, section 12.6.5
B. Murmann
EE 214 Lecture 18
11
Neutralization Caps
•
With neutralization caps
Cin = C gs + C gd 1 (1 + av ) + Cn (1 − av )
•
Letting Cn=Cgd gives
Cin = C gs + 2C gd 1
•
Great! How about cancelling Cin completely?! I.e. letting
Cn =
•
C gs + C gd 1 (1 + av )
av − 1
VERY IMPRACTICAL, because av is not well controlled
– Your circuit may end up being a great oscillator…
B. Murmann
EE 214 Lecture 18
12
215
Bode Plot of Loop Gain
Mag ( jω )
β ⋅ a (s )
a2 (s )
a1 (s )
•
ω
ω p1 ω p 2
Phase( jω )
β ⋅ a(s )
0°
If ωp1 and ωp2 are
close to each
other, the loop will
essentially have
no phase margin!
− 90°
ω
−180°
B. Murmann
EE 214 Lecture 18
13
Introducing a Dominant Pole
Mag ( jω )
β ⋅ a(s )
a2 (s )
a1 (s )
Phase[T ( jω )]
ω p1
ωc
The problem is
solved if we
somehow
manage to make
ωp1<< ωp2
– Or ωp2<< ωp1
•
Loop behaves
close first order
system around
crossover
frequency
ω
ω p2
0°
− 90°
ω
−180°
B. Murmann
•
EE 214 Lecture 18
14
216
Phase Margin
•
At the crossover frequency, the dominant pole has shifted the
phase by about -90°
•
The non-dominant pole's phase at ωc is given by -tan-1(ωc/ωp2)
⎛ ω p2 ⎞
⎟⎟
PM ≅ tan −1 ⎜⎜
⎝ ωc ⎠
⎛ω ⎞
PM ≅ 180° − 90° − tan −1 ⎜ c ⎟
⎜ω ⎟
⎝ p2 ⎠
ωp2/ωc
PM
1
45°
2
63°
3
72°
4
76°
5
79°
B. Murmann
EE 214 Lecture 18
15
Creating a Dominant Pole
•
Numerical example:
β = 0.5
f p2 =
R1 = R2 = 100kΩ
Gm1 = Gm 2 = 1mS
1
= 1.6MHz
2πR2C2
fc =
C1 =
•
f p2
3
= 530kHz
C2 = 1 pF
f p1 =
PM = 72°
fc
= 106 Hz
β ⋅ Gm1 R1 ⋅ Gm 2 R2
1
= 15nF
2π ⋅ f p1 ⋅ R1
Two issues
– Very low fc, which means low closed loop bandwidth
– Huge capacitor
• Get roughly 1fF/μm2 in CMOS technology
• C1 would occupy about 4mm x 4mm !
B. Murmann
EE 214 Lecture 18
16
217
Utilizing the Miller Effect
•
Purposely connect an additional capacitor between gate and
drain of M2 (Cc = "Compensation capacitor")
•
Two interesting things happen
– Low frequency input capacitance of second stage becomes
large – exactly what we need for low ωp1
– At high frequencies, Cc turns M2 into a diode connected
device – low impedance, i.e. large ωp2 !
B. Murmann
EE 214 Lecture 18
17
Pole Splitting
c
•
More analysis next lecture…
B. Murmann
EE 214 Lecture 18
18
218
Lecture 19
Compensation
Noise in Feedback OTAs
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 19
1
Overview
•
Reading
– 9.4.1, 9.4.2, 9.4.3 (Compensation)
•
Introduction
– In this lecture, we will continue to look at pole splitting as a
method for achieving sufficient phase margin in a two-stage
amplifier. While using a Miller capacitance for compensation
helps move the non-dominant pole to higher frequency, it
also introduces an undesired zero in the transfer function.
Today, we'll discuss several options on how to cope with this
artifact.
– In addition, we will derive useful expressions for the total
integrated noise in OTAs.
B. Murmann
EE 214 Lecture 19
2
219
Two-Stage OTA with Cc
•
Detailed analysis shows:
⎛
C ⎞
g m1 R1 ⋅ g m 2 R2 ⋅ ⎜⎜ 1 − s c ⎟⎟
gm2 ⎠
v
⎝
a(s ) = o =
vi 1 + s[(C2 + Cc )R2 + (C1 + Cc )R1 + g m 2 R2 R1Cc ] + s 2 R1 R2 (C1C2 + Cc C2 + Cc C1 )
•
Very messy…
B. Murmann
EE 214 Lecture 19
3
Dominant Pole Approximation
•
We can write the denominator as
⎛
⎛1
s ⎞ ⎛
s ⎞
1 ⎞
s2
⎜
⎟
⎜
⎟
⎜
⎟
D (s ) = ⎜ 1 − ⎟ ⋅ ⎜ 1 − ⎟ = 1 − s ⎜ + ⎟ +
p1 ⎠ ⎝
p2 ⎠
⎝
⎝ p1 p2 ⎠ p1 p2
•
Since in a practical design outcome we'll have |p1|<<|p2|, we can
approximate
⎛1⎞
s2
D(s ) ≅ 1 − s⎜⎜ ⎟⎟ +
⎝ p1 ⎠ p1 p2
•
With this simplification, we can now easily identify p1 and p2 by
comparing the coefficients with the expression from the previous
slide
B. Murmann
EE 214 Lecture 19
4
220
Final Result
⎛ s⎞
⎜1 − ⎟
⎝ z⎠
a(s ) ≅ av 0 ⋅
⎛
s ⎞ ⎛
s ⎞
⎜⎜1 − ⎟⎟ ⋅ ⎜⎜1 − ⎟⎟
p1 ⎠ ⎝
p2 ⎠
⎝
p1 ≅ −
1
R1 (C1 + Cc ) + R2 (C2 + Cc ) + g m 2 R2 R1Cc
p2 ≅ −
•
gm2
C1C2
+ C1 + C2
Cc
≅−
z=+
1
g m 2 R2 R1Cc
gm2
Cc
Questions
– How can we design an amplifier with these complex
expressions?
– What will the zero in the transfer function do to us?
B. Murmann
EE 214 Lecture 19
5
RHP vs. LHP Zero
ωz
− ωz
Phase
0°
Phase
ωz
ω
0°
− 45°
+ 45°
− 90°
+ 90°
B. Murmann
EE 214 Lecture 19
ω
ωz
6
221
OTA Transfer Function with RHP Zero
•
RHP zero will reduce
phase margin
– Unless gm2 >> βgm1
ωc ≅ β
ωz
ωz =
g m1
Cc
gm2
Cc
(assuming zero
is beyond
crossover)
ωz 1 gm2
=
ωc β g m1
B. Murmann
EE 214 Lecture 19
7
Mitigating the Impact of RHP Zero
•
Create unilateral feedback through Cc
– Source follower
– Cascode compensation
• Ahuja, IEEE J. Solid-State Ckts., 12/1983
• Ribner, IEEE J. Solid-State Ckts., 12/1984
•
"Nulling resistor"
– Push zero to infinity
– Push zero into LHP and cancel nondominant pole (!)
B. Murmann
EE 214 Lecture 19
8
222
Source Follower
•
Mitigates feedforward path issue
•
Problems: Reduced output swing, additional power dissipation
B. Murmann
EE 214 Lecture 19
9
Cascode Compensation (1)
•
Cc sees low impedance, output of first stage sees high
impedance looking into cascode device
– Reduced feedforward
B. Murmann
EE 214 Lecture 19
10
223
Cascode Compensation (2)
•
Benefits
– Tends to push ωp2 to higher frequencies, when load
capacitor is large (see text)
• Can use smaller Cc, less power in first stage
•
Issues
– Additional power dissipation
– Bias current mismatches cause input referred offset
•
The above two issues can be addressed by feeding back to
cascode device embedded in first stage (Ribner, JSSC 12/1984)
– New issue: Complex design problem (3rd order system)
B. Murmann
EE 214 Lecture 19
11
Nulling Resistor (1)
•
New transfer function becomes
⎛ 1
⎞
− Rz ⎟⎟
1 − sCc ⎜⎜
⎝ gm2
⎠
a (s ) ≅ a v 0 ⋅
⎛
s ⎞ ⎛
s ⎞ ⎛
s ⎞
⎜⎜1 − ⎟⎟ ⋅ ⎜⎜1 − ⎟⎟ ⋅ ⎜⎜1 − ⎟⎟
p1 ⎠ ⎝
p2 ⎠ ⎝
p3 ⎠
⎝
•
p1 and p2 unchanged, new pole p3, and a knob to tune the zero
B. Murmann
EE 214 Lecture 19
12
224
Nulling Resistor (2)
•
Rz=1/gm2 pushes the zero to +∞
– Can use a transistor in triode region to implement resistor
• Helps track process variations
B. Murmann
EE 214 Lecture 19
13
Nulling Resistor (3)
•
Rz=(1+C2/Cc)/gm places zero on top of ωp2!
– Cancels p2
– Good in theory, may be troublesome in practice
– If the pole and zero don't fall exactly on top of each other, we
get a so-called pole-zero doublet
• Can cause very slow settling tails in transient response
– B.Y.T Kamath, R.G. Meyer and P.R. Gray, "Relationship between frequency
response and settling time of operational amplifiers," IEEE JSSC, Vol. 9, No.
6, pp.347–352, Dec. 1974.
• See supplementary handout "Effect of Doublet on Amplifier
Settling Time" (optional)
•
Third pole
ω p3 ≅
B. Murmann
1
g
≅ m2
Rz C1 C1
ω p 3 1 g m 2 Cc
=
ωc β g m1 C1
EE 214 Lecture 19
typically >> 1
14
225
Other Compensation Methods
•
Nested Miller compensation
– >2 gain stages
– Higher order response presents design challenge
– Not (yet?) used much
– Ref: R. G. H. Eschauzier and J. H. Huijsing. Frequency
Compensation Techniques for Low-Power Operational
Amplifiers. Kluwer, 1995.
•
Lag/lead compensation
– See handout "Feedback Systems" on website
– An attempt to improve bandwidth/phase margin by adding
additional zeros to T(s)
– May introduce doublets and worsen noise performance
B. Murmann
EE 214 Lecture 19
15
Total Integrated Noise in Feedback OTAs
•
General method
– Identify noise sources
– Derive noise transfer functions (NTF)
– Find total noise power from each source by integrating
PSD·NTF from 0 to ∞
– Add up noise powers
•
Tedious, but doable…
•
Examples
– Single-stage amplifier
– Single-stage amplifier with cascode
– Two-stage amplifier
B. Murmann
EE 214 Lecture 19
16
226
Useful Integrals
B. Murmann
EE 214 Lecture 19
17
Single Stage Amplifier
VB
Cf
+
vi
-
M1
Cs
g ⎞
kT γ ⎛
⎜⎜ 1 + m 2 ⎟⎟
C Ltot β ⎝
g m1 ⎠
C gg 1 ⎞⎛
kT ⎛⎜
⎟⎜ 1 + g m 2 ⎞⎟
=
γ 1 + A∞ +
⎜
C Ltot ⎝
C f ⎟⎠⎜⎝
g m1 ⎟⎠
vo2,tot =
M2
+
CL vo
-
C Ltot = C L + C junction + C f (1 − β )
•
Make gm2 as small as possible, i.e. use small gm/ID for current
source device
– Issue: Smaller gm/ID means less available swing
•
Small Cgg1, i.e. high fT helps reduce noise
B. Murmann
EE 214 Lecture 19
18
227
Single-Stage Amplifier with Cascode
•
Analysis shows
g
ω ⎞
kT 1 ⎛⎜
1 + m2 c ⎟
C Ltot β ⎜⎝
g m1 ω p 2 ⎟⎠
kT 1 ⎛
1 g m2 ⎞
⎜⎜ 1 +
⎟
=
C Ltot β ⎝ k g m1 ⎟⎠
vo2,tot =
•
ωc = β
g m1
C Ltot
ω p2 =
gm2
Cx
k=
ω p2
ωc
Make gm2 as small as possible, i.e. use small gm/ID for
cascode device
– Reduces gm2/gm1 and Cx
– Issue: Smaller gm/ID means less available swing
B. Murmann
EE 214 Lecture 19
19
Two-Stage Amplifier
vo2,tot =
γ kT ⎛ g m11 ⎞ kT
⎟+
⋅ ⎜1 +
g m1 ⎟⎠ C Ltot
β Cc ⎜⎝
Stage 1
⎡
⎛
g m 22 ⎞⎤
⎟⎥
⎢1 + γ ⎜⎜ 1 +
g m 2 ⎟⎠⎦⎥
⎝
⎣⎢
Stage 2
•
Need large Cc for low noise
•
Stage 2 noise can be significant if CL is small and β is large
B. Murmann
EE 214 Lecture 19
20
228
SNR in Differential Circuits
SNR ∝
V̂o2
kT
C
•
In fully differential circuits, the effective output swing is doubled
– But there are two half circuits that contribute noise
•
Signal power increases by 4x, noise power increases by 2x
– 3dB win in terms of SNR
– At the cost of twice the power consumption (no free lunch…)
B. Murmann
EE 214 Lecture 19
21
Appendix
"Noise in a Two-stage OTA with Nulling Resistor"
B. Murmann
EE 214 Lecture 19
22
Noise in a Two-stage OTA with Nulling Resistor
Boris Murmann, 11/26/2006
(Initial derivations by Alireza Dastgheib)
229
kT
β Cc
1 kT
⋅γ⋅
(
) (
g m2⋅ ∆
(
) (
)
2
+ C1⋅ Cc
⎦
2⎤
)
(
) (
)
)
β Cc
1 kT
⋅γ⋅
Cc
g m1
(
∆
g m2⋅ Cc
) (
g m2⋅ ∆
ω p2 =
)
β ⋅ g m1⋅ Cc⋅ C1 + CL + g m2 − β ⋅ g m1 ⋅ ∆
ω c1 = β ⋅
Now use these expressions to simplify further:
(
ω p1
ω p2
Cc
ω p2⋅ ∆
simplify
substitute , ω p2 = k ⋅ ω c1
substitute , g m2 =
β
ω c1⋅ Cc
→ N1 =
2
1 kT
∆
⋅
⋅ γ ⋅ k⋅
2
β Cc
3
3
2
Cc ⋅ C1 + Cc ⋅ CL + k ⋅ ∆ − ∆ ⋅ Cc
∆ = C c ⋅ C L + C c ⋅ C 1 + C L⋅ C 1
substitute , g m1 =
k=
m1⋅ Cc ⋅ C1 + CL + g m2 − β ⋅ g m1 ⋅ Cc⋅ ∆
2
(
g m2⋅ ⎡CL⋅ C1 + Cc
⎣
g m2⋅ CL + C1 ⋅ Cc
kT
N3 =
⋅
CL β ⋅ g m1⋅ Cc⋅ C1 + CL + g m2 − β ⋅ g m1 ⋅ ∆
N1 =
)
β ⋅ g m1⋅ Cc C1 + CL + g m2 − β ⋅ g m1 ⋅ ∆
N2 =
⋅γ⋅
CL
β⋅g
N1 =
N1 and N2 are the total integrated noise at the output due to M1 and M2, respectively. N3 is the total noise due to R.
Letting R=1/gm2 simplifies these expressions slightly:
230
(
1 kT
⋅
⋅γ⋅
β Cc
)
1+
2
1 kT
⋅
⋅γ⋅
β Cc
1−
)
(
C c ⋅ C L⋅ C 1
2
1
)
2
substitute , ∆ = Cc⋅ CL + Cc⋅ C1 + CL⋅ C1
2
→ −Cc ⋅ CL⋅ C1
simplify
2
k ⋅ C c ⋅ C L + C c ⋅ C 1 + C L⋅ C 1
2
k⋅ ∆
2
1 kT
⋅
⋅γ
β Cc
N2 looks messy unless we let C1->0...
N1 =
The ratio of capacitors in this expression is always <1, also k usually 3...4. For C1->0, the last term above disappears. Hence, the noise
contribution from M1 is well approximated by
N1 =
(
1
Cc ⋅ Cc C1 + CL − ∆ ⋅ Cc
Cc ⋅ C c⋅ C 1 + C L − ∆ ⋅ Cc
2
N1 =
231
=
γ
1
⋅
2
(
(
) (
)
2
+ C1⋅ Cc
⎦
2⎤
)
(
Ntot = N1 + N2 + N3 =
2
)2 + C1⋅ Cc2⎤⎦
)
Cc
ω p2⋅ ∆
β
ω c1⋅ Cc
simplify
substitute , C1 = 0
substitute , ∆ = Cc⋅ CL + Cc⋅ C1 + CL⋅ C1
substitute , ω p2 = k ⋅ ω c1
substitute , g m2 =
substitute , C1 = 0 N3
1
→
=
N2
γ
simplify
1 kT
kT
⋅
⋅γ +
⋅ (γ + 1)
CL
β Cc
g m2⋅ ⎡CL⋅ C1 + Cc
⎣
(
g m2⋅ CL + C1 ⋅ Cc
m1⋅ Cc ⋅ C1 + CL + g m2 − β ⋅ g m1 ⋅ Cc⋅ ∆
Total Noise for C1->0:
N2
N3
kT
N2 =
⋅γ
CL
kT
N2 =
⋅γ⋅
CL
β⋅g
g m2⋅ ⎡CL⋅ C1 + Cc
⎣
substitute , g m1 =
kT
→ N2 =
⋅γ
CL
232
233
Lecture 20
OTA Design Considerations
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 20
1
Overview
•
Introduction
– Today, we will review a possible design strategy for twostage OTAs. Since there exist many degrees of freedom in
this topology, a thorough hand analysis is essential, and can
help minimize, if not completely eliminate, the time needed
for a large number of Spice iterations.
B. Murmann
EE 214 Lecture 20
2
234
Two-stage OTA Design Strategy
•
The design equations for a two-stage OTA are fairly complex,
and involve several capacitances that may not be known a priori
•
Solution: Iterative hand analysis/design
– Most efficiently done using e.g. Excel, Matlab or MathCAD
•
Possible design flow
1. Pick Cgg1, Cgg2 based on heuristics (see following slides)
2. Use initial guesses for junction capacitances
3. Calculate Cc based on noise spec
4. Find gm, fT, gm/ID
5. Iterate, using different choices in step 1
6. Find W for Spice verification
7. Resolve potential discrepancies
– E.g. refine Cjunction estimates if necessary
B. Murmann
EE 214 Lecture 20
3
A Note on Discrepancies
•
Discrepancies between design script and Spice are usually on
the order of 10-20%
– Mostly due to
• Bias point dependence (VDS); charts generated for VDS=VDD/2
• Inaccuracy in Cjunction estimate
• First order nature of design equations
•
Good news
– It is always possible to track discrepancies down if needed
• Look at gm/ID , CGG, Cjunction in the bias point output
•
Big difference to square law design using μCox, VOV, …
– These quantities simply don’t exist in your circuit…
B. Murmann
EE 214 Lecture 20
4
235
Component Identification
C1 = C gg 2 + C junction1
C2 = C L + (1 − β )C f + C junction 2
•
β=
Cf
C f + Cs + C gg1
For simplicity, we'll first neglect the junction caps in the following
discussion
– Straightforward to include in your optimization script (see
example)
B. Murmann
EE 214 Lecture 20
5
Loop Crossover Frequency
ωc ≅
g
1
⋅ β ⋅ g m1 R1 g m 2 R2 = β m1
g m 2 R2 R1Cc
Cc
β=
Cf
C f + Cs + C gg1
•
Small Cgg1 helps increase ωc
– But there is diminishing return if Cgg1 becomes small
compared to Cs, Cf
•
Typically a good starting point: Cgg1 ≈ Cf + Cs
– Sometimes optimum, see final exam 2005
B. Murmann
EE 214 Lecture 20
6
236
Nondominant Pole
ω p2 ≅
gm2
C Ltot C gg 2
Cc
1
ωp2
CLtot ≅ C L + (1 − β )C f
+ C gg 2 + C Ltot
⎛ C gg 2 C Ltot
≅ ⎜⎜
+
⎝ gm2 gm2
C Ltot C gg 2
⎛
⎜
C + C Ltot
⎞⎜
⎟⎟⎜ 1 + gg 2
Cc
⎠⎜
⎜
⎝
⎞
⎟
⎟
⎟
⎟
⎟
⎠
•
Heuristic: Start optimization with Cgg2 ≅ CLtot
•
For a given ωp2 target and fixed CLtot
– Choosing Cgg2 much smaller than CLtot means excess
gm2/Cgg2=ωT2 and therefore small (gm/ID)2
– Choosing Cgg2 much larger than CLtot will cost excess gm2
(power) to meet ωp2 target
B. Murmann
EE 214 Lecture 20
7
Choice of Cc
•
In practice, and particularly for high DR designs, Cc is often set
by noise requirements
•
For designs that are not constraint by noise, it is Interesting to
assume Cgg2=CLtot (not necessarily optimum) to develop further
qualitative insight
1
2 ⎛ C Ltot ⎞
– With this choice, we have
⎟
⎜1 +
≅
ω p 2 ωT 2 ⎜⎝
2Cc ⎟⎠
•
Cc large means that we can use lower ωT2 (higher gm/ID) and
save power in the second stage
– But larger Cc also requires larger gm1 and thus more power in
the first stage
– This implies that there will be a design-dependent optimum
for Cc
B. Murmann
EE 214 Lecture 20
8
237
CMFB Loop (1)
•
How to pick g?
B. Murmann
EE 214 Lecture 20
9
CMFB Loop (2)
•
CMFB loop is a feedback circuit that can be handled in the same way
we analyzed the differential loop
– Draw small signal model, find loop crossover, PM, etc.
•
The value of g will set the CMFB bandwidth
– How fast should we make the CMFB loop?
•
Bandwidth of CMFB loop should be "comparable" to differential loop
– So that any significant disturbances on Voc can recover quickly
•
Typically OK to make CMFB bandwidth ~30% of differential loop
bandwidth
– In a typical switched capacitor circuit with 10 time constants of
differential settling, this means that the common mode has about 3
time constants to settle
• Enough time to remove ~95% of common mode disturbance
B. Murmann
EE 214 Lecture 20
10
238
CMFB Loop (3)
•
Systematic error in Voc
ΔVoc ≅
I D (MB ) − I D (MTa ) − I D (MTb ) ΔI D
=
g
g
•
E.g. ΔID=100μA, g=10mS ⇒ ΔVoc=10mV
•
Bounds for g
– Upper bound due to stability
– Lower bound due to CMFB bandwidth requirements
B. Murmann
EE 214 Lecture 20
11
Appendix
2-Stage OTA Design Example ("Simple Approach")
2-Stage OTA Design Example ("Accurate Approach")
B. Murmann
EE 214 Lecture 20
12
239
Reference:C:\Documents and Settings\Murmann\My Documents\lib\Mathcad\defaults.mcd
2-Stage OTA Design Example ("Simple Approach")
Cf
CL
Cs
Vsd
+
+
Vod
-
Vid
Cs
CL
Cf
M4a,b
M1a,b
M3a,b
M2a,b
(Miller compensation, neutralization caps and CMFB not shown)
240
Technology Data
VDD := 3V
γ :=
2
3
gm/ID, gm/Cgg lookup tables:
Reference:C:\Documents and Settings\Murmann\My Documents\teaching\ee214_autumn07\otaexample\gm
Design Objectives
fc := 200MHz
PM := 75deg
DR := 72 (dB)
G := 2
ε s := 0.5%
Given Component Values
Cs := 400fF
CL := 200fF
Cf :=
Cs
G
Design Decision: Channel lengths
Return factor estimate assuming Cgg1 is ~equal to sum of feedback caps
Cgg1 := Cs + Cf
avo :=
1
β est :=
Cf
Cf + Cs + Cgg1
β est = 0.167
3
avo = 1.2 × 10
ε s⋅ β est
For simplicity, assume that each stage contributes same gain (not necessarily accurate/optimal)
avostage :=
avo
avostage = 34.641
For simplicity, assume that intrinsic gain of both signal path devices is 2x stage gain (not
necessarily accurate/optimal). Intrinsic gain requirements:
gmro := 2avostage
gmro = 69.282
Based on gmro design charts, we decide:
L1 := 0.55μm
L2 := 0.4μm
For simplicity, choose same lengths for pmos/nmos loads (not necessarily optimal)
L3 := L2
L4 := L1
241
Design Decision: gm/ID
S
gmID1 := 10
A
S
gmID2 := 10
A
Design Decision: Relative gm/ID of active loads, output swing
Ratio of gm/ID in load devices vs. signal path (often want g1,2 to be <1, to minimize noise and
Cjunction; however, very small g1,2 may impose DR limitations)
g 31 =
gmID3
g 31 := 1
gmID1
g 42 =
gmID4
g 42 := 1
gmID2
Allocated Vdsmin for M2 and M4 and resulting output swing:
VDSmin := 500mV
Vodmax := VDD − 2 ⋅ VDSmin
(Differential peak amplitude)
Step 1: Calculate Cc based on DR spec
(
)
C2 := CL + 1 − β est ⋅ Cf
C2 = 366.667 fF
Total stage 2 load
2
Ntot :=
0.5⋅ Vodmax
Ntot = 355.234 μV
DR
10
10
k B ⋅ Tr
⎡ 1 k B ⋅ Tr
⎤
Ntot = 2 ⋅ ⎢ ⋅
⋅ γ ⋅ 1 + g 31 +
⋅ ⎡γ ⋅ 1 + g 42 + 1⎤ ⎥
⎦
CL ⎣
⎣ β Cc
⎦
(
1
Cc := 2 ⋅
β est
Ntot −
)
(
(
⋅ k B⋅ Tr⋅ γ ⋅ 1 + g 31
k B ⋅ Tr
C2
(
)
)
)
⋅ ⎡γ ⋅ 1 + g 42 + 1⎤
⎣
⎦
Cc = 0.659 pF
242
Step 2: Calculate transconductances
g m1 :=
1
β est
k := tan⎛⎜ π ⋅
⎝
⋅ 2 ⋅ π ⋅ fc⋅ Cc
PM
g m1 = 4.966 mS
⎞ k = 3.732
⎟
180deg ⎠
g m2 := k ⋅ 2 ⋅ π ⋅ fc⋅ C2
g m2 = 1.72 mS
(approx.)
Step 3: Find ID, W
g m1
ID1 :=
gmID1
g m2
ID2 :=
gmID2
(
IDW1 := pidw L1 , gmID1
)
ID1 = 0.497 mA
(
IDW2 := nidw L2 , gmID2
ID1
W1 :=
IDW1
ID2
W2 :=
IDW2
)
IDW1 = 1.947
A
m
W1 = 255.015 μm
ID2 = 0.172 mA
IDW2 = 8.893
A
m
W2 = 19.336 μm
Calculate the width ratios (w31 = W3/W1) for the chosen gm/ID
(
)
(
)
pidw L1 , gmID1
w31 :=
nidw L3 , gmID1 ⋅ g 31
(
nidw L2 , gmID2
w42 :=
pidw L4 , gmID2 ⋅ g 42
(
W3 := W1 ⋅ w31
)
w31 = 0.219
)
w42 = 4.567
W4 := W2 ⋅ w42
W3 = 55.835 μm
55
3
= 18.333
W4 = 88.313 μm
88
6
= 14.667
243
OTA circuit for simulation (feedback network, CMFB, Cc, etc. not shown)
85uA
85uA
170uA
1020uA
M4c
15/0.55
M=12
15/0.55
M4d
M4a,b
15/0.55
M=2
M1a,b
255/0.55
18/0.4 18/0.4
M3c
M3d
M3a,b
18/0.4
M=6
M2a,b
19/0.4
Simulation results (first run, without any tweaking):
Magnitude [dB]
f c=130.75MHz, PM=55.78deg, T0=245
80
60
40
20
0
-20
-40 -2
10
10
0
10
2
10
4
Phase [degrees]
f [MHz]
0
-50
-100
-150
10
-2
10
0
10
f [MHz]
2
10
4
170uA
PSD [V2/Hz]
244
10
-20
Sqrt(Integral) [μVrms]
10
5
10
10
f [Hz]
Integral=417.87uVrms, DR=70.59dB (for Vodmax=2.00V)
500
0
10
5
10
10
f [Hz]
Loop crossover and phase margin are way off! Why? Look at .op output:
gm1 = 5.12mS
gm2 = 1.88mS
Not bad. What's wrong?
Junction caps are fairly large and comparable to other caps. E.g. Cdb1~300fF; this will
significantly impact nondominant pole.
Cgg1=688fF, means that beta is smaller than what we had budgetd above (beta,est)
....
Bottom line: Hard to get "reasonable" matching between hand calculations and Spice using
simplified expressions and ignoring junctions caps. Let's fix this...
245
subckt
x1
x1
x1
x1
x1
x1
element 1:m1a
1:m1b
1:m2a
1:m2b
1:m3d
1:m3c
model
0:pch214
0:pch214
0:nch214
0:nch214
0:nch214
0:nch214
region
Saturati
Saturati
Saturati
Saturati
Saturati
Saturati
id
-515.7219u -515.7219u 209.9559u 209.9559u
85.0000u 100.5239u
ibs
0.
0.
0.
0.
0.
0.
ibd
0.
0.
0.
0.
0.
0.
vgs
-1.0731
-1.0731
795.4179m 795.4179m 726.2152m 726.2152m
vds
-1.2776
-1.2776
1.4843
1.4843
726.2152m
1.9455
vbs
926.9464m 926.9464m
0.
0.
0.
0.
vth
-924.8901m -924.8901m 594.7855m 594.7855m 602.4640m 590.0494m
vdsat
-186.1804m -186.1804m 162.1420m 162.1420m 116.5494m 123.5860m
vod
-148.1634m -148.1634m 200.6324m 200.6324m 123.7513m 136.1659m
beta
31.1696m
31.1696m
11.1159m
11.1159m
10.5036m
10.4989m
gam eff 458.3707m 458.3707m 894.1238m 894.1238m 894.1238m 894.1238m
gm
5.1229m
5.1229m
1.8862m
1.8862m
1.1167m
1.2474m
gds
67.8360u
67.8360u
20.4147u
20.4147u
14.0027u
12.6163u
gmb
903.7574u 903.7574u 495.7571u 495.7571u 298.8398u 328.6908u
cdtot
419.4667f 419.4667f
25.7526f
25.7526f
28.4889f
22.8775f
cgtot
681.3669f 681.3669f
37.5110f
37.5110f
34.9915f
35.1430f
cstot
938.5513f 938.5513f
69.9117f
69.9117f
65.4864f
65.6493f
cbtot
767.3791f 767.3791f
69.2340f
69.2340f
69.7566f
64.1621f
cgs
528.6546f 528.6546f
27.3328f
27.3328f
25.2053f
25.3129f
cgd
122.2292f 122.2292f
4.3896f
4.3896f
4.1602f
4.1588f
subckt
x1
x1
x1
x1
x1
x1
element 1:m3b
1:m3a
1:m4d
1:m4c
1:m4b
1:m4a
model
0:nch214
0:nch214
0:pch214
0:pch214
0:pch214
0:pch214
region
Saturati
Saturati
Saturati
Saturati
Saturati
Saturati
id
515.7219u 515.7219u -100.5239u
-1.1886m -209.9559u -209.9559u
ibs
0.
0.
0.
0.
0.
0.
ibd
0.
0.
0.
0.
0.
0.
vgs
726.2152m 726.2152m
-1.0545
-1.0545
-1.0545
-1.0545
vds
795.4179m 795.4179m
-1.0545 -926.9464m
-1.5157
-1.5157
vbs
0.
0.
0.
0.
0.
0.
vth
601.7593m 601.7593m -739.0005m -740.0639m -735.1556m -735.1556m
vdsat
116.9465m 116.9465m -311.3437m -310.5103m -314.3549m -314.3549m
vod
124.4559m 124.4559m -315.5066m -314.4432m -319.3516m -319.3516m
beta
63.0200m
63.0200m
1.8139m
21.7668m
3.6278m
3.6278m
gam eff 894.1238m 894.1238m 472.5709m 472.5709m 472.5709m 472.5709m
gm
6.7553m
6.7553m 559.8132u
6.6300m
1.1578m
1.1578m
gds
81.4843u
81.4843u
10.9880u 147.0550u
17.4548u
17.4548u
gmb
1.8061m
1.8061m 127.5074u
1.5107m 263.4207u 263.4207u
cdtot
167.9230f 167.9230f
29.4059f 361.5686f
54.5446f
54.5446f
cgtot
210.0083f 210.0083f
39.8665f 478.3955f
79.7341f
79.7341f
cstot
392.9864f 392.9864f
66.0755f 792.9036f 132.1528f 132.1528f
cbtot
415.5359f 415.5359f
62.0693f 753.5347f 119.8678f 119.8678f
cgs
151.2783f 151.2783f
30.0686f 360.8146f
60.1421f
60.1421f
cgd
24.9604f
24.9604f
7.0195f
84.2342f
14.0390f
14.0390f
246
Reference:C:\Documents and Settings\Murmann\My Documents\lib\Mathcad\defaults.mcd
2-Stage OTA Design Example ("Accurate Approach")
Cf
CL
Cs
Vsd
+
+
Vod
-
Vid
Cs
CL
Cf
M4a,b
M1a,b
M3a,b
M2a,b
(Miller compensation, neutralization caps and CMFB not shown)
247
Technology Data
2
VDD := 3V
γ :=
Lmin := 0.35μm
k dbn := 0.65
k dbp := 0.8
3
(approximate ratios of Cdb/Cgg for minimum length device)
gm/ID, gm/Cgg lookup tables:
Reference:C:\Documents and Settings\Murmann\My Documents\teaching\ee214_autumn07\otaexample\gm
Design Objectives
fc := 200MHz
PM := 75deg
DR := 72 (dB)
G := 2
ε s := 0.5%
Given Component Values
Cs := 400fF
CL := 200fF
Cf :=
Cs
G
Design Decision: Channel lengths
Return factor estimate assuming Cgg1 is ~equal to sum of feedback caps
Cgg1 := Cs + Cf
avo :=
1
β est :=
Cf
Cf + Cs + Cgg1
β est = 0.167
3
ε s⋅ β est
avo = 1.2 × 10
For simplicity, assume that each stage contributes same gain (not necessarily accurate/optimal)
avostage :=
avo
avostage = 34.641
For simplicity, assume that intrinsic gain of both signal path devices is 2x stage gain (not
necessarily accurate/optimal). Intrinsic gain requirements:
gmro := 2avostage
gmro = 69.282
Based on gmro design charts, we decide:
L1 := 0.55μm
L2 := 0.4μm
248
For simplicity, choose same lengths for pmos/nmos loads (not necessarily optimal)
L3 := L2
L4 := L1
Design Decision: Relative gm/ID of active loads, output swing
Ratio of gm/ID in load devices vs. signal path (often want g1,2 to be <1, to minimize noise and
Cjunction; however, very small g1,2 may impose DR limitations)
g 31 =
gmID3
g 31 := 1
gmID1
g 42 =
gmID4
g 42 := 1
gmID2
Allocated Vdsmin for M2 and M4 and resulting output swing:
VDSmin := 500mV
Vodmax := VDD − 2 ⋅ VDSmin
(Differential peak amplitude)
Usually have to leave margins for PVT and gain drop at swing. Depending on final design outcome
(optimum gm/ID), it may be worth re-visiting this assumtion.
ITERATION
Adjust following parameters to minimize total current computed in step 4
c1 =
c2 =
Cgg1
Cs + Cf
Cgg2
(good starting point: c1=1)
c1 := 1
(good starting point: c2=1)
c2 := 1
CL
Step 1: Estimate/calculate return factor and all capacitances
β :=
Cf
(Cf + Cs)⋅ (1 + c1)
(
)
β = 0.167
Cgg1 := c1 ⋅ Cs + Cf
Cgg1 = 600 fF
Cgg2 := c2 ⋅ CL
Cgg2 = 200 fF
249
Lmin
Cdb1 := k dbp ⋅ Cgg1 ⋅
L1
Cdb1 = 305.455 fF
To find Cdb3, we estimate the width ratio w31 = W3/W1, assuming gm/ID1=10S/A (not known at
this point; also won't matter much...)
w31 :=
pidw⎛⎜ L1 , 10
S⎞
⎟
⎝
nidw⎛⎜ L3 , 10
⎝
A⎠
S
A
w31 = 0.219
⋅ g 31⎞⎟
⎠
Lmin
Cdb2 := k dbn ⋅ Cgg2 ⋅
L2
w42 :=
nidw⎛⎜ L2 , 10
pidw⎛⎜ L4 , 10
⎟
⎝
A⎠
S
A
Cdb3 = 54.338 fF
k dbp
Cdb4 := w42⋅ Cdb2 ⋅
k dbn
Cdb4 = 639.428 fF
Cdb2 = 113.75 fF
S⎞
⎝
k dbn
Cdb3 := w31⋅ Cdb1 ⋅
k dbp
w42 = 4.567
⋅ g 42⎞⎟
⎠
C1 := Cdb1 + Cdb3 + Cgg2
C1 = 559.793 fF
Total stage 1 load
C2 := CL + Cdb2 + Cdb4 + ( 1 − β ) ⋅ Cf
C2 = 1.12 × 10 fF
3
Total stage 2 load
Step 2: Calculate Cc based on DR spec
2
Ntot :=
0.5⋅ Vodmax
Ntot = 355.234 μV
DR
10
10
k B ⋅ Tr
⎡ 1 k B ⋅ Tr
⎤
Ntot = 2 ⋅ ⎢ ⋅
⋅ γ ⋅ 1 + g 31 +
⋅ ⎡γ ⋅ 1 + g 42 + 1⎤ ⎥
⎣
⎦
CL
⎣ β Cc
⎦
(
1
Cc := 2 ⋅
β
Ntot −
)
(
(
)
)
⋅ k B⋅ Tr⋅ γ ⋅ 1 + g 31
k B ⋅ Tr
C2
(
)
Cc = 0.56 pF
⋅ ⎡γ ⋅ 1 + g 42 + 1⎤
⎣
⎦
(Note: for low DR designs, Cc is not neccesarily set by noise and becomes part of the optimization
process (e.g. add variable c3=cc/CL to the iteration loop)
250
Step 3: Calculate transconductances and transit frequencies
g m1 :=
1
β
⋅ 2 ⋅ π ⋅ fc⋅ Cc
k := tan⎛⎜ π ⋅
g m1 = 4.222 mS
⎞ k = 3.732
⎟
⎝ 180deg ⎠
PM
⎛ C2⋅ C1
g m2 := k ⋅ 2 ⋅ π ⋅ fc⋅ ⎜
⎝
fT1 :=
1
⋅
Cc
⎞
+ C1 + C2⎟
g m2 = 13.128 mS
⎠
g m1
fT1 = 1.12 GHz
2 ⋅ π Cgg1
1 g m2
fT2 :=
⋅
2 ⋅ π Cgg2
fT2 = 10.447 GHz
Step 4: Find gm/ID, ID
(
)
(
gmID1 := pgmid L1 , fT1 gmID2 := ngmid L2 , fT2
g m1
ID1 :=
gmID1
g m2
ID2 :=
gmID2
IDtotal := ID1 + ID2
)
1
1
gmID1 = 10.404
gmID2 = 6.21
V
V
ID1 = 0.406 mA
ID2 = 2.114 mA
IDtotal = 2.52 mA
This design (with c1=1, c2=1 -> IDtotal=2.52mA) will certainly work, but there's lots of room for
power optimization. With the above script, it is quite easy to iteratively step c1 and c2 up/down to
search for a power minimum.
For c1=1, c2=1, it is interesting to note that there is lots of self-loading (The Cdb's present a large
fraction of the caps that set fc and fp2). Hence, going to smaller c1 and/or c2 may help lower
power.
Another try with c1=0.5, c2=1 yields IDtotal=1.819mA. Making the pmos smaller reduces its
contributed capacitance faster than gm/ID drops; beta also improves for small c1. Hence, there is
a net reduction in power.
Similarly, using yet another try with c1=0.5, c2=0.5 yields IDtotal=1.497mA.
Manual iterations are very useful for developing intuition; but it is also possible to automate the
search process using an optimization function:
251
(
)
f c1 , c2 :=
Cf
β←
(Cf + Cs)⋅ (1 + c1)
Cgg1 ← c1 ⋅ ( Cs + Cf)
Cgg2 ← c2 ⋅ CL
Lmin
Cdb1 ← k dbp ⋅ Cgg1 ⋅
L1
pidw⎛⎜ L1 , 10
S⎞
⎟
⎝
Cdb3 ←
nidw⎛⎜ L3 , 10
⎝
S
A
A⎠
⋅ g 31⎞⎟
⋅ Cdb1 ⋅
⎠
k dbn
k dbp
Lmin
Cdb2 ← k dbn ⋅ Cgg2 ⋅
L2
S⎞
nidw⎛⎜ L2 , 10
⎟
⎝
Cdb4 ←
pidw⎛⎜ L4 , 10
⎝
S
A
A⎠
⋅ g 42⎞⎟
⋅ Cdb2 ⋅
⎠
k dbp
k dbn
C1 ← Cdb1 + Cdb3 + Cgg2
C2 ← CL + Cdb2 + Cdb4 + ( 1 − β ) ⋅ Cf
1
β
Cc ← 2 ⋅
Ntot −
g m1 ←
1
β
k B ⋅ Tr
C2
⎣
PM
)
⎦
⎞
⎟
⎛ C2 ⋅ C1
⎝ Cc
fT2 ←
(
⋅ ⎡γ ⋅ 1 + g 42 + 1⎤
180 ⋅ deg ⎠
g m2 ← k ⋅ 2 ⋅ π ⋅ fc⋅ ⎜
fT1 ←
)
⋅ 2 ⋅ π ⋅ fc⋅ Cc
k ← tan⎛⎜ π ⋅
⎝
(
⋅ k B⋅ Tr⋅ γ ⋅ 1 + g 31
⎞
+ C1 + C2 ⎟
⎠
g m1
1
⋅
2 ⋅ π Cgg1
1
⋅
g m2
2 ⋅ π Cgg2
1
1
gmID1 ← if ⎛⎜ pgmid L1 , fT1 > 4 ⋅ , pgmid L1 , fT1 , 0.1⋅ ⎞⎟
V
V⎠
⎝
(
⎛
)
(
1
)
1⎞
252
gmID2 ← if ⎛⎜ ngmid L2 , fT2 > 4 ⋅
⎝
(
)
1
V
(
)
, ngmid L2 , fT2 , 0.1⋅
1⎞
⎟
V⎠
g m2
g m1
+
IDtotal ←
gmID2
gmID1
IDtotal
Initial guess for optimization
c1 := 1
c2 := 1
Given
c1 > 0
c2 > 0
(
Copt := Minimize f , c1 , c2
Copt =
⎛ 0.145 ⎞
⎜
⎟
⎝ 0.379 ⎠
)
f⎛ Copt , Copt ⎞ = 0.989 mA
0
1⎠
⎝
M := CreateMesh( f , 0.01 , 0.8 , 0.01 , 0.8 , 40 , 40)
M
Comments:
- Shallow power minimum for small c1 and c2
- the optimum is close to the "steep cliff" imposed by limiting gm/ID to practical values >4S/A in the
objective function (in the power minimum, gm/ID1=4.7S/A, gm/ID2=7.7S/A)
253
Re-calculate components with optimizer result:
c1 := Copt
0
c2 := Copt
1
Estimate/calculate return factor and all capacitances
β :=
Cf
β = 0.291
(Cf + Cs)⋅ (1 + c1)
(
)
Cgg1 := c1 ⋅ Cs + Cf
Cgg1 = 87.269 fF
Cgg2 := c2 ⋅ CL
Cgg2 = 75.833 fF
Lmin
Cdb1 := k dbp ⋅ Cgg1 ⋅
L1
Cdb1 = 44.428 fF
w31 :=
pidw⎛⎜ L1 , 10
⎝
nidw⎛⎜ L3 , 10
⎝
S
A
S⎞
⎟
A⎠
w31 = 0.219
⋅ g 31⎞⎟
⎠
Lmin
Cdb2 := k dbn ⋅ Cgg2 ⋅
L2
w42 :=
nidw⎛⎜ L2 , 10
⎝
pidw⎛⎜ L4 , 10
⎝
S
A
k dbn
Cdb3 := w31⋅ Cdb1 ⋅
k dbp
Cdb3 = 7.903 fF
k dbp
Cdb4 := w42⋅ Cdb2 ⋅
k dbn
Cdb4 = 242.447 fF
Cdb2 = 43.13 fF
S⎞
⎟
A⎠
w42 = 4.567
⋅ g 42⎞⎟
⎠
C1 := Cdb1 + Cdb3 + Cgg2
C1 = 128.164 fF
Total stage 1 load
C2 := CL + Cdb2 + Cdb4 + ( 1 − β ) ⋅ Cf
C2 = 627.376 fF
Total stage 2 load
Step 2: Calculate Cc based on DR spec
1
Cc := 2 ⋅
β
Ntot −
(
)
⋅ k B⋅ Tr⋅ γ ⋅ 1 + g 31
k B ⋅ Tr
C2
(
)
⋅ ⎡γ ⋅ 1 + g 42 + 1⎤
⎣
⎦
Cc = 0.34 pF
254
Step 3: Calculate transconductances and transit frequencies
g m1 :=
1
β
⋅ 2 ⋅ π ⋅ fc⋅ Cc
k := tan⎛⎜ π ⋅
⎝
PM
g m1 = 1.469 mS
⎞
⎟
k = 3.732
180deg ⎠
⎛ C2⋅ C1
g m2 := k ⋅ 2 ⋅ π ⋅ fc⋅ ⎜
⎝ Cc
fT1 :=
1
⋅
⎞
+ C1 + C2⎟
g m2 = 4.652 mS
⎠
g m1
fT1 = 2.679 GHz
2 ⋅ π Cgg1
1 g m2
fT2 :=
⋅
2 ⋅ π Cgg2
fT2 = 9.763 GHz
Step 4: Find gm/ID, ID
(
)
(
gmID1 := pgmid L1 , fT1 gmID2 := ngmid L2 , fT2
g m1
ID1 :=
gmID1
)
g m2
ID2 :=
gmID2
1
gmID1 = 4.639
V
1
gmID2 = 6.916
V
ID1 = 0.317 mA
ID2 = 0.673 mA
Step 5: Calculate device widths
(
IDW1 := pidw L1 , gmID1
ID1
W1 :=
IDW1
W3 := W1 ⋅ w31
)
(
IDW2 := nidw L2 , gmID2
)
IDW1 = 9.535
A
m
IDW2 = 17.356
A
m
ID2
W2 :=
IDW2
W1 = 33.208 μm
W2 = 38.755 μm
W4 := W2 ⋅ w42
W3 = 7.271 μm
W4 = 177.009 μm
255
OTA circuit for simulation (feedback network, CMFB, Cc, etc. not shown)
50uA
100uA
700uA
700uA
M4c
25/0.55
M=7
25/0.55
M4d
M4a,b
25/0.55
M=7
M1a,b
33/0.55
1/0.4
M3d
M=2
1/0.4
M3c
M3a,b
1/0.4
M=7
M2a,b
39/0.4
Simulation results (first run, without any tweaking):
Magnitude [dB]
f c=181.34MHz, PM=75.68deg, T0=298
80
60
40
20
0
-20
-40 -2
10
10
0
10
2
10
4
Phase [degrees]
f [MHz]
0
-50
-100
-150
10
-2
10
0
10
f [MHz]
2
10
4
700uA
PSD [V2/Hz]
256
10
-20
Sqrt(Integral) [μVrms]
10
5
10
10
f [Hz]
Integral=351.77uVrms, DR=72.09dB (for Vodmax=2.00V)
400
200
0
10
5
10
10
f [Hz]
Very close to specs!
- fc is about 10% lower than expected. This is mostly due to the dominant pole approximation; the
second pole pulls fc to lower frequencies. Also, Cgd of M2 adds additional compensation
capacitance, which also reduces fc.
- PM and DR are essentially right on target.
The discrepancies, in general, can be resolved in two ways: (1) Spice tweaking (OK for small
changes), (2) Re-visit above calculations and improve assumtions and equations. E.g. factor in the
expected error from the dominant pole approximation.
An advantage of the presented methodology is that most, if not all discrepancies/errors can be
tracked down by comparing the Spice component values (gm, gm/ID, Cdb, ... from .op) with those
used in the optimization routine.
Close inspection of the .op values below reveals that most small signal parameters calculated
above agree with Spice to within 10-20%.
257
subckt
element
model
region
id
ibs
ibd
vgs
vds
vbs
vth
vdsat
vod
beta
gam eff
gm
gds
gmb
cdtot
cgtot
cstot
cbtot
cgs
cgd
x1
1:m1a
0:pch214
Saturati
-344.6871u
0.
0.
-1.2816
-1.4099
718.4015m
-885.7614m
-384.5490m
-395.8371m
3.9883m
461.1502m
1.5295m
27.5794u
282.3301u
55.0457f
88.4114f
126.0118f
104.3431f
68.7789f
15.6526f
x1
x1
x1
x1
x1
1:m1b
1:m2a
1:m2b
1:m3d
1:m3c
0:pch214
0:nch214
0:nch214
0:nch214
0:nch214
Saturati
Saturati
Saturati
Saturati
Saturati
-344.6871u 772.2910u 772.2910u
50.0000u 104.2676u
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
-1.2816
871.7327m 871.7327m
1.0988
1.0988
-1.4099
1.4986
1.4986
1.0988
2.0205
718.4015m
0.
0.
0.
0.
-885.7614m 595.0075m 595.0075m 584.3535m 574.9695m
-384.5490m 205.6017m 205.6017m 304.2131m 307.9633m
-395.8371m 276.7252m 276.7252m 514.4807m 523.8647m
3.9883m
23.0145m
23.0145m 531.3387u
1.0629m
461.1502m 894.1238m 894.1238m 894.1238m 894.1238m
1.5295m
4.9575m
4.9575m 161.8125u 331.7534u
27.5794u
56.8387u
56.8387u
2.9301u
4.0365u
282.3301u
1.2918m
1.2918m
42.9472u
86.8132u
55.0457f
52.2020f
52.2020f
2.0215f
3.4641f
88.4114f
77.6753f
77.6753f
1.7924f
3.5848f
126.0118f 143.1825f 143.1825f
4.4518f
8.8976f
104.3431f 140.4000f 140.4000f
5.1705f
9.7628f
68.7789f
56.7937f
56.7937f
1.3590f
2.7114f
15.6526f
9.0726f
9.0726f 207.4045a 414.8090a
subckt
x1
x1
x1
x1
x1
x1
element 1:m3b
1:m3a
1:m4d
1:m4c
1:m4b
1:m4a
model
0:nch214
0:nch214
0:pch214
0:pch214
0:pch214
0:pch214
region
Saturati
Saturati
Saturati
Saturati
Saturati
Saturati
id
344.6871u 344.6871u -104.2676u -703.2478u -772.2910u -772.2910u
ibs
0.
0.
0.
0.
0.
0.
ibd
0.
0.
0.
0.
0.
0.
vgs
1.0988
1.0988 -979.5312m -979.5312m -979.5312m -979.5312m
vds
871.7327m 871.7327m -979.5312m -718.4015m
-1.5014
-1.5014
vbs
0.
0.
0.
0.
0.
0.
vth
586.6658m 586.6658m -739.8673m -742.0442m -735.5167m -735.5167m
vdsat
303.2846m 303.2846m -251.4877m -249.7596m -254.9415m -254.9415m
vod
512.1684m 512.1684m -239.6639m -237.4870m -244.0145m -244.0145m
beta
3.7191m
3.7191m
3.0506m
21.3547m
21.3541m
21.3541m
gam eff 894.1238m 894.1238m 472.5709m 472.5709m 472.5709m 472.5709m
gm
1.1179m
1.1179m 741.4965u
5.0238m
5.4115m
5.4115m
gds
27.2459u
27.2459u
13.0870u 115.7219u
74.4836u
74.4836u
gmb
297.8439u 297.8439u 170.1165u
1.1533m
1.2402m
1.2402m
cdtot
14.9146f
14.9146f
49.3740f 364.8318f 316.8690f 316.8690f
cgtot
12.5467f
12.5467f
67.0284f 469.1707f 469.2508f 469.2508f
cstot
31.1678f
31.1678f 109.9526f 769.6267f 769.7437f 769.7437f
cbtot
36.9572f
36.9572f 103.2448f 741.9384f 693.9432f 693.9432f
cgs
9.5188f
9.5188f
50.4655f 353.2113f 353.3468f 353.3468f
cgd
1.4518f
1.4518f
11.8133f
82.6932f
82.6926f
82.6926f
258
Lecture 21
Step Response
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 21
1
Overview
•
Reading
– 7.5 (Relation Between Frequency and Time Response)
•
Reference
– H.C. Yang and D.J. Allstot, "Considerations for fast settling
operational amplifiers," IEEE Trans. Ckts. and Syst., March
1990, pp. 326 – 334.
•
Introduction
– OTAs with capacitive feedback are primarily used in
switched capacitor circuits, where fast settling to voltage
steps at the input is critical. In today's lecture we will look at
first and second order behavior in the step response of
feedback OTAs.
B. Murmann
EE 214 Lecture 21
2
259
Motivation
•
In switched capacitor circuits, the amplifiers have to respond to
voltage steps
•
How fast of a switched capacitor circuit can we build?
B. Murmann
EE 214 Lecture 21
3
Analysis
•
Assuming a single pole system, we have
V (s)
C T0
1
A(s ) ≅ out
=− s
⋅
Vin ( s )
C f 1 + T0 1 + s
ωc
B. Murmann
T0 = β ⋅ Gm Ro
ωc ≅ β ⋅
EE 214 Lecture 21
Gm
C Ltot
β=
Cf
C f + Cs + Cin
C Ltot = C L + (1 − β ) ⋅ C f
4
260
Step Response
Vout ( s ) = A( s ) ⋅ Vin ( s )
Vout ( t ) = L−1 {A( s ) ⋅ Vin ( s )}
Vstep ⎫
⎧
Cs
T
Vout ( t ) = L−1 ⎨ A( s ) ⋅
⋅ Vstep ⋅ 0 ⋅ 1 − e −t / τ
⎬=−
s ⎭
Cf
1 + T0
⎩
(
Ideal
Response
B. Murmann
)
τ=
Static
Error
Dynamic
Error
⇒ ε0
⇒ εd
1
ωc
EE 214 Lecture 21
5
Graphical Illustration
Static
Error
Dynamic
Error (t)
1
Vout/Vout,ideal
0.8
0.6
0.4
0.2
0
0
2
4
6
8
10
t/τ
B. Murmann
EE 214 Lecture 21
6
261
Design Considerations (1)
•
Need large DC loop gain for small static error
– |ε0| ~ 1/T0
– E.g. need T0>1000 for better than 0.1% precision
•
Need small τ (large bandwidth) for fast settling
•
Can define settling time ts based on tolerable dynamic error
− ε d ,tol = −e −t s / τ
t s = −τ ⋅ ln(ε d ,tol )
ts = −
B. Murmann
1
ωc
⋅ ln(ε d ,tol )
EE 214 Lecture 21
7
Design Considerations (2)
•
εd,tol
ts/τ
1%
4.6
0.1%
6.9
0.01%
9.2
10-6
13.8
Going from 1% dynamic precision to 10-6 necessitates only ~3x
increase in settling time
B. Murmann
EE 214 Lecture 21
8
262
Design Considerations (3)
•
A switched capacitor circuit operates in two clock phases
•
Fitting the required number of time constants within ½ period
lets us relate fCLK to a minimum bandwidth requirement
ts = −
B. Murmann
fc
1
1 1
⋅ ln(ε d ,max ) <
2π ⋅ f c
2 f CLK
f CLK
>−
1
π
ln(ε d ,max )
ε
fc/fCLK
1%
1.5
0.1%
2.2
0.01%
2.9
10-6
4.4
EE 214 Lecture 21
9
How Fast Can We Go?
•
fc cannot be larger than about 1/3 of the nondominant amplifier
pole frequency (stability)
– In a cascoded amplifier, the nondominant pole occurs at a
frequency around fT/3
• Assuming that two junction caps equal to Cgg are present
– At a reasonable bias, the NMOS transit frequency in our
technology is roughly 8GHz (nominal process and
temperature)
•
Putting all of this together, and assuming that we'll need to settle
to within 0.01% (~9 time constants), we have
f CLK ,max =
B. Murmann
f
1 fT
⋅
≅ T = 266 MHz
2.9 9
30
EE 214 Lecture 21
10
263
Practical Aspects
•
In practice, it is hard to achieve fCLK> fT/50
– Amplifier topology restriction
• E.g. sometimes forced to use PMOS in signal path
– Restrictions on power dissipation
• Near the technology limits, power tends to grow out of bounds
– Timing overhead to produce non-overlapping clocks
• Have somewhat less than half clock cycle to settle
– Other transient effects
• We'll look at some of those next…
B. Murmann
EE 214 Lecture 21
11
Impact of Non-Dominant Pole
[Yang, IEEE TCAS 3/1990]
(see also appendix)
B. Murmann
EE 214 Lecture 21
12
264
Simulation Example
•
Using simple single stage (single pole) OTA
•
Parameters
– Cs=Cf=500fF, CL=10pF, β=0.48, Gm=1mS, GmRo=85, Vidstep=-10mV
B. Murmann
EE 214 Lecture 21
13
Result
τ=
1 C L + (1 − β )C f
= 21ns
Gm
β
Vod , final = −Vidstep
β ⋅ Gm R0
= 9.76 mV
1 + β ⋅ Gm R0
1000
-V
Voltage [mV]
800
id
Vod (simulation)
V
600
od
(theory?)
400
200
0
0
B. Murmann
50
100
Time [ns]
EE 214 Lecture 21
150
200
14
265
Another Run
•
Changed CL from 10pF to 300fF
10
Voltage [mV]
-V
V
5
od
(simulation)
Vod (theory?)
0
-5
0
•
id
5
10
15
20
25
Time [ns]
What's this ?
B. Murmann
EE 214 Lecture 21
15
Capacitive Feedforward
•
In the first instant after the input step has been applied, the
output is completely determined by capacitive voltage division
•
Half circuit during initial transient:
Vodstep
Vidstep
B. Murmann
=
Cs
Cs + Cin +
C f CL
⋅
C f + CL
EE 214 Lecture 21
Cf
C f + CL
16
266
Analysis
•
Can analyze this effect in two (equivalent) ways
– Using capacitive divider to find new starting point of
exponential
– Using inverse Laplace transform of A(s) with feedforward
zero included
•
Recall from lecture 14, that A(s) is more precisely given by
s
1−
Cs
z ⋅ T0
A(s ) = −
C f 1 − s 1 + T0
p
B. Murmann
z=
Gm
Cf
p=−
βGm
C L + (1 − β )C f
EE 214 Lecture 21
17
New Result
Vstep ⎫
⎧
Cs
T
Vod ( t ) = L−1 ⎨ A( s ) ⋅
⋅Vidstep ⋅ 0
⎬=−
s ⎭
Cf
1 + T0
⎩
1−
•
CL + C f
p CL + (1 − β )C f + βC f
=
=
=
z
C L + (1 − β )C f
C L + (1 − β )C f
New
1
1− β
Cf
C f + CL
For our example:
1
= 1.4
500 fF
1 − 0.48
500 fF + 300 fF
•
⎛ ⎡
⎞
p⎤
⋅ ⎜⎜ 1 − ⎢1 − ⎥ e −t / τ ⎟⎟
z⎦
⎝ ⎣
⎠
⇒ Vod ( t = 0 ) ≅ 10 mV (1 − 1.4 ) = −4 mV
Good agreement with simulation
B. Murmann
EE 214 Lecture 21
18
267
New Settling Time
ts = −
⎛
⎡
C f ⎤ ⎞⎟
⋅ ln⎜ ε d ,tol ⎢1 − β ⋅
⎥
ωc ⎜⎝
C f + CL ⎦⎥ ⎟⎠
⎣⎢
1
<1
•
Settling time for given precision increases due to feedforward,
since the settling range is artificially enlarged
•
E.g. in our simulation example, the time to settle within 0.1%
dynamic error increases from 6.9τ to 7.3τ
– Not extremely significant, especially when β is low and CL is
at least comparable to Cf
B. Murmann
EE 214 Lecture 21
19
Appendix
"Bandwidth and Settling of a Two-Pole Feedback Amplifier"
B. Murmann
EE 214 Lecture 21
20
268
Bandwidth and Settling of a Two-Pole Feedback Amplifier
Boris Murmann
11/11/2006
Single Pole Amplifier (for reference)
β ⋅ T0
T( s) =
1−
T( s) =
s
β ⋅ Gm⋅ R
1
=
1 + sRC
1
β ⋅ Gm⋅ R
p1
C
β ⋅ Gm
1
T( s) =
For large DC loop gain:
+ s⋅
s⋅
C
β ⋅ Gm
Unity gain frequency of T(s)
1
j ⋅ ω c1⋅
=1
C
ω c1 = β ⋅
β ⋅ Gm
Gm
C
Phase margin
PM1 = 180deg + arg⎛⎜
⎞
⎟
⎟
ω c1 ⎟
⎠
1
⎜j
⎜
⎝
PM1 := 90deg
ω c1
Closed loop 3-dB frequency (neglecting feedforward effects)
(
T jω3dB
(
)
1 + T jω3dB
)
=
1
1
2
1+j
ω 3dB
=
1
1
2
=
⎛ ω 3dB ⎞⎟
1+⎜
⎜⎝ ω c1 ⎟⎠
ω c1
2
Linear settling time
−
Step response:
u( t) = 1 − e
t
τ
τ =
−
Dynamic error at t=ts
εd = e
1
ω c1
ts
τ
( )
ts = −τ ⋅ ln ε d
1
2
ω 3dB = ω c1
269
Two- Pole Amplifier
1
T( s) =
s
ω c1
⋅ ⎛⎜ 1 −
s
⎞
⎠
p2 ⎟
⎝
Unity gain frequency of T(s)
1
j
⎛
ω c1 ⎜
⎝
ω c2
ω c2 ⎞
⋅⎜1 + j
=1
m=
⎟
⎠
ω p2 ⎟
2
m⋅ 1 +
⎛ m⎞ = 1
⎜n⎟
⎝ ⎠
m( n ) :=
ω c2
ω c1
1
2
n=
2
ω p2
ω c1
⋅ 2 ⋅ n ⋅ n + 4 − 2n
2
0.9
m( n)
0.8
0.7
1
2
3
4
5
n
Non-dominant pole moves crossover point to slightly lower frequencies (as one would predict
from first order model). E.g. for a nondominant pole with a frequency of ~three times the
crossover frequency (n=3), the crossover shifts to ~0.95*wc1.
Phase margin
PM2 = 180deg + arg⎡⎢
⎢j
⎢
⎣
⎤ = 180 ⋅ deg − 90deg + arg⎛
⎥
⎜
ω c2 ⎛
ω c2 ⎞
⎥
⎜1 +
⎟
⋅⎜1 + j
⎥
⎜
ω c1 ⎜
ω p2 ⎟
⎝
⎠⎦
⎝
1
⎞
⎟
ω c2
⎟
j⋅
ω p2 ⎟
⎠
1
k=
ω p2
ω c2
270
PM2( k ) := 90deg − atan⎛⎜
1⎞
⎟
⎝k⎠
90
80
70
PM2( k) ⋅
180
π
60
50
40
1
2
3
4
5
6
65
70
7
8
k
1
k ( PM2) :=
⎛ π − PM2⎞
⎟
⎝2
⎠
tan⎜
6
5
4
π ⎞
⎛
k⎜PM⋅
⎟
180 ⎠
⎝
3
2
1
45
50
55
60
PM
75
80
271
Closed loop 3-dB frequency (neglecting feedforward effects)
A( ω ) =
1
ω
1+j
A( w , n ) :=
ω c1
w=
⎛
⎞
⎟
ω p2
⎠
ω
⋅⎜1 + j
⎝
1
1 + j w⋅ ⎛⎜ 1 + j
⎝
c( w) :=
w⎞
ω
n=
ω c1
ω p2
ω c1
1
2
⎟
n⎠
1.5
A ( w , 1)
A ( w , 2)
1
A( w , 100)
c( w )
0.5
0
0.1
1
10
w
1
1 + j r⋅ ⎛⎜ 1 + j
⎝
r⎞
⎟
=
1
r=
2
ω 3dB
n=
ω c1
n⎠
ω p2
ω c1
2
2⎞
⎛
⎜ 1 − r ⎟ + r2 = 2
n⎠
⎝
r( n ) :=
1
2
2
2
⋅ 4⋅ n − 2⋅ n + 2⋅ n⋅ n − 4⋅ n + 8
1.4
r( n)
1.3
1.2
1.1
2
4
6
8
10
n
For wp2=3*wc1, closed loop bandwidth is about 35% higher than first order prediction
272
Linear settling time
2
U( s) =
1
s
1 U( s) = k ⋅ ω c1
= ⋅
s
2
2
k ⋅ ω c1 + s⋅ k ⋅ ω c1 + s
1
⋅
1+
2
s
ω c1
⋅ ⎛⎜ 1 +
⎝
⎞
ω p2 ⎟
⎠
s
s1
2
k ⋅ ω c1 + s⋅ k ⋅ ω c1 + s = 0
ω c1
=−
k
2
2
+
s2
⎛k⎞ − k
⎜2⎟
⎝ ⎠
k ⎛
s1n( k ) := − ⋅ ⎜ 1 +
2 ⎝
1−
k=
ω c1
4⎞
=−
k
2
ω p2
ω c1
2
⎛k⎞ − k
⎜2⎟
⎝ ⎠
−
k ⎛
s2n( k ) := − ⋅ ⎜ 1 −
2 ⎝
⎟
k⎠
2
1
U( s) =
s
⋅
k ⋅ ω c1
a = s1n⋅ ω c1
(s − s1n⋅ ω c1)⋅ (s − s2n⋅ ω c1)
2
a⋅ t
b⋅ t ⎞
⎛ 1
e
e
⎟
+
+
⎝ a⋅ b a( a − b ) b ( b − a) ⎠
tn = t⋅ ω c1
u ( t) = k ⋅ ω c1 ⋅ ⎜
(
)
k⋅ e
k
b = s2n⋅ ω c1
s 1n( k) ⋅ tn
k⋅ e
s2n( k) ⋅ tn
u tn , k :=
+
+
s1n( k ) ⋅ s2n( k )
s1n( k ) s1n( k ) − s2n( k )
s2n( k ) s2n( k ) − s1n( k )
( )
u( tn , 2)
u( tn , 3)
u tn , 1
(
)
(
1.1
1
0.9
0
2
4
6
tn
8
10
)
1−
4⎞
⎟
k⎠
273
(
)
(
)
ε d t n , k := u tn , k − 1
(
)
(
t sn ε dspec , k :=
)
t s ← −ln ε dspec + 10
(
)
while ε d t s , k < ε dspec
t s ← ts − 0.01
return
(
ts
−ln ε dspec
)
k := 1 , 1.1 .. 8
Relative settling time versus k=wp2/wc1 for various dynamic error specs
1.4
1.2
1
tsn( 0.01% , k)
tsn( 0.1% , k)
0.8
tsn( 1% , k)
1
0.6
0.4
0.2
0
1
2
3
4
5
k
6
7
8
274
k ( PM2) :=
1
PM := 50 , 51 .. 90
⎛ π − PM2⎞
⎟
⎝2
⎠
tan⎜
1.4
1.2
π ⎞⎞
⎛
⎛
tsn⎜0.01% , k⎜PM⋅
⎟⎟
180
⎝
⎝
⎛
⎝
⎛
⎝
tsn⎜0.1% , k⎜PM⋅
⎠⎠
π
⎞⎞
⎟⎟
180 ⎠⎠
π ⎞⎞
⎛
⎛
tsn⎜1% , k⎜PM⋅
⎟⎟
180
⎝
⎝
⎠⎠
1
0.8
0.6
1
0.4
0.2
0
50
60
70
80
PM
Conclusion: Try to target phase margin between 70...75 degrees; amplifier then
settles ~30% faster than single pole system (ignoring second order effects).
90
275
Lecture 22
Slewing
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 22
1
Overview
•
Reading
– 9.6.1, 9.6.2, 9.6.5 (Slew Rate)
•
Introduction
– Today we'll complete our discussion of transient behavior in
OTA circuits with capacitive feedback. Aside from the
feedforward artifact we've discovered last time, there exists
another effect called "slewing". Whenever the differential
input pair is driven outside its "linear range", the differential
output current is limited by the available tail bias. In typical
switched capacitor circuits, slewing can cause a significant
speed reduction, and is therefore an important effect to
consider.
B. Murmann
EE 214 Lecture 22
2
276
Simulation Example from Last Lecture
•
•
Using simple single stage OTA (See lecture 15, slide 3)
Parameters
– Cs=Cf=500fF, CL=10pF, β=0.48, Gm=1mS, GmRo=85
B. Murmann
EE 214 Lecture 22
3
Another Simulation
•
Set Vidstep=-1V (CL=10pF ⇒ insignificant feedforward to output)
1000
Voltage [mV]
800
-Vid
Vod (simulation)
600
Vod (theory?)
400
200
0
0
•
50
100
Time [ns]
150
200
What causes this discrepancy ?
B. Murmann
EE 214 Lecture 22
4
277
Capacitive Divider at OTA Input
•
Half circuit during initial transient:
Vxdstep = Vidstep
•
Cs
Cs + Cin +
≅ −1V
C f CL
500 fF
= −480 mV
500 fF + 40 fF + 500 fF
C f + CL
Initially -480mV across differential pair input!
B. Murmann
EE 214 Lecture 22
5
Differential Pair Characteristics
•
Differential output current begins to saturate ~|Vid| > 1.4·2/(gm/ID)
•
Beyond this point, current will be much less than⎝ that predicted
by linear model (slope at origin)
1
Slope = 1
Iod/ITAIL 0
1
2
1
-√2
B. Murmann
0
1
√2
Vid/(2/[gm/ID])
EE 214 Lecture 22
2
6
278
Differential Pair Input Voltage vs. Output Current
0
Vxd [mV]
-100
-200
"Slewing"
"Linear Settling"
-300
-1.4·2/gm/ID)
-400
-500
0
50
100
150
100
150
Time [ns]
Diff. pair I od [μA]
0
-100
-200
-300
0
50
Time [ns]
B. Murmann
EE 214 Lecture 22
7
Slewing
•
During "slewing", the amplifier drives its output with a constant
current (equal to tail bias)
•
The slewing behavior ends when |Vid| has become smaller than
about 1.4·(2/gm/ID)
– This is the point when the differential pair re-enters its "linear
region"
– Hence, the remaining portion of the settling is often called
"linear settling"
• Even though the output voltage really settles with a (1-et/τ)
relationship
•
The total settling time of the amplifier in presence of slewing can
be calculated as shown in the following derivation
B. Murmann
EE 214 Lecture 22
8
279
Slew Rate
•
In order to find the time it takes to complete slewing, we can first
calculate the "ramp speed" at which the output changes
– This quantity is called "Slew Rate" (SR)
SR =
B. Murmann
dVod I TAIL
I TAIL
=
=
dt
CLtot CL + (1 − β )C f
EE 214 Lecture 22
9
Slewing Time
•
The input of the differential pair changes at a rate equal to β·SR,
where β is given by the usual capacitive feedback divider
•
Hence, the time it takes to complete slewing is given by
t slew =
•
Vxstep − 2.8 / (g m / I D )
β ⋅ SR
In our example, we have
SR =
t slew =
B. Murmann
I TAIL 200 μA
V
≅
= 20
C Ltot 10 pF
μs
480 mV − 280 mV
= 21ns
V
0.48 ⋅ 20
μs
EE 214 Lecture 22
10
280
Subsequent Linear Settling
•
Once slewing is completed, the differential output voltage is
Vod ,slew = t slew ⋅ SR = 420 mV
•
The final settling value in our example is roughly 1V
– Almost half way there after slewing
•
This means that the dynamic error budget for the remaining
settling portion has increased
– E.g. if we wanted to settle within 0.1% of the final value
(~1V), we only need to complete the remaining transient to
within 0.1%·1V/0.58V=0.17%
– Not a very big win, usually a negligible change in the number
of required time constants
• 0.1%→6.9τ, 0.17% → 6.4τ
B. Murmann
EE 214 Lecture 22
11
Complete Expression for Settling Time
t s = t slew + tlin ≅
SR =
β ⋅ SR
I TAIL
I TAIL
=
C Ltot C L + (1 − β )C f
Vxdstep = Vidstep
•
Vxstep − 2.8 / ( g m / I D )
Cs
Cs + Cin +
C f CL
β=
− τ ln(ε d ,tol )
Cf
C f + Cs + Cin
≅ Vidstep
C f + CL
Cs
Cs + Cin + C f
<1
Note that circuits with large closed loop gain tend to slew less
– Since Vidstep cannot be larger than Voutputswing/Gain
– E.g. Voutputswing=2V, Gain =8 ⇒ Vxdstep < Vidstep < 250mV
• Won’t slew at all if gm/ID < 2.8/250mV= 11.2V-1
B. Murmann
EE 214 Lecture 22
12
281
Slewing in a Two-Stage OTA (1)
•
Nodes V1p and V1m stay roughly constant
– These nodes move at a rate equal to the slew rate Vop, Vom
divided by ~ intrinsic gain of second stage
– Second stage acts as "integrator"
B. Murmann
EE 214 Lecture 22
13
Slewing in a Two-Stage OTA (2)
•
Must design circuit such that slew rate is limited by ITAIL
– IB2 limitation would cause asymmetric slewing
• One branch slew limited by IB2, other slew limited by ITAIL
– Asymmetric slewing causes CM shift
• Causes slow transients due to slow CMFB
B. Murmann
EE 214 Lecture 22
14
282
Slewing in a Two-Stage OTA (3)
•
The maximum slew rate at which output Vom can move down
SR− ,MAX =
•
d (Vom − V1 p )
dt
⎧I
/ 2 I B2 / 2 ⎫
= min⎨ TAIL
,
⎬
Cc + C Ltot ⎭
⎩ Cc
The maximum slew rate at which output Vop can move up
SR+ ,MAX =
•
d ( Vop − V1m )
dt
I
/2
= TAIL
Cc
To make slew rates equal, we need
ITAIL
I B2
≤
Cc
Cc + C Ltot
B. Murmann
EE 214 Lecture 22
15
Slewing in "Continuous Time" Circuits
•
Slewing is not only an issue in switched capacitor circuits, it can
also limit the large signal performance of continuous time circuits
•
Example:
vo (t ) = Vˆo sin (ωt )
iout
Vin
Vout
CL
iout (t ) = CL
dvo
dt
= ω ⋅ C L ⋅ Vˆo cos(ωt )
îout = ω ⋅ C L ⋅V̂o
•
At large frequency and/or amplitude, the peak output current needed
can exceed the maximum current available from the amplifier
B. Murmann
EE 214 Lecture 22
16
283
Resulting Waveform
B. Murmann
EE 214 Lecture 22
17
Design Considerations
•
When slewing is an issue, it can be mitigated by biasing the
relevant transistors at lower gm/ID
– Increase ID, keep gm constant
– Slewing performance improves, because of larger ID and
also because the differential pair input range increases
(2.8/[gm/ID])
– Small signal performance remains virtually unchanged or
improves if fT is a limiting factor (since fT increases)
– Issue: Lower gm/ID means higher power consumption
B. Murmann
EE 214 Lecture 22
18
284
How to Incorporate Slewing in Design Flow
•
Set up a spreadsheet for small signal design as usual
– Settling time requirement translates into minimum "linear"
bandwidth spec (based on linear analysis, without slewing)
•
Introduce a bandwidth spec scale factor K≥1 in your design
script
– If the circuit slews, we will need more bandwidth than
predicted from linear analysis
•
Perform design optimization as usual, begin with K=1
•
Calculate slewing time, add to linear settling time
– Done if tslew=0
•
Increase K until design meets settling time spec
– In the process, you may consider to optimize gm/ID values to
minimize power in presence of slewing
B. Murmann
EE 214 Lecture 22
19
285
Lecture 23
Feedback and Port Impedances
OTA Variants, CMFB Implementation
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 23
1
Overview
•
Reading
– 8.8.2 (Closed-Loop Impedance Formula using Return Ratio)
– 6.7 (MOS Active-Cascode Amplifiers)
– 12.5 (CMFB Circuits)
– 12.6 (Fully Differential Amplifiers)
– 9.4.4 (Compensation of Single Stage CMOS OTAs)
•
Reference
– K. Bult, G.J.G.M. Geelen, "A fast-settling CMOS op-amp for SC circuits with
90-dB DC gain," IEEE J. Solid-State Circuits, Dec. 1990, pp. 1379 – 1384.
•
Introduction
– In order to complete our framework for feedback circuit analysis, we will
study the effect of feedback on port impedances in today's lecture. A useful
analytical tool for quick calculations is Blackman's Impedance Formula,
which allows us to find port impedances based on a superposition of simple
sub-analyses.
– Next, we survey common OTA architectures and associated implementation
aspects. Most OTAs used in practice are derivatives of the basic single- or
two-stage topology, with add-ons such as simple cascodes or gain boosted
cascodes. The survey concludes with a brief analysis of the most commonly
used common mode feedback implementations.
B. Murmann
EE 214 Lecture 23
2
286
Using Feedback to Modify Port Impedances
•
Our initial motivation for using feedback was to build precise
gain elements that are insensitive to gmR variations
•
In addition, feedback can be used to increase/decrease port
impedances
– In fact, we have already seen one example of such behavior
• Closed loop bandwidth of OTA feedback amplifier was (1+T)
higher than open loop bandwidth
• This means the impedance seen by the load capacitor must
have dropped in presence of feedback
•
We can calculate the port impedances of arbitrary feedback
circuits using "Blackman's Impedance Formula"
– Based on loop gain calculations
– Extremely useful and easy to use
B. Murmann
EE 214 Lecture 23
3
Blackman's Impedance Formula
Z port = Z port (k = 0 ) ⋅
1 + T ( port shorted )
1 + T ( port open )
1. First, find port impedance with feedback loop broken
– E.g. set gm=0
2. Calculate loop gain in circuit with port under consideration shorted
3. Calculate loop gain in circuit with port under consideration open
•
In many cases, either the "shorted" or "open" loop gain is zero
B. Murmann
EE 214 Lecture 23
4
287
Example 1
Rout (k = 0 ) = Rout (av = 0 ) =
1
gm
T ( port shorted ) = 0
T ( port open ) = av
Rout =
B. Murmann
1 1
g m 1 + av
EE 214 Lecture 23
5
Example 2: Shunt-Shunt Stage
Rin ( g m = 0 ) = RF + ro
Rout (g m = 0 ) = ro
T (input shorted ) = 0
T (output shorted ) = 0
T (input open ) = g m ro
T (output open ) = g m ro
Rin = (RF + ro )
B. Murmann
1
1 ⎛ RF ⎞
⎟
⎜1 +
≅
1 + g m ro g m ⎜⎝
ro ⎟⎠
EE 214 Lecture 23
Rout = ro
1
1
≅
1 + g m ro g m
6
288
Example 3: Active Cascode
•
Also referred to as "Regulated Cascode" or "Gain Boosting"
technique
Rout (av = 0 ) ≅ r01 ⋅ g m 2 ro 2
T ( port open ) = 0
T ( port shorted ) ≅ av
gm2
≅ av
g m 2 + g mb 2
Rout ≅ r01 ⋅ g m 2 ro 2 ⋅ (av + 1)
≅ r01 ⋅ g m 2 ro 2 ⋅ av
B. Murmann
EE 214 Lecture 23
7
Basic Implementation
•
Can use simple CS stage as
auxiliary amplifier
– Issue: Costs headroom
•
See literature for more advanced
implementations
– E.g. using fully differential
folded cascode amplifier as an
auxiliary amplifier
Rout ≅ r01 ⋅ g m 2 ro 2 ⋅ g m 3 ro 3
B. Murmann
EE 214 Lecture 23
8
289
Properties
•
Must compensate local feedback loop such that its crossover
frequency ωc occurs before non-dominant pole at source of M2.
•
Consideration of the total impedance at the output node
(including CL) shows that gain boosting introduces a pole zero
doublet around ωc
– Can result in slow step response, if not designed carefully
– See Bult, JSSC 12/1990 for design considerations
•
In typical designs, "gain boosting" adds about 20-30% to the
total power dissipation of an OTA
B. Murmann
EE 214 Lecture 23
9
OTA Variants: Telescopic OTA
B. Murmann
EE 214 Lecture 23
•
Approximately same gain
as two stage amplifier,
with only two current legs
– Maximum power
efficiency
•
Issue: Low swing
– Especially if input and
output common mode
cannot be chosen
freely
10
290
Biasing
B. Murmann
•
Typically use at least 20% of tail
current in auxiliary biasing branch
– Must avoid slow recovery of
this node during transients
•
Can use several device in series
to implement MB1
– "1/3rd or 1/5th" device
– Helps avoid extremely small
width for MB1
EE 214 Lecture 23
11
Folded Cascode OTA
B. Murmann
EE 214 Lecture 23
•
Large input
common mode rage
•
Slightly improved
output range
•
Folding adds power
dissipation
12
291
Compensation
AC Half Circuit for Telescopic OTA
AC Half Circuit for Folded Cascode OTA
•
Nondominant pole ωp2=gm1a/Cp
– Cp=Cgs1a+Cj1a+Cj1
•
Easy to compensate
– Make k·β·gm1/CL < ωp2
– k=3 for 72° phase margin
B. Murmann
EE 214 Lecture 23
13
Current Mirror OTA
k
1
1
•
Gm=k·gm1,2, ωp2 ~ ωT/(1+k)
•
Large swing
•
Good for low speed, low power applications
– See e.g. Yao, JSSC 11/2004
B. Murmann
EE 214 Lecture 23
k
14
292
All NMOS Signal Path (1)
[Feldman et al., JSSC 10/1998]
•
Capacitive level shift allows NMOS in second stage
B. Murmann
EE 214 Lecture 23
15
All NMOS Signal Path (2)
[Yang et al., JSSC 12/2001]
•
Differential pair and separate common mode feedback in
second stage
B. Murmann
EE 214 Lecture 23
16
293
Gain Boosted Gain Boosters
[Chiu et al., ISSC 2004]
•
Gain ~ gmro6, design achieved av0=130dB in 0.18μm technology
B. Murmann
EE 214 Lecture 23
17
Common Mode Feedback
•
Implementation aspects
– How to sense
– How to compare to desired value
– How to provide a "knob" for adjusting Voc
B. Murmann
EE 214 Lecture 23
18
294
Knob
•
Typically generate ~80% of tail current with fixed bias, leave
remaining 20% as tuning range for CMFB loop
B. Murmann
EE 214 Lecture 23
19
Comparison Circuit
•
Low frequency loop gain T0 ≅ 0.25·gmx·rop1 · gmp2/gmx
– Loop will control Voc more accurately if Mp1 is cascoded
B. Murmann
EE 214 Lecture 23
20
295
Sensing
•
•
Using a resistive divider may destroy differential gain
Solutions
– Use source followers to drive divider (headroom issue)
– Purely capacitive sensing
B. Murmann
EE 214 Lecture 23
21
CMFB Implementation Example
•
Circuit uses switched
capacitors (CM) to set the
voltage across sensing
capacitors (CCM)
[Feldman et al., JSSC 10/1998]
B. Murmann
EE 214 Lecture 23
22
296
"Passive" CMFB (1)
•
•
During φ1: Initialize voltage across Ccmfb to Voc,desired - VB
During φ2: Activate feedback loop
– If Voc>Voc,desired, Vcntrl becomes >VB and lowers Voc
B. Murmann
EE 214 Lecture 23
23
"Passive" CMFB (2)
•
OTA cannot be used during φ1, because the common mode
feedback mechanism is inactive
– Often not a problem in switched capacitor circuits, where the
OTA is active only during one half-cycle
•
Can use switched capacitor scheme shown on slide 22 to
enable uninterrupted common mode feedback
•
Unfortunately, this simple circuit cannot be used if an additional
inversion is needed in the common mode feedback loop
– E.g. won't work for a two-stage OTA that uses a single
common mode feedback loop (see e.g. slide 9, lecture 18)
– Will work for the two-stage OTA with separate CMFB loops
as shown on slide 16
B. Murmann
EE 214 Lecture 23
24
297
Common Mode Half Circuit
•
Low frequency loop gain:
g
T0 ≅ mx rop ⋅
2
•
Loop crossover frequency
ωc ≅
•
1
2
Ccmfb
g mx
Cx
Ccmfb ⋅ 0.5C x
Ccmfb +
C +
2 L Ccmfb + 0.5C x
Nondominant pole
ω p2 ≅
B. Murmann
Ccmfb
C
Ccmfb + x
2
g mn
Cy
EE 214 Lecture 23
25
Design Considerations
•
The required bandwidth of the common mode loop strongly
depends on the amount of expected imbalance, common mode
transients or ac components
– In an ideal world, the common mode is not affected by the
signal and hence stays constant
• In this case, the bandwidth of the CMFB loop is unimportant
•
For robustness in practical implementations, the bandwidth of
the common mode loop is often chosen to be about 30% of the
differential signal path bandwidth
– In a typical switched capacitor circuit with 10 time constants
differential settling, this means that the common mode has
about 3 time constants to settle
• Enough time to remove 95% of common mode disturbance
B. Murmann
EE 214 Lecture 23
26
298
Lecture 24
OTAs with Single Ended Outputs
Output Stage Examples
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 24
1
Overview
•
Reading
– 4.3.5 (Differential Pair with Current Mirror Load)
– 6.3.3 (Systematic Offset in Two-Stage Amplifier)
– Chapter 5 (Output Stages)
•
Introduction
– This lecture concludes our discussion of amplifier implementations
by looking at a number of missing bits and pieces. First, we will
discuss subtleties of OTA implementations that use a single ended
output. While rarely used in the signal path of high performance
integrated circuits, OTAs with single ended outputs can be useful
as auxiliary amplifiers in biasing circuits. Finally, we will take a brief
look at output stages that are suitable for driving resistive loads.
While most on-chip loads tend to be capacitive in nature, low
impedance drivers are often needed when interfacing to off-chip
components and wire lines.
B. Murmann
EE 214 Lecture 24
2
299
Single Ended OTA
•
Current mirror performs "differential to single-ended conversion"
B. Murmann
EE 214 Lecture 24
3
Mirror Doublet
•
Half of the output current comes directly from the differential
pair, the other half goes through a current mirror with finite
bandwidth
– Result: Pole-zero doublet
⎞
⎛
⎟
⎜
1 1 1 ⎟
⎜
io = g m vid +
s⎟
⎜2 2
1− ⎟
⎜
p⎠
⎝
s
1−
2p
= g m vid
s
1−
p
•
In the circuit on the previous slide p ≅ -ωTp/2 (neglecting
junctions)
B. Murmann
EE 214 Lecture 24
4
300
Unity Gain Buffer
•
Closed loop gain ≅ 1
•
Output impedance ≅ 1/gm
•
Output range equal to common mode input range
– Advantageous to use a folded cascode OTA architecture
B. Murmann
EE 214 Lecture 24
5
Inverting Amplifier
•
Cx = ?, T(s) = ?
– Big mess…
B. Murmann
EE 214 Lecture 24
6
301
Two-Stage OTA with Single Ended Output
B. Murmann
EE 214 Lecture 24
7
Systematic Offset
B. Murmann
EE 214 Lecture 24
•
No offset if VGS6=VGS3
balances the current in
the output branch
•
Input referred
systematic offset is
(VGS6,balance-VGS3)/av1,
where av1 is the gain of
the first stage
8
302
Output Stages
•
Needed to drive resistive loads (low R)
– Integrated continuous time RC-filters
– Off-chip resistive loads
– Line drivers
• E.g. twisted pair (Ethernet, ISDN, ADSL)
•
Solutions
– Use OTA + source follower output stage
• Swing issue
– Use OTA + "low gain" common source stage
• E.g. make gm·RL~1
• E.g. 20mS·50Ω =1, IBIAS=20mS·10V-1 = 2mA
– "Sophisticated" output stages
• Examples on following slides
B. Murmann
EE 214 Lecture 24
9
Output Stage Nomenclature (1)
•
Class-A
– Output devices conduct for entire cycle of output sine wave
– E.g. source follower with constant current source bias
•
Class-B
– Output devices conduct for ≅50% of sine wave cycle
– E.g. back-to back PMOS/NMOS source followers
• NMOS and PMOS each conduct for about one half cycle
• No quiescent current during zero crossing of sine wave
•
Class-AB
– Output devices conduct for >50%, but <100% of cycle
– E.g. a simple inverter
• Problem: How to set quiescent current around zero crossing
• Needs local or global feedback to mitigate nonlinearity
B. Murmann
EE 214 Lecture 24
10
303
Output Stage Nomenclature (2)
Class-A
Class-B
Class-AB
B. Murmann
Class-C
EE 214 Lecture 24
11
Selected References on Output Stages
•
D. M. Monticelli, "A quad CMOS single-supply op amp with railto-rail output swing," IEEE J. Solid-State Ckts., pp. 1026-1034,
Dec. 1986.
•
R. Hogervorst, J. P. Tero, R. G. H. Eschauzier, and J. H.
Huijsing, "A compact power efficient 3 V CMOS rail-to-rail
input/output operational amplifier for VLSI cell libraries," IEEE J.
Solid-State Ckts., pp. 1505 - 1513, Dec. 1994.
•
G. Palmisano, G. Palumbo, and R. Salerno, "CMOS Output
Stages for Low-Voltage Power Supplies," IEEE Trans. Ckts. and
Syst. II, pp. 96-104, Feb. 2000.
B. Murmann
EE 214 Lecture 24
12
304
Class-AB Output Stage
[Hogervorst]
B. Murmann
EE 214 Lecture 24
13
Current [A]
Output Current (Vo=0)
Iout
•
Output transistors
never turn off
•
Quiescent current
set by transistor
ratios
•
Large drive
capability
ID(M26)
ID(M25)
Iin1=Iin2 [A]
B. Murmann
EE 214 Lecture 24
14
305
Complete OpAmp
[Hogervorst]
B. Murmann
EE 214 Lecture 24
15
Low Voltage Variant
<Vtn+|Vtp|
[Palmisano]
B. Murmann
EE 214 Lecture 24
16
306
Complete OpAmp
<Vtn+|Vtp|
[Palmisano]
B. Murmann
EE 214 Lecture 24
17
A Note on General Purpose OpAmps
•
Issue: Compensation
– The designer of a general purpose OpAmp does not know
anything about the feedback network of the particular
application
– General purpose OpAmps are typically compensated for the
"worst case", i.e. unity feedback configuration
• Tends to be wasteful, since much less compensation is needed
for smaller return factors
– Some general purpose OpAmps provide an external pin to
let the user decide on the required compensation capacitor
B. Murmann
EE 214 Lecture 24
18
307
Lecture 25
Supply Insensitive Biasing
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 25
1
Overview
•
Reading
– 4.4.2 (Supply Insensitive Biasing)
•
Introduction
– Throughout this course, we have ignored the question on
how to generate the reference bias currents used in most of
our circuits. Today, we will examine a variety of current
reference implementations, with primary focus on supply
independence.
B. Murmann
EE 214 Lecture 25
2
308
Poor Man's Bias
I OUT ≅ I IN =
•
VDD − Vt − VOV
R
Issue: Current is essentially proportional to VDD
– E.g. if VDD varies by X%, bias current will roughly vary by the
same amount
B. Murmann
EE 214 Lecture 25
3
"Vt" Referenced Bias
Vt +
I OUT =
•
VGS 1 Vt + VOV
≅
≅
R2
R2
By using a sufficiently large device,
we can make VOV << Vt, and achieve
I OUT ≅
•
B. Murmann
2 I IN
W
μCox
L
R2
Vt
R2
Question: By how much will IOUT
change given some variation in VDD?
EE 214 Lecture 25
4
309
Sensitivity
•
The sensitivity of a parameter y to a change in parameter x can
be approximately found using
Δy / y ∂y / y x ∂y ! y
≅
=
= Sx
Δx / x ∂x / x y ∂x
•
In our case, we are looking for
OUT
IN
SVIDD
= SVIDD
⋅ S IIINOUT
≅ 1⋅
≅
•
I IN ∂I OUT
⋅
I OUT ∂I IN
1 VOV 1
2 Vt + VOV 1
e.g .
1
0.1V
= 7.1%
2 0.6V + 0.1V
Not bad, but also not all that great…
B. Murmann
EE 214 Lecture 25
5
BJT Version
I OUT =
OUT
SVIDD
B. Murmann
VBE 1
1 kT ⎛ I IN ⎞
⎟
ln⎜
=
R2
R2 q ⎜⎝ I S ⎟⎠
kT
q
=
VBE
EE 214 Lecture 25
e.g .
26 mV
= 3.7%
700 mV
6
310
Stability
T (s ) ≅ g m1 R1 ⋅
g m 2 R2
1
1
⋅
⋅
s
s
1 + g m 2 R2 1 +
1+
ω p1
•
ω p2
Loop gain greater 1 at low
frequencies, two poles
– Means that we must make one of
the poles dominant to guarantee
sufficient phase margin
• E.g. use large capacitance to
ground at drain of T1
B. Murmann
EE 214 Lecture 25
7
Self Biasing
•
In the above discussed bias generator circuits, the supply
sensitivity is still fairly high, because IIN is essentially directly
proportional to VDD
•
Idea: Mirror output current back to input instead of using supply
dependent input current!
B. Murmann
EE 214 Lecture 25
8
311
Start-Up Circuit
(WEAK)
•
Unfortunately, self-biasing comes with a built in "chicken and
egg problem"
– There exists a stable operating point with all currents =0
– Can use a simple start-up circuit to solve this problem
B. Murmann
EE 214 Lecture 25
9
VBE Reference
I OUT =
•
VBE 1
R
Utilizes "parasitic" substrate PNP transistor available in any
CMOS technology
B. Murmann
EE 214 Lecture 25
10
312
Temperature Dependence (1)
•
Similar to the sensitivity to supply variations, we can establish
an expression for temperature variations
– Fractional temperature coefficient
∂I OUT
TC F = ∂T
I OUT
•
For the circuit on the previous slide we have
I OUT =
∂I OUT 1 ∂VBE 1 VBE 1 ∂R
=
− 2
∂T
R ∂T
R ∂T
⎛ 1 ∂VBE 1 1 ∂R ⎞
⎟
= I OUT ⎜⎜
−
R ∂T ⎟⎠
⎝ VBE 1 ∂T
VBE 1
R
B. Murmann
EE 214 Lecture 25
11
Temperature Dependence (2)
•
Final result
TC F =
•
1 ∂VBE 1 1 ∂R
−
VBE 1 ∂T
R ∂T
Temperature dependence is usually quite high
1 ∂VBE 1 − 2mV / K
≅
= −0.33% / K
VBE 1 ∂T
600 mV
TC F = −0.53% / K
1 ∂R
≅ +0.2% / K (poly resistor)
R ∂T
•
E.g. ΔT=100K ⇒ ΔI=-53% (!)
B. Murmann
EE 214 Lecture 25
12
313
ΔVBE Reference
I OUT ⋅ R = VBE 1 − VBE 2
kT ⎛ I IN ⎞ kT ⎛ I OUT
⎟−
ln⎜
ln⎜
q ⎜⎝ I S 1 ⎟⎠ q ⎜⎝ I S 2
kT ⎛ I IN I S 2 ⎞
⎟
ln⎜
=
q ⎜⎝ I S 1 I IN ⎟⎠
=
=
kT
ln(n )
q
I OUT =
B. Murmann
1 kT
ln(n )
R q
EE 214 Lecture 25
13
TC of ΔVBE Reference
∂I OUT
= ln(n )
∂T
= ln(n )
TC F =
R
∂VT
∂R
− VT
∂T
∂T
2
R
1 ∂VT 1 ∂R 1 1 ∂R
−
= −
VT ∂T R ∂T T R ∂T
Example
•
TC of resistor and kT/q partially cancel!
B. Murmann
kT
q
VT ⎛ 1 ∂VT 1 ∂R ⎞
⎜
⎟
−
R2 ⎜⎝ VT ∂T R ∂T ⎟⎠
•
TC F =
VT =
1
− 0.2% / K = 0.13% / K
300 K
EE 214 Lecture 25
14
⎞
⎟⎟
⎠
314
ΔVGS Reference (1)
I REF ⋅ R2 = VGS 1 − VGS 2
= VOV 1 − VOV 2
1
m
I REF
⎛
≅ VOV 1 ⎜ 1 −
⎝
⎛
VOV 1 ⎜ 1 −
⎝
≅
R2
1 ⎞
⎟
m⎠
1 ⎞
⎟
m⎠
[Lee, 2nd ed. p.326]
•
Strange result, why is this useful?
B. Murmann
EE 214 Lecture 25
15
ΔVGS Reference (2)
•
The transconductance of M1 is approximately
g m1 =
2 I D1 2 I REF 1
=
VOV 1
VOV 1
1 ⎞
⎛
2 ⋅ ⎜1 −
⎟
m⎠
= ⎝
R2
•
Transconductance of M1 and other devices biased using this
circuit depend only on m and R2!
– This is why this bias circuit is more appropriately called
"constant gm reference"
•
Design aspects
– Can use off-chip resistor to set gm precisely
– Large VOV helps reduce mismatch errors
– Small I⋅R2 makes circuit less sensitive to body effect
B. Murmann
EE 214 Lecture 25
16
315
Constant Settling Time Bias (1)
•
Reference
– I. E. Opris, L. D. Lewicki, "Bias optimization for switched
capacitor amplifiers,"IEEE TCAS II, pp. 985-989, Dec. 1997.
Settling time with constant bias current
B. Murmann
EE 214 Lecture 25
17
Constant Settling Time Bias (2)
•
Settling time of amplifier is given by
t s = t slew + tlin ≅
Vo ,slew ⋅ C
C
+N
I
β ⋅ gm
•
Constant gm bias helps, but is not optimum for minimizing
process variations
•
Really want a bias circuit that keeps ts constant
B. Murmann
EE 214 Lecture 25
18
316
Constant Settling Time Bias (3)
I = I0 +
ts ≅
I0 ≅
B. Murmann
R
Vo ,slew ⋅ C
C
+N
I
β ⋅ gm
⇒I ≅
•
ΔV
C⎛
I ⎞
⎜⎜Vo ,slew + N
⎟
ts ⎝
β ⋅ g m ⎟⎠
Can make setting time
"constant" by choosing
Vo ,slew ⋅ C
ts
1 ⎞
⎛
2 ⋅ β ⋅ ts ⋅ ⎜ 1 −
⎟
k⎠
⎝
R≅
N ⋅C
EE 214 Lecture 25
19
Constant Settling Time Bias (4)
B. Murmann
EE 214 Lecture 25
20
317
Lecture 26
Bandgap Reference
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 26
1
Overview
•
Reading
– 4.4.3 (Temperature Insensitive Biasing)
•
Introduction
– In this lecture we will introduce the basic idea behind the
frequently used "bandgap" voltage reference. Conceptually,
a bandgap reference simply combines two quantities with
opposite temperature behavior to generate a voltage with
(approximately) zero TC.
B. Murmann
EE 214 Lecture 26
2
318
Key Idea
•
kT/q has a positive temperature coefficient
– "PTAT" proportional to absolute temperature
•
VBE of a BJT decreases with temperature
– "CTAT" complementary to absolute temperature
•
Can combine PTAT + CTAT to yield an approximately zero TC
voltage reference
– Useful in circuits that require a stable reference voltage
• E.g. A/D converters
B. Murmann
EE 214 Lecture 26
3
Conceptual Block Diagram
B. Murmann
EE 214 Lecture 26
4
319
A Closer Look at VBE
VBE =
•
kT ⎛ I C ⎞
ln⎜ ⎟
q ⎜⎝ I S ⎟⎠
Even though kT/q increases with temperature, VBE decreases
because IS itself strongly depends on temperature
VBE ≅
kT ⎛ I C VG 0 /( kT / q ) ⎞
⎟⎟
ln⎜ e
q ⎜⎝ I 0
⎠
≅ VG 0 −
kT ⎛ I 0 ⎞
ln⎜ ⎟
q ⎜⎝ I C ⎟⎠
•
I0 is a device parameter, which is (unfortunately) not completely
independent of temperature
– We'll ignore this for now
•
VG0 is the bandgap voltage of silicon "extrapolated to 0°K"
B. Murmann
EE 214 Lecture 26
5
Extrapolated Bandgap
[Pierret, Advanced
Semiconductor
Fundamentals, p.85]
1.205eV
VG 0 =
B. Murmann
EE 214 Lecture 26
1.205eV
= 1.205V
q
6
320
Temperature Coefficient of VBE
VBE ≅ VG 0 −
•
kT ⎛ I 0
ln⎜
q ⎜⎝ I C
Assuming that both I0 and IC are constant over temperature
dVBE
k ⎛I
≅ − ln⎜⎜ 0
dt
q ⎝ IC
•
⎞
⎟⎟
⎠
⎞ VBE − VG 0
⎟⎟ =
T
⎠
Example
dVBE 0.6V − 1.205V
mV
≅
= −2.02
dt
300 K
K
B. Murmann
EE 214 Lecture 26
7
CTAT + PTAT
•
Returning our initial idea, we can now find the condition that
gives us a temperature independent voltage
VBE + M
kT
kT ⎛ I 0 ⎞
kT
ln⎜⎜ ⎟⎟ + M
≅ VG 0 −
q
q ⎝ IC ⎠
q
≅ VG 0 +
•
kT
q
⎛
⎛ ⎞⎞
⎜ M − ln⎜ I 0 ⎟ ⎟
⎜ I ⎟⎟
⎜
⎝ C ⎠⎠
⎝
Combining VBE and an appropriately scaled version of kT/q
produces a temperature independent voltage, equal to VG0
B. Murmann
EE 214 Lecture 26
8
321
Simple CMOS Realization
I1 = I 2 =
ΔVBE
Vout = VBE 1 +
B. Murmann
R1
=
1 kT
ln(n )
R1 q
R2 kT
ln(n )
R1 q
EE 214 Lecture 26
9
Choice of n
•
Usually make n=integer2-1, e.g. n=8
•
Layout:
B. Murmann
EE 214 Lecture 26
10
322
Design Example
•
From measurement data, we know that |VBE| of a unit device is
700mV at room temperature and I=50μA
•
Decided to use n=8
R1 =
1 kT
1
ln(n ) =
26 mV ln(8 ) = 1.08 kΩ
I2 q
50 μA
R2 = R1
VGO − VBE 1
1.205V − 0.7V
= 1.08 kΩ
= 9.34 kΩ
kT
(
)
26
mV
ln
8
ln(n )
q
B. Murmann
EE 214 Lecture 26
11
Nonidealities
•
"Curvature"
– Temperature dependence of I0
•
Offset voltages and TC of offset voltages
•
Resistor mismatch and TC
•
Finite β and β mismatch
•
More next lecture…
B. Murmann
EE 214 Lecture 26
12
323
Lecture 27
Bandgap Reference
(Continued)
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 27
1
Overview
•
Reading
– 4.4.3 (Temperature Insensitive Biasing)
•
Introduction
– Today's lecture will cover several important details that
we've left out in our previous analysis of bandgap
references. We will discuss nonidealities such as "curvature"
and the impact of offset voltages. Finally, we will take a brief
look at state-of-the art implementations and performance.
B. Murmann
EE 214 Lecture 27
2
324
Selected References (1)
•
R. J. Widlar, "New developments in IC voltage regulators," IEEE
J. Solid-State Circuits, pp. 2-7, Feb. 1971.
– First report, LM309 5V regulator
•
A. P. Brokaw, "A simple three-terminal IC bandgap reference,"
IEEE J. Solid-State Circuits, pp. 388-393, Dec. 1974.
– A classic implementation
•
C. Palmer and R. Dobkin, "A curvature corrected micropower
voltage reference," IEEE Int. Solid-State Conference, pp. 58-59,
Feb. 1981.
•
G. Nicollini et al., "A CMOS bandgap reference for differential
signal processing," IEEE J. Solid-State Circuits, pp. 41-50, Jan.
1991.
– Offset compensated amplifier
B. Murmann
EE 214 Lecture 27
3
Selected References (2)
•
T.L. Brooks et al., "A low-power differential CMOS bandgap
reference," IEEE Int. Solid-State Conf., pp. 248-249, Feb. 1994.
– Differential output, stacked diodes
•
H. Banda et al. "A CMOS bandgap reference circuit with sub-1V operation," IEEE J. Solid-State Circuits, pp. 670 - 674, May
1999 .
•
P. Malcovati et al., "Curvature-compensated BiCMOS bandgap
with 1-V supply voltage," IEEE J. Solid-State Circuits, pp. 10761081, July 2001.
B. Murmann
EE 214 Lecture 27
4
325
VBE Revisited
•
Last lecture, we assumed that
VBE ≅ VG 0 −
•
kT ⎛ I 0 ⎞
ln⎜ ⎟
q ⎜⎝ I C ⎟⎠
A more accurate, but empirical model is given by
VBE ≅ VG 0 −
kT ⎛ K 1 ⋅ T r
ln⎜
q ⎜⎝ I C
⎞
⎟⎟
⎠
•
The temperature dependence inside the logarithm slightly
curves the VBE vs. temperature characteristic
– TC of VBE is not quite independent of temperature
•
Parameter r depends on technology, typically 3…6
B. Murmann
EE 214 Lecture 27
5
Curvature
[Lee, 2nd ed., p.322]
B. Murmann
EE 214 Lecture 27
6
326
Collector Current
•
Another superficial assumption was to assume that the current
IC is independent of temperature
– Actually PTAT in the example circuit from last lecture
– Also affected by TC of resistors
•
To capture the temperature behavior of IC, we can introduce yet
another empirical fudge factor and write
VBE ≅ VG 0 −
•
kT ⎛ T r ⎞
⎟
ln⎜ K
q ⎜⎝ T n ⎟⎠
The factor n is 1 for ideal PTAT current behavior
B. Murmann
EE 214 Lecture 27
7
Modified TC of VBE
•
With this refinement, we have
dVBE
d ⎡ kT ⎛ T r ⎞⎤
⎟⎥
≅
ln⎜ K
⎢−
dT
dT ⎣ q ⎜⎝ T n ⎟⎠⎦
k ⎛ T r ⎞ kT K (r − n )T r − n −1
≅ − ln⎜⎜ K n ⎟⎟ −
Tr
q ⎝ T ⎠ q
K n
T
⎤
k ⎡ ⎛ Tr ⎞
≅ − ⎢ln⎜⎜ K n ⎟⎟ − (r − n )⎥
q⎣ ⎝ T ⎠
⎦
kT
VG 0 − VBE + (r − n )
q
=−
T
B. Murmann
EE 214 Lecture 27
8
327
Modified Condition for Zero-TC
0=M
kT dVBE
k
+
=M −
q
dT
q
VG0 − VBE + (r − n )
kT
q
T
⎡
kT ⎤
⎢VG0 − VBE + (r − n ) q ⎥
⎦
⇒M = ⎣
kT
q
•
Or, more interestingly
Vout = VG 0 + (r − n )
•
kT
q
Must design for Vout that is a few kT/q higher than VGO
B. Murmann
EE 214 Lecture 27
9
Modified Output Voltage
[Lee, 2nd ed., p.323]
•
Getting good temperature stability typically requires some
tweaking or calibration
– Some semiconductor foundries provide tried and true
bandgap cells that are optimized for a particular technology
B. Murmann
EE 214 Lecture 27
10
328
Curvature Compensation
[Palmer]
•
Control current such that the impact of "(r-n) term" is minimized
•
In practice, curvature compensation may not be all that effective
if other nonidealities dominate…
B. Murmann
EE 214 Lecture 27
11
Offset Voltage
ΔVBE
=
⎞
1 ⎛ kT
⎜⎜
ln(n ) − VOS ⎟⎟
R1 ⎝ q
⎠
Vout = VBE 1 + VOS +
⎞
R2 ⎛ kT
⎜⎜
ln(n ) − VOS ⎟⎟
R1 ⎝ q
⎠
I1 = I 2 =
= VBE 1 +
B. Murmann
EE 214 Lecture 27
R1
⎛R
⎞
⎞
R2 ⎛ kT
⎜⎜
ln(n )⎟⎟ − VOS ⎜⎜ 2 − 1 ⎟⎟
R1 ⎝ q
⎠
⎝ R1
⎠
12
329
Issues
•
VOS appears amplified at the bandgap output and can cause a
large absolute error in Vout
– Since R2/R1-1 ≅ 8, this means that for VOS=5mV, the error in
Vout will be about 40mV or roughly 3.3%
•
If the bandgap is trimmed after manufacturing, e.g. to yield an
output of VGO+2kT/q, then this is no longer the point of zero TC
in presence of VOS
•
In CMOS, VOS drift is typically 1…10μV/K
– This means Vout will drift at least 8μV/K, which corresponds
to about 6.6 ppm/K
• Good CMOS bandgaps achieve about 10…50ppm/K
•
Possible solutions
– Mitigate impact of offset by stacking two VBE
– Cancel offset or use low offset BJT differential pair
B. Murmann
EE 214 Lecture 27
13
SC Bandgap with Offset Cancellation
[Nicollini]
B. Murmann
EE 214 Lecture 27
14
330
Bandgap with stacked VBE
[Brooks]
B. Murmann
EE 214 Lecture 27
15
Sub-1-V Bandgap
≅1.2V
Vref
[Banba]
•
kT
⎛
⎞
ln( N ) ⎟
⎜
V
q
⎟
= R4 ⎜ BE +
⎜ R2
R3 ⎟
⎜
⎟
⎝
⎠
Idea: Add currents proportional to VBE and kT/q, instead of
stacking voltages
B. Murmann
EE 214 Lecture 27
16
331
Variant with Curvature Compensation
Low offset amplifier
[Malcovati]
B. Murmann
EE 214 Lecture 27
17
Performance
B. Murmann
EE 214 Lecture 27
18
332
Lecture 28
Technology Scaling
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 28
1
Overview
•
Introduction
– The trend of continuously shrinking feature sizes in
integrated circuits has resulted in enormous performance
gains in both digital and analog circuits. However, since
device scaling necessitates the use of smaller supply
voltages, it is often argued that noise limited circuits can no
longer benefit from scaling. In this lecture, we will take a
closer look at this argument and also review basic analog
device performance trends in light of feature size reduction.
B. Murmann
EE 214 Lecture 28
2
333
The Age of "Moore's Law"
•
In 1965, Gordon Moore predicted that there will be an
exponential growth in the number of transistors per integrated
circuit
B. Murmann
EE 214 Lecture 28
3
… And He was Right
[Moore, ISSCC 2003]
B. Murmann
EE 214 Lecture 28
4
334
Minimum Feature Size
Smaller
B. Murmann
EE 214 Lecture 28
5
Faster
B. Murmann
EE 214 Lecture 28
6
335
Average Transistor Price
Cheaper
B. Murmann
EE 214 Lecture 28
7
Misnomer
•
The term "Moore's Law" was coined by the press
– Of course, the exponential progress rate is not set by
fundamental law
– Merely a rate of progress that makes sense for the industry
and keeps things predictable for all involved players
•
•
•
•
•
Device technologists
Makers of manufacturing equipment
Circuit designers
Sales and marketing
The dirty truth is that Moore's law is mostly just a gigantic
economic feedback loop
– With lot's of great innovation fueled by $$$
B. Murmann
EE 214 Lecture 28
8
336
Moore's Law in Action
$$$
B. Murmann
EE 214 Lecture 28
9
Worldwide Semiconductor Sales
[European
Nanotechnology
Roadmap]
B. Murmann
EE 214 Lecture 28
10
337
State-of-the-Art Semiconductor Fab
•
Cost ~ $3,000,000,000
B. Murmann
EE 214 Lecture 28
11
State-of-the-Art Silicon Technology
•
90nm feature sizes, electrical channel length of 50nm
•
Gate oxide thickness 1.2nm, roughly 5 atomic layers
•
8 Layers of metal routing
B. Murmann
EE 214 Lecture 28
12
338
State-of-the-Art Chips
Single-Chip 802.11
Transceiver
[Zargari, ISSCC 2004]
Pentium 4 Processor,
125 Million Transistors
[Schutz, ISSCC 2004]
B. Murmann
EE 214 Lecture 28
13
Impact of Technology Scaling
•
Scaling is great from a digital perspective!
•
Can show that scaling down features and voltages achieves
three things simultaneously
– Higher speed
– More transistors/area
– Lower energy per operation
•
How about analog circuits?
B. Murmann
EE 214 Lecture 28
14
339
Quotes
•
[Vertregt, ESSCIRC 2004]
– "Significant power efficiency improvements are predicted as a
result of scaling to deep sub-micron technology nodes."
•
[Annema, IEEE J. Solid-State Circuits, 12/2005 ]
– "In summary: unlike digital designs, analog circuits can benefit from
technology scaling if the supply voltages are not scaled down."
•
[Nauta, ESSCIRC 2005]
– "The evolution of CMOS technology will continue for many years to
come, which is beneficial for digital circuits but which is not so for
analog."
B. Murmann
EE 214 Lecture 28
15
List of Concerns
•
Reduced supply voltage
•
Low intrinsic gain
•
Variability
•
Distortion
•
Gate leakage
•
Isolation
•
…
•
Cost (mask & wafer)
•
Model accuracy
•
…
B. Murmann
EE 214 Lecture 28
16
340
gm/ID and fT trends
(a)
160
gm /ID [S/A]
30
180nm
130nm
90nm
20
10
0
80
40
-0.2
-0.1
0
0.1
0.2
VGS-Vt [V]
0.3
0.4
0
0.6
0.5
(b)
1200
(gm /ID)*fT [GHz*S/A]
120
180nm
130nm
90nm
1000
800
•
gm/ID essentially
unaffected by
scaling
•
Very high fT in
recent technologies
– Enables RF
CMOS
fT [GHz]
40
600
400
200
0
-0.4
-0.2
0
0.2
VGS-Vt [V]
B. Murmann
0.4
0.6
EE 214 Lecture 28
17
Available Signal Swing
5
VDD [V]
4
3
2
1
0
0.50um
0.35um
0.25um
0.18um
0.12um
90nm
65nm
45nm
32nm
Technology Node
VDD
> 4kT/q
Available Swing < VDD − 8
kT
q
> 4kT/q
B. Murmann
EE 214 Lecture 28
18
341
Noise Limited Circuit Performance
g
BW ∝ m
C
P ∝ VDD ⋅ I D
Swing 2
DR ∝
kT / C
2
⎛ Swing ⎞ g m
BW ⋅ DR
⎟⎟ ⋅
∝ VDD ⋅ ⎜⎜
P
⎝ VDD ⎠ I D
•
Low VDD is generally bad news, but
– Analog designers have worked hard to maintain or even
improve Swing/VDD
• Typical ADC in 0.5μm: Swing/VDD=2/5
• Typical ADC in 90nm: Swing/VDD=0.5/1
– How about gm/ID?
B. Murmann
EE 214 Lecture 28
19
Leveraging fT
40
160
120
180nm
130nm
90nm
20
80
10
0
fT [GHz]
gm /ID [S/A]
30
40
-0.2
-0.1
0
0.1
0.2
VGS-Vt [V]
0.3
0.4
0.5
0
0.6
•
Example
– fT = 50GHz, 130nm: gm/ID = 8S/A, 90nm: gm/ID = 16S/A
•
For "fixed-speed" applications, high fT can be leveraged to mitigate low
VDD penalty
B. Murmann
EE 214 Lecture 28
20
342
Further Considerations
•
Analog building blocks are never completely limited by thermal noise
– Not uncommon to have ~50% dynamic power
• Decreases with scaling
•
Designers are continuing to develop/refine low-voltage design
techniques
– Recent publications show very good analog building block
performance at 1V
•
Bottom line
– Analog design is challenging at 1V, but it's neither impossible
nor detrimental
B. Murmann
EE 214 Lecture 28
21
Intrinsic Gain
(a)
250
180nm
130nm
90nm
ID [μA]
200
•
A real issue
– How to design a highgain op-amp with devices
that have intrinsic gain of
~10?
•
How much worse does this
get at 45nm/65nm?
EE 214 Lecture 28
22
150
100
(VGS-Vt=100mV)
50
0
0
0.2
0.4
0.6
0.8
1
VDS [V]
1.2
1.4
1.6
(b)
50
gm /gds
40
30
20
180nm
130nm
90nm
10
0
0
0.2
B. Murmann
0.4
0.6
0.8
1
VDS [V]
1.2
1.4
1.6
343
Intrinsic Gain in the Near Future
gm /gds
15
10
5
0
45nm (TCAD)
65nm (TCAD)
90nm (BSIM4)
(VGS-Vt=100mV)
0
0.2
0.4
0.6
0.8
1
VDS [V]
•
Pretty bad…
•
Solutions
– Use non-minimum length device (NML-device)
– Use asymmetric device without drain-side pocket implant (Adevice)
– Or, don't try to build op-amps in these technologies…
• E.g. "Digitally Assisted ADC" research in my group
B. Murmann
EE 214 Lecture 28
23
HoleImpactIonization
HoleImpactIonization
Signed Log
Signed Log
26.52
23.2
23.2
20.3
20.3
-0.12
-0.12
26.58
-0.16
-0.16
A-Device vs. Standard Device
17.4
17.4
14.5
14.5
5.8
2.9
2.9
-0.04
-0.04
8.7
5.8
0.4
0.36
0.32
0.28
0.24
0.2
0.16
0.12
0.08
0.04
0
0
0.44
0.44
0.4
0.36
0.32
0.28
0.24
0.2
0.16
0.12
0.08
0.04
0
0
-0.16
•
11.6
-0.08
-0.08
11.6
8.7
-0.12
-0.08
-0.04
0
0.04
0.08
0.12
0.16
-0.16
-0.12
-0.08
-0.04
0
0.04
0.08
0.12
0.16
Removing halo widens depletion region
– Reduces impact ionization and improves output resistance
B. Murmann
EE 214 Lecture 28
24
344
Intrinsic Gain of Alternate Devices (45nm)
( )
gm /gds
150
Min. length
NML-device
A-device
100
(VGS-Vt=100mV)
50
0
(VGS-Vt=100mV)
0
0.2
0.4
0.6
0.8
1
VDS [V]
•
For both NML and A-device Lphysical=80nm (Lphysical=24nm for
minimum length device)
•
Great, lots of gain!
– But how about fT?
B. Murmann
EE 214 Lecture 28
25
gm/ID and fT for Alternate Devices (45nm)
(a)
40
800
600
Minimum length
NML-device
A-device
20
10
0
fT much lower than
for minimum length
45-nm device
– But still better
than minimum
length device in
90nm…
•
Who needs fT >
200GHz in an opamp…?
200
-0.2
-0.1
0
0.1
0.2
VGS-Vt [V]
0.3
0.4
0
0.5
(b)
5000
(gm /ID)*fT [GHz*S/A]
400
•
fT [GHz]
gm /ID [S/A]
30
Minimum length
NML-device
A-device
4000
3000
2000
1000
0
B. Murmann
-0.2
-0.1
0
0.1
0.2
VGS-Vt [V]
0.3
0.4
0.5
EE 214 Lecture 28
26
345
Variability (1)
[Courtesy A. Bowling,
Texas Instruments]
B. Murmann
EE 214 Lecture 28
27
Variability (2)
•
Must keep in mind that even if we draw nice colorful rectangles
in our CAD tools, things won't come out like that…
•
Device shape strongly depends on surroundings
•
Good old layout recipes become more important and must be
applied even when only "moderate" matching in required
– Use unit devices
– Match surroundings of unit devices
B. Murmann
EE 214 Lecture 28
28
346
Variability (3)
[Marcel Pelgrom, NXP]
•
Device mismatch larger than process corner variations!
– For small "digital" transistors…
B. Murmann
EE 214 Lecture 28
29
Variability (4)
Digital
Analog
•
•
A well known problem
– Designers are used to
"caring" about mismatch
•
Lots of options and potential
solutions
– Layout techniques,
analog or digital
calibration, dynamic
element matching, larger
device area, …
•
Usually care about matching
for a few up to a few hundred
transistors
•
A "new" problem
– Significant impact on
achievable performance,
yield, design
methodology, EDA, …
Big difference compared to
analog
– Care about millions if not
billions of devices!
B. Murmann
EE 214 Lecture 28
30
347
An Interesting Hike Lies Ahead…
D
Bag of
Tricks
B. Murmann
A
EE 214 Lecture 28
31
Cost – The "Real" End of The Roadmap?
[Marcel Pelgrom, NXP]
•
Reference point: 30 mm2 die in 0.12μm CMOS
B. Murmann
EE 214 Lecture 28
32
348
Lecture 29
Class Summary
Project Discussion
Boris Murmann
Stanford University
murmann@stanford.edu
Copyright © 2007 by Boris Murmann
B. Murmann
EE 214 Lecture 29
1
Transistor Models
•
Any model is an approximation of the real world
– Must leave many details out
– Must retain the important details (to be useful)
– Appropriate level depends on questions you want to answer
• BSIM model/Spice
• gm/ID approach
• Long channel equations
•
When designing and analyzing circuits, we are usually forced to
use much simpler models than the ones available in Spice
– gm/ID methodology partially closes this gap
•
A "good" IC designer is always on the lookout for modeling
limitations!
B. Murmann
EE 214 Lecture 29
2
349
Transistor Figures of Merit
•
Transit Frequency
ωT =
•
gm
C gg
Current Efficiency
gm
ID
•
Intrinsic Gain
gm
g ds
B. Murmann
EE 214 Lecture 29
3
Circuit Analysis
•
Once we have appropriate models, we can in principle analyze
any circuit using KCL/KVL
– Usually too difficult and also won’t allow us to reason about
design choices and tradeoffs
•
Crutches for circuit analysis
– Small signal approximation
– Zero value time constant method
– Miller approximation
– Return ratio analysis
– Blackman's impedance formula
– …
•
Again, a good designer will always be on the lookout for
potential limitations of the respective analysis method
B. Murmann
EE 214 Lecture 29
4
350
The "Atoms" of Analog Circuit Design
•
Common source
– Basic voltage amplifier
•
Common gate
– Good for "shielding"
– Can help boost output impedance, mitigate Miller effect
•
Common drain
– Buffer, level shifter
•
Differential pair
B. Murmann
EE 214 Lecture 29
5
Feedback (1)
•
Desensitizes circuit to (forward-) gain variations
•
Modifies port impedances
•
Central quantity of interest: Loop gain T
– Quantifies static error
– Used to assess stability, phase and/or gain margin
– Helpful in calculating closed loop port impedances
•
Finding T is simple
– E.g. break loop at transconductor, inject test current,
calculate ratio of return and test current
•
Typically need a dominant pole
– Hard to use more than 2-3 amplifier stages and maintain
sufficient phase margin
B. Murmann
EE 214 Lecture 29
6
351
Feedback (2)
•
Impact of phase margin on step response
– Excessive "ringing" for phase margin <60°
– Fast settling for phase margin ~70°
•
Impact of phase margin on closed loop AC response
– Gain peaking and slightly larger 3-dB bandwidth for small
phase margin
•
Compensation techniques
– Can use simple load compensation for single stage OTAs
– Miller compensation is most popular for two-stage designs
• Can push parasitic zero to infinity using nulling resistor
• Beware of tricks such as pushing the zero into the LHP to avoid
pole-zero doublets
– Cascode compensation
• For designers with experience…
B. Murmann
EE 214 Lecture 29
7
OTAs with Capacitive Feedback
•
Primary application: Switched capacitor circuits
•
Important performance metric: Settling time
– Small signal model fails when input exceeds "linear range"
– Slew rate ~ I/C
– ts = tlin + tslew
•
High performance OTAs are usually implemented as fully
differential circuits
– Need CMFB
– Better XXX-rejection than single ended circuits
– Easy to analyze, since circuit is "perfectly" balanced
• Exceptions occur during transients, in presence of mismatch,
etc.
B. Murmann
EE 214 Lecture 29
8
352
Electronic Noise
•
Tends to set achievable power dissipation in circuits with high
DR requirements
•
Fundamental noise
– Thermal noise
•
Technology related noise
– 1/f noise
•
Increasing (noise limited) precision by "one bit" quadruples
power dissipation!
B. Murmann
EE 214 Lecture 29
9
References
•
Self-biasing concept
– Beware of stability issues!
•
Current references
– VDD, VGS, ΔVGS, VBE, ΔVBE - based approaches
– To first order, ΔVGS bias yields constant gm over temperature
and process
•
Voltage reference
– Bandgap
– Useful as a reference whenever there's a need to convert
"bits to Volts" or "Volts to bits"
– Conceptually simple, but lots of second order issues
B. Murmann
EE 214 Lecture 29
10
Download