Advanced Analog Integrated Circuit Design (B. Murmann)
Advanced Analog Integrated Circuit Design (B. Murmann)
- Winter 2014 -
Boris Murmann
Stanford University
[email protected]
Table of Contents
Chapter 1 Introduction
Chapter 2 Bipolar Junction Transistors
Chapter 3 Elementary BJT Amplifier Stages
Chapter 4 MOS Transistor Modeling
Chapter 5 gm/ID-Based Design
Chapter 6 Electronic Noise
Chapter 7 Feedback: Introduction
Chapter 8 Feedback: TIA Example
Chapter 9 Feedback: Cherry-Hooper Example
Chapter 10 Feedback: Root Locus and Frequency Compensation
Chapter 11 Low-Frequency Distortion Analysis
Chapter 12 High-Frequency Distortion Analysis
Chapter 13 Mismatch
(This page is intentionally left blank)
Chapter 1
Introduction
Boris Murmann
Stanford University
Winter 2013-14
Design of mixed-signal
and RF building blocks
EE314B
Advanced RF
Integrated Circuit
EE114/EE214A EE214B Design
Fundamentals of Advanced Analog
Analog Integrated Integrated Circuit
Circuit Design Design EE315A
VLSI Signal
Conditioning Circuits
EE315B
VLSI Data
Conversion Circuits
EE 214A
Bandwidth
Power Dissipation
EE 214B
Advanced Technologies
Signal-to-Noise Ratio
2
Psignal Vsignal
SNR = ∝ 2
Pnoise Vnoise
http://www.soe.ucsc.edu/~htakeda/kernelreg/kernelreg.htm
Nonlinear Distortion
vo
vi
Small-signal approximation
For a single tone input, the nonlinear terms in a circuit’s transfer function
primarily result in signal harmonics
Main objective
– Acquire the basic tools and intuition needed to analyze noise and
distortion in electronic circuits
– Look at a few specific circuit examples to “get a feel” for situations
where noise and/or distortion may matter
Technology
EE214A
EE214B
C. Knochenhauer et al., “40 Gbit/s transimpedance amplifier with high linearity range in 0.13 µm SiGe
BiCMOS, Electronics Letters, May 12, 2012.
Trends
Instructor
– Boris Murmann
Teaching assistant
– Valerie Barry
Administrative Assistant
– Ann Guerra, Allen-207
Lectures are available online
– But please come to class to keep the discussion interactive!
Web page
– http://coursework.stanford.edu/homepage/W14/W14-EE-214B-01.html
– Check regularly, especially the discussion board
– Register for online access to grades and solutions
Required textbook
– Chan Carusone, Johns, Martin, Analog Integrated Circuit Design, 2nd
Edition, Wiley, 2011
– Errata: http://analogicdesign.com/students/errata/
Reference texts
– B. Razavi, Design of Integrated Circuits for Optical Communications,
2nd Edition, McGraw-Hill, 2012
– Gray, Hurst, Lewis and Meyer, Analysis and Design of Analog
Integrated Circuits, 5th Edition, Wiley, 2008
Course prerequisite: EE114/214A or equivalent
– Basic device physics and models
– Frequency response, dominant pole approximation, ZVTC
– Biasing, small-signal models
– Common source, common gate, and common drain stages
– Port impedance calculations
– Feedback basics
EE214B EE315A
EE315B
Simulation
– HSpice + circuit netlists
– Cscope and/or Matlab for post-processing
– You can use your own tools/setups “at own risk“
Getting started
– Read “CAD basics” handout provided on the course website
EE214B Technology
– 0.18-µm SiGe BiCMOS
– BSIM3v3 models provided under /usr/class/ee214b/hspice
– Models correspond to a typical technology as described in
• Wada, et al., “A manufacturable 0.18-um SiGe BiCMOS
technology for 40-Gb/s optical communication LSIs,” BCTM 2002
Assignments
Homework (20%)
– Handed out on Wednesdays, due following Wednesdays in class
– Late submission penalty: 0.5 dB/hour
– Lowest HW score will be dropped
– Policy for off-campus students
• Fax or email to SCPD before deadline stated on handout
Midterm Exam (30%)
Design Project (20%)
– Design of an amplifier using HSpice (no layout)
– Work in teams of two
– OK to discuss your work with other teams, but no file exchange!
Final Exam (30%)
Course Outline
Boris Murmann
Stanford University
Winter 2013-14
History
VBE
Device acts as a voltage
IC ∝ e kT / q controlled current source
– VBE controls IC
The base-emitter junction is
forward biased and the base-
collector junction is reverse
IB << IC n- C
biased
p B VCE The device is built such that
– The base region is very thin
VBE n+ E
– The emitter doping is much
higher than the base doping
– The collector doping is much
lower than the base doping
Outline of Discussion
Built-in Potential
The built in potential sets up an electric field that opposes the diffusion of
mobile holes and electrons across the junction
dp
(Drift ) qµ ppE = qDp (Diffusion )
dx
p n N N kT
⇒ ψ 0 = VT ln p 0 = VT ln n0 ≅ VT ln A 2 D VT =
pn0 np 0 ni q
n N ND
VBE VBE
n2 BE
V
The result on the previous slide shows that forward biasing increases
the concentration of electrons at the “right” edge of the depletion region
by a factor of exp(VBE/VT)
The same holds for holes at the “left” edge of the depletion region
VBE VBE
ni2
pn (0) = pn0 ⋅ e VT ≅ ⋅ e VT
ND
Since ND >> NA, it follows that pn(0) << np(0), i.e. the concentration of
minority carriers is much larger at the lightly doped edge
The carriers would “like” to diffuse further into the neutral regions, but
quickly fall victim to recombination
The number of minority carriers decays exponentially, and drops to 1/e
of the at the so-called diffusion length (Lp or Ln, on the order of microns)
Lots of electrons being injected into the p-region, not all that many holes
get injected into the n+ region
– The heavier n-side doping, the more pronounced this imbalance
becomes
The electrons injected in the p region cause a diffusion current that
decays in the x-direction due to recombination
The recombination necessitates a flow of holes to maintain charge
neutrality; as the diffusion current decays, the hole current increases,
yielding a constant current density along the device
Near the edge of the depletion region, the electron diffusion current
dominates over the hole current that supplies carriers for recombination
– This is a very important aspect that we will come back to
“cut here”
n+ p n-
BJT Currents
http://en.wikipedia.org/wiki/Bipolar_junction_transistor
VBE qADnni2
∴ IC ≅ IS e VT IS =
WBNA
Base Current
IB1 follows from dividing the minority carrier charge in the base (Qe) by its
“lifetime” (τB)
1 VBE
Qe 2 np (0)WBqA 1 WBqAni2 VT
IB1= = = e
τb τb 2 τbNA
IB2 depends on the gradient of minority carriers (holes) in the emitter. For a
“long” emitter (all minority carriers recombine)
2 VBE − x VBE
qADp ni2 V
= −qADp i e VT e p
dpn (x) d n L
IB2= −qADp = e T
dx x =0 dx N Lp ND
D x =0
IE = − (IC + IB )
IC
βF = (ideally infinite)
IB
IC β
αF = = F (ideally one)
( −IE ) 1 + βF
The subscript “F” indicates that the device is assumed to operate in the
forward active region (BE junction forward biased, BC reverse biased, as
assumed so far)
– More on other operating regions laterB
Side note:
BJT inherently has better (higher)
ro than MOS since lower doping
on n-side (collector) has most of
the depletion region inside the
collector
∂ qADnni2
VBE
∂IC I dWB
= e VT =− C
∂VCE ∂VCE WB (VCE ) ⋅ NA
W B dVCE
IC WB
VA = =− = const. (independent of IC )
∂IC dWB
∂VCE dVCE
VBE
VT VCE
IC ≅ ISe 1 +
VA
VBE
dIC d VT VCE IC
gm = = ISe 1 + =
dVBE dVBE VA VT
I
d C
β 1 IC
= F=
1 dIB g
gπ = = = m (assuming βF = const.)
rπ dVBE dVBE βF VT βF
VBE
1 dIC d VT VCE IC
go = = = ISe 1+ ≅
r0 dVCE dVCE VA VA
Intrinsic Gain
IC VA VA
gmro ≅ ⋅ = VT ≅ 26mV (at room temperature)
VT IC VT
In the EE214B technology, the SiGe npn device has VA = 90V, thus
90V
gmro ≅ = 3460
26mV
Discussed so far
BE = forward biased
CE = reverse biased
Gummel Plot
≅ 7000 ppm / °C
Big mess!
First focus on intrinsic elements
Charge Storage
Junction Capacitance
C j0 2C j0
Cj = n
VD
1 −
ψ0
Cµ
B C
v1 rπ Cπ gmv1 ro
–
E
Cπ = Cb + C je = Cb + 2C je0
C jc0
Cµ = C jc = n
VCB
1 +
ψ 0c
Neglect
Range of numbers
re ~1-3Ω
rb ~ 50-500Ω Values at high end of these ranges may have large
impact on performance Try to minimize through
rc ~ 20-500Ω
advanced processing & technology
CCS ~ 3-200fF
Oxide isolated
Self-aligned structure (base and emitter align automatically)
Very thin base (~100nm or less) through ion implantation
Reduced breakdown voltages compared to more traditional structures
qADnnp0 qADnnp0 2
qADnniB
DnNDLp 2
WB WB WBNA niB
βF = 2
≅ 2
= 2
= 2
⋅
1 np0 WBqA qADpniE qADpniE qADpniE DpNA WB niE
+
2 τb LpND LpND LpND
Added degree of
freedom for HBT
300
2
90
3x smaller device
3.2x10-17A
1pA 5x bigger IS
2.0V
5.5V
3.3V
0.56ps 18x smaller τF
10ps
25Ω 16x smaller rb
60Ω
2.5Ω
6.26fF
0.8V
0.4
3.42fF
0.6V
0.33 Oxide isolation vs.
3.0fF
Junction isolation
0.6V
0.33
CMOS
[Texas Instruments]
[Texas Instruments]
గ ఓ
ଵ = గ 1 −
1 + గ (గ + ఓ )
=
1 + గ (గ + ఓ )
= ଵ + ఓ ଵ
1
Low frequencies: ≅ గ = ி for ≪ = ఉ
గ గ + ఓ
High frequencies: ≅ for ఉ ≪ ≪
గ + ఓ ఓ
|io/ii| gm
(asymptote)
(
ω Cπ + Cµ )
≅ βF
Note that rπ “matters” only for
frequencies up to ωβ = ωT/βF
݃
ܥఓ
gm gm
1= ⇒ ωT =
(
ω T C π + Cµ ) Cπ + Cµ
1 C C je Cµ C je Cµ
τT = = b+ + = τF + +
ωT gm gm gm gm gm
“peak fT”
gm = IC/VT increases
EE214B Technology
Assumed to be similar to a 0.18-µm BiCMOS technology featuring a
high-performance SiGe npn device
– VCC = 2.5V (BJT), VDD=1.8V (MOS)
See e.g.
– Wada et al., BCTM 2002
– Joseph et al., BCTM 2001
– IBM 7HP documentation
• https://www-01.ibm.com/chips/techlib/techlib.nsf/products/BiCMOS_7HP)
NPN PMOS/NMOS Poly resistor
http://fuji.stanford.edu/events/spring01/slides/harameSlides.pdf
/afs/ir.stanford.edu/class/ee/synopsys/B-2008.09-SP1/hspice/docs_help
PDF files:
home.pdf hspice_cmdref.pdf hspice_integ.pdf hspice_relnote.pdf hspice_sa.pdf
hspice_devmod.pdf hspice_mosmod.pdf hspice_rf.pdf hspice_si.pdf
* C B E
q1 c b 0 npn214
Vc c 0 1.25
ib 0 b 1u
.op
.dc ib dec 10 10f 100u
.probe ib(q1) ic(q1) ie(q1) 1x VCC/2
.probe gm = par('gm(q1)')
.probe go = par('g0(q1)')
.probe cpi = par('cap_be(q1)')
.probe cmu = par('cap_ibc(q1)')
.probe beta = par('beta(q1)')
Transit Frequency
150
fT [GHz]
100
50
0 -4 -3 -2
10 10 10
IC [A]
ee215 npn
4000
3500
3000
2000
1500
1000
500
0 -8 -6 -4 -2
10 10 10 10
IC [A]
gm/IC
Important to realize that gm will not be exactly equal to IC/VT at high currents
NPN (1x, A =0.7µm2, I =0.2, 0.4, ..., 1µA NMOS 2/0.18, V =0.6, 0.8, ..., 1.4V
E B GS
300 800
700
250
600
200
500
I C [µA]
I D [µA]
150 400
300
100
200
50
100
0 0
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5
V [V] V [V]
CE DS
Boris Murmann
Stanford University
Winter 2013-14
MOS
BJT
MOS
BJT
Common-Emitter Stage
VCC
Vo IB
RL VCC
RS +
Q1 VO+vo
vi ~ –
VI
Vi Vi
IB IC = βFIB
RS + + R
VI VBE Vo L VCC
– –
Is e qVBE kT
VI − VBE(on) RL
IB ≅ VO = VCC − ICRL = VCC − βFIBRL = VCC − βF
RS
( VI − VBE(on) )
RS
Simple Example
VCC = 3V
1kΩ
vO
vI
1kΩ 700Ω
2.5
2 10
0
1.5
I C [mA]
I C [mA]
0.5 10
-1
0
0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.6 0.65 0.7 0.75 0.8 0.85 0.9
VBE [V] VBE [V]
For the useful range of unit collector current between 0.162mA, VBE
changes by less than 100mV
We can therefore approximate VBE ≅ const. ≅ 0.8V for bias calculations
Ri rb 1 Cµ 2
+ +
vi ~ v1 rπ Cπ gmv1 ro RL CL vo
– –
For hand analysis, we will usually neglect rc and re (from the BJT model)
If significant, rb can be included with Ri, i.e. RS = Ri + rb
Resulting low-frequency gain
rπ
A v0 = − ⋅ gmRLtot RLtot = ro || RL
RS + rπ
=1 for MOS
Frequency Response
1 1
p1 ≅ − =−
b1 RS Cπ + Cµ (1 + gmRLtot ) + RLtot (CL + Cµ )
Emitter Degeneration
VCC
RL
RS VO+vo
Ro gm
Gm ≅
vi ~ 1 + gmRE
Ri
RE Ro ≅ ro (1 + gmRE )
VI
ve RE g R
K= ≅ = m E
vb 1
+ RE 1 + gmRE
gm
rπ rπ
Ri = ≅ = rπ (1 + gmRE )
1− K gmRE
1−
1 + gmRE
Ri ≅ rπ + RE (1 + β ) = rπ + RE (1 + gmrπ ) ≅ rπ (1 + gmRE )
Tricks of this kind are useful for reasoning about low frequency behavior
A more detailed analysis is required when investigating frequency
dependence
RS Cµ
+ +
vi ~ v1 rπ Cπ gmv1 ro RL vo
– –
+
vE RE
–
Deriving the transfer function of this circuit requires solving a 3x3 system
of equations
This is very messy, and so we will just look at an OCT-based bandwidth
estimate to gain some basic insight
Useful Expressions
RC + RE
Ro ≅ ro (for β → ∞ )
1 + gmRE
R RC
RB RB + RE
R π ≅ rπ
Ro 1 + gmRE
RS Cµ
+ +
vi ~ v1 rπ Cπ gmv1 ro RC vo
– –
+
vE (neglect)
RE
–
≅ RS + RC + GmRCRS = RS (1 + A v0 ) + RC
RS + RE RS + RE
R π ≅ rπ ≅
1 + gmRE 1 + gmRE
RE
1+ 1
R
τ = RS (1 + A v0 ) + RC Cµ + 1 + g RS RSCπ ω−3dB ≅
τ
m E
τ ≅ RS (1 + A v0 ) + RC Cµ + RSCπ
Adding RE can help improve the bandwidth, provided that gm > 1/RS
− Note, however, that gm (and hence the power dissipation must be
increased) to maintain the same Av0
R C R
τ ≅ 1 + S π ω−3dB ≅ ωT / 1 + S Near ωT for small RS
RE gm RE
VCC
RS
Q1
vi ~
VO+vo
VI IB RL CL
1 R 1 RS
Ro ≅ + S ≅ 1 +
gm β + 1 gm rπ
RS vb
vi
vo
Ri ≅ rπ (1 + gmRL ) RL
vo vb vo rπ (1 + gmRL ) gmRL
A v0 = = ≅
vi vi v b rπ (1 + gmRL ) + Rs 1 + gmRL
gmRL
≅ for rπ (1 + gmRL ) >> Rs
1 + gmRL
≅1 for gmRL >> 1 and rπ (1 + gmRL ) >> Rs
Common-Base Stage
VCC
io iC β
RL Ai = A i0 = =
ii iE β + 1
VO+vo
io Ro
Neglecting rb, rc, re and rπ, we have
Ri Ro ≅ ro (1 + gmRS )
Ii + ii RS (large) 1 RL 1
Ri ≅ 1+ ≅
gm ro gm
IC1 IC2 I
IIN = IC1 + IB1 + IB2 = IC1 + + ≅ IC2 1 + 2 C2 for IS1 = IS2
β β β
IOUT 1 2
≅ ≅ 1−
IIN 2 β
1 + β
IIN VCC
IOUT
Q3
+ IC1 IC2 I
IE3 − IE3 = + ≅ 2 C2 assumin g IS1 = IS2
Q1 Q2 VOUT β β β
IOUT 1 2
≅ ≅ 1−
IIN 2 β2
1+ 2
β +β
IIN IOUT
VCC
RC1
Differential Input Voltage
RC2
ܸௗ = ܸଵ − ܸଶ
Vo1 Vo2
IC1 IC2 Differential Collector Current
ܫௗ = ܫଵ − ܫଶ
Vi1 Q1 Q2 Vi2
Differential Output Voltage
ܸௗ = ܸଵ − ܸଶ
ITAIL RTAIL
VEE
The following large signal analysis neglects rb, rc, re, finite REE and
assumes that the circuit is perfectly symmetric
Vbe1 Vbe2
VT VT
Vi1 − Vbe1 + Vbe2 − Vi2 = 0 IC1 ≅ IS1 e IC2 ≅ IS2 e
1 αITAIL αITAIL
ITAIL = − (Ie1+ Ie2) = (Ic1+ Ic2) ⇒ Ic1 = V
Ic2 = V
α − id + id
VT VT
1+ e 1+ e
1 1 Vid
Icd = Ic1 − Ic2 = αFITAIL Vid
− V = αFITAIL tanh 2V
−
VT
+ id
T
1 + e 1 + e VT
V
Vod = IodRL = αITAILRL tanh id
2VT
VCC
RC RC
Vo1 Vo2
Vi1 Q1 Q2 Vi2
RE RE
ITAIL
VEE
I I
VT ln i = VT ln o
Is1 Is 2
Ii I
= o
Is1 Is 2
Is 2
Io = Ii
Is1
Square-Root Circuit
I I I I
VT ln B + VT ln i = VT ln o + VT ln o
Is1 Is 2 Is 3 Is 4
IB Ii I I
= o o
Is1 Is 2 Is 3 Is 4
Is 3 Is 4
Io = Ii IB
Is1 Is 2
IB1 Ii Ii I I I
= B2 B2 o
Is1 Is 5 Is 2 Is 3 Is 4 Is 6
IB1 Is 3 Is 4 Is 6
Io = Ii2
IB2 2 Is1 Is 5 Is 2
Boris Murmann
Stanford University
Winter 2013-14
Outline
How to calculate the drain current (ID) current as a function of VGS, VDS?
Simplifying Assumptions
ID = Qn ⋅ v ⋅ W
v = µ ⋅E
W VDS
ID = µCox ( VGS − Vt ) − ⋅ VDS
L 2
S D
VDS = 0
Qn (y) = Cox [ VGS − V(y) − Vt ]
n+ n+
G
G
The saturation current has
Qn(L) = 0
S D
some dependence on VDS due
to channel length modulation
VDS = VGS - VTn
n+ n+
G
Qn(L- L) = 0 1 W
S D
ID = µCox (VGS − Vt )2 (1 + λVds )
2 L
VDS > VGS - VTn
n+ n+
“Lambda Model”
L
dID W 2I
gm = = µCox VOV (1 + λVDS ) = D VOV = VGS − Vt
dVGS L VOV
dID 1 W λID
go = = µCox VOV 2 ⋅ λ = ≅ λID
dVDS 2 L 1 + λVDS
dID gm γ
gmb = =
dVBS 2 2φ f + VSB
Capacitances
Intrinsic
Text, p. 31
AD ⋅ CJ PD ⋅ CJSW
Cdb = MJ
+ MJSW
AD = WLdiff
VDB VDB
1 + PB 1 + PB PD = W + 2Ldiff
10/0.18
10um
0.5um 0.3um
* HSpice Netlist
* d g s b
mp1 0 in out out pmos214 L=0.18um W=10um
d1 0 out dwell 37.4p
Col = WC'ol
“Overlap Capacitance”
Chipworks
For the most recent generation of MOS transistors, the overlap and
fringe capacitances from gate to drain/source are about as large as the
intrinsic gate capacitance!
3.5 2
3 Square Law
2.5 1.5
SQRT(I D [mA])
ID [mA]
2
1
1.5
1
0.5
0.5
0 0
0 0.5 1 1.5 0 0.5 1 1.5
VGS [V] VGS [V]
Two observations
– The transistor does not abruptly turn off at some Vt
– The current is not perfectly quadratic in (VGS–Vt)
40
NMOS214
30 Square Law (2ID/VOV)
BJT (q/kT)
gm/ID [S/A]
20
10
0
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
VGS [V]
The square law fails miserably at predicting gm/ID for low VGS
0.2 0.48
0.19 0.46
ID⋅L [mA⋅µm]
0.18 0.44
Vt [V]
0.17 0.42
0.16 0.4
0.15 0.38
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
L [µm] L [µm]
The current does not scale perfectly with 1/L (ID⋅L ≠ const.)
The threshold voltage of the device depends on the channel length
0
10
NMOS214
Square Law
-2
ID, IC [mA] 10 NPN214
~90mV/decade
-4 ~60mV/decade
10
-6
10
What is Vt, anyway? The device does not turn off at all, but really
approaches an exponential IV law for low VGS
What determines the current at low VGS?
Definition of Vt
Before inversion occurs, the electrostatic field from the gate forward-
biases the source-side pn junction at the surface
Physics governed by the “gated diode” model
Cox
Cjs
టೞ టೞ ିವೄ
0 = =
0 −
=
ഗೞ ೇ
ଵ ି ವೄ
= ೇ (1 − ೇ
)
௦ ௫ 1
= =
ீௌ ௦ + ௫
0
10
NMOS214
-2 Weak Inversion
10
Square Law
ID [mA]
-4
10
-6
10
30
NMOS214
25 Weak Inversion
Square Law
20
ID [mA]
15
10
0
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
VGS [V]
We now have a better idea now about the maximum possible gm/ID, but
this does not help in the transistor region between the two IV laws
Moderate Inversion
In the transition region between weak and strong inversion, the drain
current consists of both drift and diffusion currents
One can show that the ratio of drift/diffusion current in moderate
inversion and beyond is approximately (VGS-Vt)/(kT/q)
This means that the square law equation (which assumes 100% drift
current) does not work unless the gate overdrive is several kT/q
– Recall that in EE214A, you used the square law model only for
VGS-Vt > 150mV ≅ 6 kT/q
Is there a simple expression that works for all three regions (weak,
moderate and strong inversion)?
The so-called EKV model was developed with this intent
– C. Enz and E. A. Vittoz, Charge-Based MOS Transistor Modeling -
The EKV Model for Low-Power and RF IC Design, Wiley, 2006
= 2்ଶ ∙ ௫ ௦ଶ +
( ௦ )
௧ ௨௦
ீௌ − ௧ = ் 2 ௦ − 1 + ln ௦
Comparison
0
10
-2
10
ID [mA]
-4
10
NMOS214
-6
10 Square Law
Basic EKV
0 0.5 1 1.5
VGS [V]
Many more issues exist; we will once again only discuss the most
relevant subset
In the derivation of the square law model, it is assumed that the carrier
velocity is proportional to the lateral E-field, v=µE
Unfortunately, the speed of carriers in silicon is limited (vscl ≅ 105 m/s)
– At very high fields (high voltage drop across the conductive part of
the channel), the carrier velocity saturates
VOV 200mV V
E = e.g. = 1.11⋅ 106
L 0.18µm m
Field Estimates
m
105
v s = 6.7 ⋅ 106 V
Ec = scl ≅
µ cm2 m
150
Vs
Therefore
V
1.11⋅ 106
E m
= ≅ 0.16
Ec V
6.7 ⋅ 106
m
This means that for VOV on the order of 200mV, the carrier velocity is
somewhat reduced, but the impairment is relatively small
The situation changes when much larger VOV are applied, as the case in
digital circuits
A simple equation that captures the moderate deviation from the long
channel drain current can be written as
1 W 2 1
ID ≅ µCox VOV ⋅
2 L VOV
1+
EcL
V
Minimum-length NMOS: EcL = 6.7 ⋅ 106 ⋅ 0.18µm = 1.2V
m
V
Minimum-length PMOS: EcL = 16.75 ⋅ 106 ⋅ 0.18µm = 3V
m
Reality Check
2
NMOS214
Square Law
1.5 Square Law With Velocity Saturation
SQRT(I D [mA])
0.5
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
VGS [V]
1 W 2 1
ID ≅ µCox VOV ⋅ 1 (see text, page 45)
2 L
(1 + [ θVOV ] )
m m
The parameters m and θ are “fudge factors” that allow us to fit this
expression to measured data
In the square law model, we attributed the ID-VDS dependence (and thus
finite intrinsic gain) primarily to channel length modulation
In a short channel device, it turns out that the intrinsic gain is strongly
affected by another effect called DIBL
In essence, the drain can be viewed as an additional gate that
modulates the inversion charge
SCBE = substrate
current induced body effect
[BSIM3 manual]
DIBL
0.42
0.415
0.41
0.405
RSCE
Vt [V]
0.4
SCE
0.395
0.39
Boris Murmann
Stanford University
Winter 2013-14
Specifications
Circuit
Results
Unfortunate Consequence
* /usr/class/ee214b/hspice/techchar.sp
.inc '/usr/class/ee214b/hspice/ee214_hspice.sp'
.inc 'techchar_params.sp'
.param ds = 0.9
.param gs = 0.9
% HSpice toolbox
addpath('/usr/class/ee214b/matlab/hspice_toolbox')
>> nch
>> size(nch.ID)
ans =
22 73 73 11
ans =
8.4181e-006
In usage scenarios (1) and (2) the input parameters (L, VGS, VDS, VS) can be
listed in any order and default to the following values when not specified:
Key Question
Square Law
Transconductance efficiency
– Want large gm, for as little current
gm 2
=
as possible ID VOV
Intrinsic gain gm 2
– Want large gm, but no go ≅
go λVOV
40
fT
30 Moderate Inversion
gm/ID [S/A], f T [GHz]
gm/ID
20
Weak Inversion Strong Inversion
10
0
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5
VOV [V]
250
200
gm/ID⋅f T [S/A⋅GHz]
150 Moderate Inversion
100
50
Weak Inversion Strong Inversion
0
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5
VOV [V]
Design in a Nutshell
ID
gm/ID
The inversion level is also fully defined once we pick gm/ID, so there is no
need to know VOV
ID 40
30
f T [GHz]
gm/ID 20
10
fT 0
5 10 15 20 25
gm/ID [S/A]
L=0.18um
L=0.5um
L=0.5um
L=0.18um
L=0.18um
L=0.5um
VDS Dependence
VDS dependence
is relatively weak
Typically OK to
work with data
generated for
VDD/2
NMOS, L=0.18um
1
Cdd/Cgg
Cgd/Cgg
0.8 Again, usually OK
to work with
0.60
estimates taken at
0.6
VDD/2
0.4
0.24
0.2
0
0 0.5 1 1.5
VDS [V]
0.5
0.4
0.3
0.2
0.1
0
0.2 0.25 0.3 0.35 0.4 0.45 0.5
L [µm]
0.5
0.4
0.3
0.2
0.1
0
0.2 0.25 0.3 0.35 0.4 0.45 0.5
L [µm]
ID [mA]
VGS=0.8V
saturation 0.4
0
It is important to note that 0 0.5 1 1.5
VDsat is not crisply defined VDS [V]
in modern devices 40
– Gradual increase of
gm/gds with VDS 30
gm/gds
20
10
0
0 0.5 1 1.5
VDS [V]
0.6
ID [mA]
0.4
0.2 Computed
2/(gm/ID)
0 values
0 0.5 1
VDS [V]
40
0
0 0.5 1
VDS [V]
4 gm 4mS S
A v 0 ≅ gmRL = 4 ⇒ gm = = 4mS = = 13.3
1kΩ ID 300µA A
A v0 = gm (RL || ro )
−1
1 1
= gm +
RL ro
1 1 1
= +
A v0 gmRL gmro
1 1 1
= +
4 gmRL gmro
gm
High frequency zero ωz = >> ωT
(negligible) Cgd
b1
Nondominant pole ωp2 ≅
b2
L=0.18um
16.9 GHz
= =
.
.
܌
= ܌ = . ∙
. = .
܌܌
= ܌܌ = . ∙
. = .
Device Sizing
−2
gm 2 ID 1 1 2 1 1 gm
Square law: = = µCox VOV = µCox
ID VOV W 2 L L 2 ID
gm ID g
General case: = f ( VOV ) = g ( VOV ) = g f −1 m
ID W
ID
% Specs
Av0 = 4; RL = 1e3; CL = 50e-15; Rs = 10e3; ITAIL = 600e-6;
% Component calculations
gm = Av0/RL;
gm_id = gm/(ITAIL/2);
wT = lookup(nch, 'GM_CGG', 'GM_ID', gm_id);
cgd_cgg = lookup(nch, 'CGD_CGG', 'GM_ID', gm_id);
cdd_cgg = lookup(nch, 'CDD_CGG', 'GM_ID', gm_id);
cgg = gm/wT;
cgd = cgd_cgg*cgg;
cdd = cdd_cgg*cgg;
cdb = cdd - cgd;
cgs = cgg - cgd;
% pole calculations
b1 = Rs*(cgs + cgd*(1+Av0))+RL*(CL+cgd);
b2 = Rs*RL*(cgs*CL + cgs*cgd + CL*cgd);
fp1 = 1/2/pi/b1
fp2 = 1/2/pi*b1/b2
% device sizing
id_w = lookup(nch, 'ID_W', 'GM_ID', gm_id);
w = ITAIL/2 / id_w
ID 300µA
Device width W= = = 18.6µm
ID 16.1A / m
W
Simulation circuit
-vid/2
600 A
1V
Circuit Netlist
* gm/id design example
.op
.ac dec 100 1e6 1000e9
.end
xfmr
= .ends balun
Simulated AC Response
20
214 MHz
0 11.4 dB (3.7)
Magnitude [dB]
-20
5.0 GHz
-40
-60
-80 6 8 10 12
10 10 10 10
Frequency [Hz]
Calculated values: |Av0|=12 dB (4.0), fp1 = 200 MHz, fp2= 5.8 GHz
clear all;
close all;
addpath('/usr/class/ee214b/matlab/hspice_toolbox');
h = loadsig('gm_id_example1.ac0');
lssig(h)
f = evalsig(h,'HERTZ');
vod = evalsig(h,'vod');
magdb = 20*log10(abs(vod));
av0 = abs(vod(1))
f3dB = interp1(magdb, f, magdb(1)-3, 'spline')
figure(1);
semilogx(f, magdb, 'linewidth', 3);
xlabel('Frequency [Hz]');
ylabel('Magnitude [dB]');
axis([1e6 1e12 -80 20]);
grid;
Netlist statement
.pz v(vod) vid
Output
***************************************************
input = 0:vid output = v(vod)
Comparison
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 J. Solid-State Circuits, Sep. 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, Sep. 2002.
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
P. Jespers, The gm/ID Methodology, a sizing tool for low-voltage analog
CMOS Circuits, Springer, 2010.
T. Konishi, K. Inazu, J.G. Lee, M. Natsu, S. Masui, and B. Murmann,
“Optimization of High-Speed and Low-Power Operational
Transconductance Amplifier Using gm/ID Lookup Table Methodology,”
IEICE Trans. Electronics, Vol. E94-C, No.3, Mar. 2011.
Boris Murmann
Stanford University
Winter 2013-14
Outline
Noise in components
– Resistors
– MOSFETs
– BJTs
Noise analysis in circuits
– Output and input referred noise with ideal voltage drive
– Total integrated noise
– Output and input referred noise with finite source resistance
– Equivalent voltage and current generators for arbitrary source
resistance
Signal-to-Noise Ratio
2
Psignal Vsignal
SNR = ∝ 2
Pnoise Vnoise
Topics
− How to model noise of circuit components
− How to calculate/simulate the noise performance of a complete circuit
• In which circuits and applications does thermal noise matter?
Ideal Resistor
i(t)
1V/1kΩ
i(t)
1V/1kΩ
Text, p. 364
Average Power
T/2
1
Pav = ∫
T − T/2
i2 ( t ) ⋅ R ⋅ dt
T/2
1
in2 = lim ∫ in ( t ) ⋅ dt
2
T →∞ T
− T/2
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
PSD ( f ) = n0 = 4 ⋅ kT
Pn 1
v n2 = Pn ⋅ R = 4kT ⋅ R ⋅ ∆f in2 = = 4kT ⋅ ⋅ ∆f
R R
vn2 V2 in2 A2
= 16 ⋅ 10−18 = 16 ⋅ 10−24
∆f Hz ∆f Hz
vn2 in2
= 4nV / Hz = 4pA / Hz
∆f ∆f
2
(
v n2 = vn1 − vn2 ) 2
= vn1 2
+ vn2 − 2 ⋅ v n1 ⋅ vn2
vn2 = vn1
2 2
+ vn2 = 4 ⋅ kT ⋅ (R1 + R2 ) ⋅ ∆f
∑
[C. Enz, EPFL] ∙ ∙∙
PSD of V is equal to Fourier transform
of autocorrelation function
i2d = 4kT ⋅ γ ⋅ gm ⋅ ∆f
[Scholten, 2003]
Measured γ
relative to 2/3
G.D.J. Smit, A.J. Scholten, R.M.T. Pijper, R. van Langevelde, L.F. Tiemeijer, and
D.B.M. Klaassen, "Experimental Demonstration and Modeling of Excess RF
Noise in Sub-100-nm CMOS Technologies," IEEE Electron Device Letters,
vol.31, no.8, pp. 884-886, Aug. 2010.
γ ≅ 0.85
γ ≅ 0.7
vd dd 0 0.9
vm dd d 0
vg g 0 dc 0.7 ac 1 0.9V
mn1 d g 0 0 nmos214 L=0.18u W=10u dd
h1 c 0 ccvs vm 1 c
0V
d
.op Current Controlled
.ac dec 100 10k 1gig Voltage Source
1V/A
.noise v(c) vg
(gm=3.14mS)
2
2 K f gm ∆f
i1/ f =
Cox W ⋅ L f
By definition, the frequency at which the flicker noise density equals the
thermal noise density
2
K f gm ∆f
= 4kTγ ⋅ gm ⋅ ∆f
Cox W ⋅ L fco
Kf 1 gm Kf 1 1 gm ID
⇒ fco = =
4kTγ Cox W ⋅ L 4kTγ Cox L ID W
For a given gm/ID the only way to achieve lower fco is to use longer
channel devices
− In the above expression, both 1/L and ID/W are reduced for
increasing L
Example
– EE214B NMOS, L = 0.18µm, gm/ID = 12 S/A , ⇒ID/W = 20 A/m
⇒fco = 560 kHz
In newer technologies, fco can be on the order of 10 MHz
Just as with white noise, the total 1/f noise contribution is found by
integrating its power spectral density
f2 2
K f gm ∆f
∫ Cox W ⋅ L f
2
i1/ f,tot =
f
1
2
K f gm f K gm2
f
= ln 2 = f 2.3log 2
Cox W ⋅ L f1 Cox W ⋅ L f1
Noiseless!
(merely a modeling
resistor that lets us
account for finite
i2d 2
K f gm 1 dID/dVDS)
= 4kT ⋅ γ ⋅ gm ⋅ +
∆f Cox W ⋅ L f
Gate noise
– "Shot noise" from gate leakage current
– Noise due to finite resistance of the gate material
– Channel-induced gate noise (coupling via Cgs)
• Relevant only at very high frequencies
• See EE314A
Bulk noise
Source barrier noise in very short channels
– Shot noise from carriers injected across source barrier
– R. Navid, C. Jungemann, T. H. Lee and R. W. Dutton, “High-frequency noise
in nanoscale metal oxide semiconductor field effect transistors,” Journal of
Applied Physics, vol. 101(12) , pp. 101-108, June 15, 2007.
– J. Jeon, J. Lee, J. Kim, C. H. Park, H. Lee, H. Oh, H.-K. Kang, B.-G. Park,
and H. Shin, “The first observation of shot noise characteristics in 10-nm
scale MOSFETs,” in Proc. Symp. VLSI Technol., 2009, pp. 48-49.
i2 = 2qID ⋅ ∆f
Constant
(“white”)
PSD
In a bipolar transistor, the flow of DC current into the base and collector
causes shot noise
The noise can be modeled via equivalent current generators
2 1
v out = 4kT ∆f + 2qIC ∆f ⋅ R2
R
(
= 2kT∆f 2R + gmR2 )
2
= 2kTgm∆f ⋅ R2 + 1
gmR
Shot noise due to base current is absorbed by the input source and does
not contribute to noise at the output
For large gain (gmR), the collector shot noise dominates
2
v out = A 2v vin
2
where A v = gmR
We can write
2
2kTgm∆f ⋅ R2 + 1
2
vin = gmR = 2kT 1 ∆f 2 + 1
( gmR )2 gm gmR
Spice Simulation
.op
.ac dec 100 100 10e9
.noise v(vo) vi
.options post brief
.inc ‘ee214_hspice.sp'
.end
2
2 1 1
v out = 4kT ∆f + 2qIC ∆f ⋅ R ||
R jω C
2
1 1
= 4kT ∆f + 2qIC ∆f ⋅ R2
R 1 + jω RC
2
2 1
= 2kTgm∆f ⋅ R2 + 1 ⋅
gmR 1 + jω RC
Same calculation as before, except that now the noise current drops into
the parallel combination of R and C
Output PSD is shaped by squared magnitude of first order response
2 2
2 v out (ω) 2 1
vin (ω) = where A v ( jω ) = A v (0)
A v ( jω )
2 1 + jω RC
2
2 1
2kTgm∆f ⋅ R2 + 1 ⋅
= gmR 1 + jω RC
2
1
A 2v ( 0 )
1 + jω RC
2 1 2
∴ v in = 2kT ∆f + 1
gm gmR
.op
.ac dec 100 100 10e9
.noise v(vo) vi
.options post brief
.inc 'ee214_hspice.sp'
.end
Signal-to-Noise Ratio
1/f tangent
To find out at which frequency
most of the noise rolls in, drop
a 1/f tangent in the log-log PSD
plot (from the top)
The frequency range where the
tangent touches the PSD is the
strongest contributor to the
integrated noise
See text (page 377) for more
info
∞ 2
1
2
v out,tot = ∫ 4kTR ⋅ df
0
1 + j2 π f ⋅ RC
∞
df du
= 4kTR ∫ 2
; ∫ 1 + u2 = tan
−1
u
0 1+ ( 2πfRC )
1
= 4kTR ⋅
4RC
kT
=
C
Interesting result
– The total integrated noise at the output depends only on C (even
though R is generating the noise)
Effect of Varying R
Increasing R increases
the noise power spectral
density, but also
decreases the
bandwidth
– R drops out in the
end result
For C=1pF (example to
the right), the total
integrated noise is
approximately 64µVrms
The equipartition theorem says that each degree of freedom (or energy
state) of a system in thermal equilibrium holds an average energy of
kT/2
In our circuit, the quadratic degree of freedom is the energy stored on
the capacitor
1 1
Cv out 2 = kT
2 2
kT
v out 2 =
C
fENBW
Text, p. 374
∞ 2
1
∫( )
2
v out,tot = 4kTR + 2kTgmR2 ⋅ df
0
1 + j2πf ⋅ RC
2 kT 4kTR + 2kTgmR2
v out,tot = ⋅
C 4kTR
kT 1
= 1 + gmR
C 2
Taking the BJT’s collector shot noise into account, the total integrated
noise becomes a multiple of kT/C
Assumptions
– Output carries a sinusoid with 1V peak amplitude
– We observe the output without significant band limiting and thus use
the total integrated noise in the SNR expression
1 2
v̂ out
Psignal 2 0.5V 2 0.5V 2
SNR = = = = = 8.59 ⋅ 106
Pnoise kT 1 kT 1 2
1+ g R 1 + 3.67mS ⋅ 10kΩ ( 763µV )
C 2 m 10pF 2
(
SNR [ dB] = 10log 8.59 ⋅ 106 = 69.3dB )
Typical system requirements
– Audio: SNR ≅ 100dB
– Video: SNR ≅ 60dB
– Gigabit Ethernet Transceiver: SNR ≅ 35dB
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
– Increase collector current by 4x
Bottom line
– Improving the SNR in a noise limited circuit by 6dB ("1bit")
QUADRUPLES power dissipation !
MDS and DR
Psignal,max
DR =
MDS
If the noise level in the circuit is independent of the signal level (which is
often, but not always the case), it follows that the DR is equal to the
"peak SNR," i.e. the SNR with the maximum signal applied
Rules of thumb
– Up to SNR ~ 30-40dB, integrated circuits are usually not limited by
thermal noise
– Achieving SNR >100dB is extremely difficult
• Must usually rely on external components, or reduce bandwidth
and remove noise by a succeeding filter
• See e.g. oversampling ADCs in EE315B
In our analysis, we showed that the circuit’s noise can be lumped into a
single input referred voltage noise generator
It is important to remember that this result was based on the assumption
of ideal voltage drive
To see why this matters, consider driving the circuit with a current
source The circuit would appear noiseless at the output!
2 1 2
v in = 2kT ∆f + 1
gm gmR
v out gmR
H(s) = =−
vs 1 + sRsCπ
2
v out 2 1
∆f
( )
= 4kTRS + 2qIBRS2 H(s) + 4kT + 2qIC ⋅ R2
R
2
v out 2 1
∆f
( )
= 4kTRS + 2qIBRS2 H(s) + 4kT + 2qIC ⋅ R2
R
2
v out 1 2
v s2 4kT R + 2qIC ⋅ R
= ∆f 2 = 4kTRS + 2qIBRS2 +
∆f H(s) H(s)
2
v 2s 1
= 4kTRS + 2qIBRS2 + 4kT + 2qIC ⋅ R2
∆f R
v 2s g R2 R 1
= 4kT RS + m S + +
∆f 2β ( gmR )2 2gm
Note that this equation has a minimum for a specific value of gm (and thus IC)
Additional Examples
vo g R io gm io g R
Av = = m Gm = = Ai = = m
vi 1 + gmR vi 1 + gmR ii 1 + gmR
Neglecting finite ro, backgate effect and flicker noise for simplicity
2
1 i2d + ir2 2
(
v o2 = i2d + ir2 ) ⋅ = 2 ⋅ Av
gm + 1 gm
R
Often negligible
i2d ir2 1 1
vi2 = 2
+ 2
= 4kT ∆f γ +
gm gm gm gmR
gm
io = id + ( ir − id )
1
gm +
R
gmR 1
= ir + id
1 + gmR 1 + gmR
Gm
= ir GmR + id
gm
i2 i2d 1
io2 d 2 2 2
= 2 + ir R ⋅ Gm vi2 = + ir2R2 = 4kTγ ∆f + 4kTR∆f
gm 2
gm gm
The input referred voltage noise consists of drain current noise, reflected
through gm, plus the resistor’s voltage noise
gmR 1
io = ir + id
1 + gmR 1 + gmR
Ai
= ir A i + id
gmR
Often negligible
i2 i2d 1 γ
io2 = 2 d 2 + ir2 ⋅ A i2 ii2 = + ir2 = 4kT ∆f 1 +
gmR 2 2
gmR R gmR
The input referred current noise from the transistor is often negligible (at
low frequencies)
The noise tends to be dominated by the devices providing the source
and drain bias currents (resistors or current sources)
2
1
+ jω C
1 R
ii2 = 4kT ∆f + 4kTγgm ∆f 2
R gm
2
1 ωC
≅ 4kT ∆f + 4kTγgm ∆f
R gm
The input referred current noise from the transistor can be significant at
high frequencies (near the cutoff frequency of the current transfer)
We will illustrate this method for the purpose of carrying out a general
noise performance characterization (and comparison) of MOSFETs and
BJTs
The procedure can be applied in the same way to any other device or
circuit
To find input referred voltage generator, short the input of both circuit
models and equate output noise
Neglecting Cµ, rc and re for simplicity
2
io1 ≅ ic2 + gm (
2 2 2
ib ⋅ rb + vb2 )
2 2 2
io2 ≅ gm vi
2 2 ic2
io1 = io2 ⇒ vi2 ≅ vb2 + 2
+ ib2 ⋅ rb2
gm
vi2 2qI
≅ 4kTrb + 2C + 2qIBrb2
∆f gm
2qIC gm 2 2
rb
≅ 4kTrb + 1 +
2
gm β
2qIC
≅ 4kTrb + 2
gm
1
≅ 4kT rb +
2gm
To find input referred current generator, open circuit the input of both
circuit models and equate output noise
2 2
io1 = ic2 + gm
2 2
⋅ ib ⋅ z π
2 2 2 2
io2 = gm ⋅ ii ⋅ z π
2 2 ic2
io1 = io2 ⇒ ii2 = ib2 + 2
2
gm zπ
2qIC
= 2qIB + 2
β ( jω )
2
1+ ω
ωβ ω
= 2qIB 1 + where ωβ =
1
≅− T
β0 rπCπ β0
The term due to IC is negligible at low frequencies, but becomes
comparable to the base current contribution at
ωT
ωb = ωβ β0 ≅
β0
2
io1 = ic2
2 2 2
io2 = gm vi
ic2 vi2 1 Kf 1
vi2 = 2
= 4kTγ +
gm ∆f gm WLCox f
2
2 1 ω
io1 = ic2 2
+ gm ⋅ i2g ⋅ ≅ ic2 + ig2 ⋅ T
ω2Cgs
2
ω
2
2 ω
io2 ≅ ii2 ⋅ T
ω
2 2
ω 2 ii2 f 2
K f gm 1
ii2 = ig2 + ic ≅ 2qIG + 4kTγgm +
ω ∆f f WLCox f
T T
Noiseless
Transistor
v2 1 v2 1 Kf 1
i ≅ 4kT rb + i ≅ 4kTγ +
∆f 2gm ∆f gm WLCox f
BJT MOS
i2 i2 2
f
IC K g2 1
i ≅ 2q IB + i ≅ 2qIG + 4kTγgm + f m
f
∆f β ( jω ) ∆f
2
MOS fT WLCox
BJT
Boris Murmann
Stanford University
Winter 2013-14
Text, p. 243
Example:
Switched Capacitor Gain Stage
Example:
Two-Stage Operational
Transconductance Amplifier
(OTA) Text, p. 243
OTA
B. Murmann EE214B Winter 2013-14 – Chapter 7 5
Example: TIA
C. Knochenhauer et al., “40 Gbit/s transimpedance amplifier with high linearity range in 0.13 µm SiGe
BiCMOS, Electronics Letters, May 12, 2012.
After developing the key tools using the TIA example, we will look into
other examples and further generalizations to broaden your
understanding
EE315A will then continue on this track by looking at “mainstream” OTA
design for filters and sensor interfaces
+ Sε
Si Σ a So
–
Sfb
f
So = a ⋅ Sε
Sfb = f ⋅ So ⇒ So = (Si − Sfb ) = a(Si − f ⋅ So )
Sε = Si − Sfb
So a
Closed-Loop Gain: = A=
Si 1 + af 1 T
⇒ A=
S f 1+ T
Loop Gain: T = af = fb
Sε
If T >> 1, then
1
A≅ (Also, note that if T << 1, then A ≅ a)
f
For large loop gain, the feedback acts to minimize the error signal
(Sε), thus forcing Sfb to track Si
a af
Sε = Si − f ⋅ So = Si − f ⋅ Si = 1− ⋅S
1+ af 1+ af i
Sε T 1 Sfb S T
∴ = 1− = and = a⋅f ε =
Si 1+ T 1+ T Si Si 1+ T
dA d a 1 d 1
= = +a
da da 1+ af 1+ af da 1+ af
(1+ af) − af 1 1
= 2
= 2
=
(1+ af) (1+ af) (1+ T)2
For a change δa in a
dA δa
δa = δa =
da (1+ T)2
δA δa 1+ T 1 δa
∴ = =
A (1+ T)2 a 1+ T a
Closed-Loop Impedances
io
vε avε
vi vo
Zi = Zo =
ii io = 0 io
vi =0
With io = 0 With vi = 0
v o = av ε v ε + fv o = v i = 0
v i = v ε + fv o = (1+ af)v ε v o − av ε 1
= (1+ T)v ε
io =
zo
=
zo
( )
1+ af v o
vε 1 1 1
ii = = v (1+ T)v o
zi 1+ T i
=
zi zo
vi vo z
Zi = = z i (1 + T) Zo = = o
ii io 1+ T
1 + T(port shorted)
Zport = Zport ( gain set to zero ) ⋅
1 + T(port open)
a0
Example : a(s) =
s
1−
p1
a(s) a0 1 1 T0 1
A(s) = = ⋅ = ⋅
1 + a(s)f 1 + a0 f 1 − s f 1 + T0 1 − s
p1 (1 + a0 f ) A0 p1 (1 + T0 )
20 log10 a0
20 log10 |a(jω)|
20 log10 (1+T0)
20 log10 |A(jω)|
20 log10 A0
|p1|
ω3dB = (1+T0)|p1|
a0 f
a( jω3dB ) = A 0 T( jω3dB ) = a( jω3dB ) ⋅ f = A 0 f = ≅1
1 + a0 f
G2
G1
ωu=G1BW1=G2BW2
1 ω
BW1 BW2
Instability
Phase margin
– Defined at the frequency where |T(jω)| =1 ω0
Gain margin
– Defined at the frequency where Phase[T(jω)] = -180° ω180
1
GM =
T ( jω180 )
Example
20 log10 A0
Gain Margin
ω0 | p2 |
Phase Margin
1
T( jωu ) = a( jωu ) ⋅ f = 1 ⇒ a( jωu ) = (assuming f is real)
f
a( jωu )
A( jωu ) =
1 + a( jωu ) a( jωu )
a( jωu ) a( jωu )
= jφ[a( jωu )]
=
1+ e 1 + e j(PM−180° )
PM = 45o
a( jωu ) a( jωu )
A( jωu ) = − j135°
=
1+ e 1 − 0.7 − 0.7j
a( jωu ) 1.3
∴ A( jωu ) = = ≅ 1.3A 0 “Peaking”
0.76 f
a( jωu ) a( jωu )
PM = 60o A( jωu ) = =
1+ e − j120°
1 − 0.5 − 0.87j
1
∴ A( jωu ) = a( jωu ) = ≅ A0 “Flat”
f
|A|/A0 ω/ω u
Frequency Compensation
↑ (1+T)
Bandwidth Approximately equal to unity gain
frequency of loop gain
↓ (1+T)
Gain Defined primarily by the
feedback network ~1/f
Boris Murmann
Stanford University
Winter 2013-14
C. Knochenhauer et al., “40 Gbit/s transimpedance amplifier with high linearity range in 0.13 µm SiGe
BiCMOS, Electronics Letters, May 12, 2012.
25
∙ ∙
∙ 34
∙ 5.8
2 2
The smallest signal we can receive while maintaining the given BER is
.
∙ 10 40.6
This number then defines the minimum optical input power, maximum
length of the fiber, etc., typically in compliance with a certain standard
మ
= 4 =
1
= = 442
2
Assuming that transimpedance gain is fixed, both noise and f3dB are
fixed and beyond our control
– The noise happens to be much better than needed
– The bandwidth happens to be much worse than needed
This circuit lacks degrees of freedom that let us adjust gain, bandwidth
and noise independently (within some reasonable range)
= ≅
2( + + ) 2
= 4 + 2 =
1 19
Δ
Good news
– We are now meeting the gain, bandwidth and noise specification
Bad news
– The circuit has very small signal headroom: VR =1.1mA·1800Ω ≅ 2V
– The circuit cannot handle the maximum signal of 1.3mApp
– Any capacitance we connect to the output will kill the bandwidth
– Whatever we can do to try and fix these issues will drive us outside
the noise specification?
B. Murmann EE214B Winter 2013-14 – Chapter 8 6
Commonly used Feedback TIAs
CS CD
Sung Min Park and Hoi-Jun Yoo,
"1.25-Gb/s regulated cascode
CMOS transimpedance amplifier
for Gigabit Ethernet applications,"
IEEE J. Solid-State Circuits, pp.
112-121, Jan. 2004
CG “Regulated Cascode”
Optimality
Attractive property
– Both Rin and Rout are low, e.g. ~1/gm, for RD >> RF (prove this)
−
−
0= 1
1−
=
−
= −
+ +
1
1+
0=
−
=− =
=− =−
= −i R
=
v
= −
i
→
= i r = =
v
i
1
1−
=
1+ 1+ 1 +
1 +
+ = − + = −
1
1+
The result for the closed loop gain (A) matches the nodal analysis
expression perfectly
→
f 1
A
→ TA =
The two-port modeling approach insists on the “af” two-box model and
looks to identify an idealized “f” block such that
– The feedback block does not load (or alter) the forward gain
– The feedback signal sums perfectly with the input (as in the “af” model)
Key to achieving the latter is the controlled source y12, which subtracts
from the input; the other elements in the block are needed for equivalence
= +
= +
This requires
= =
1 1
= =−
మ భ
= =
1 1
=− =
మ భ
Final step
– Absorb all elements other than the ideal feedback source into the “a” block
a
RD
vo iε
+
RF ii a vo
v1 –
ii RF ifb
-v1/RF
f
-vo/RF vo
f
= ∙ భ = − − =−
ూ ీ 1
R
భ ಷ ూ ీ
1
1 1−
= = − =
1 R
R
R
+ R 1+
= −
1
1+
Interpretation
The obtained closed-loop gain matches the result from KCL and return
ratio analysis perfectly
However, the loop gain T that we obtained from the two-port model does
not match the return ratio result
This is to be expected, because the two-port model lumped the
feedforward effect into the forward path of the loop; but it is in reality not
part of the loop (see closed-loop return ratio model)
– The return ratio result is the “golden standard” for the purpose of loop
stability analysis
So then, why bother with the two-port model?
– It turns out that the two-port model is proper as long as feedforward
is insignificant
• In any practical circuit this is consistent with our design intent –
we want to build a circuit whose forward gain depends mostly on
the amplifier, not on an unwanted path
– The two-port model is “nice,” since everything we know about “af”
two-box models can be directly applied
Assuming RF >> RD
= − − ≅ −
ూ ీ
R
ూ ీ
ಷ
= ≅
Generalization
absorb
in “a”
absorb
in “a”
absorb
in “a”
Port Impendaces
vo
M1
ii CD
RF
Two-Port Model
=−
1
= −
1
1 +
∙
= =
≅
1
=
1 +
≅
≅
=
1 1 1 1
1 +
≅
Closed-Loop Bandwidth
= −
1
1 +
∙
= =
≅
, = (1 + ) ,
≅
1 1
, =
4
1
∆
4! ∆
4! ∆
4
1
4
1
∆
∆
All noise sources, except the noise of M1 and RF at the input are
negligible in practice
We already saw in our CE noise example that the RD noise is negligible
as long as gmRD is large
– All other noise sources are irrelevant by the same argument
4! ∆
4
1
∆
4!
1
∆
4
1
∆
4 4! + "
1 1 1
∆ ∆
≅ 4∆ +!
1
4!
1
4
1
∆
∆
+
+
≅ 4∆
+! ≅ 4∆ +!
1 1
≅ 4∆
1
+ 4!
We can now solve for the frequency where the noise of M1 takes over
= 4!
1
!
! =2
RD
M2
vo
M1
ii CD
RF
Cgd2
RD
Cgd1 Cdb1
M2
Csb2
M1 Cgs2
vo
Cgs1
ii
CD
RF
Cgd2
RD
Cgd3 Cdb3
M2
Csb2
Cgs2
Cgs3
Cgd1
Cdb1+Csb3
M1 vo
Cgs1
ii
CD
RF
RD
= −
1
M2
C2
= −
C4 1
M3
IB IB IB
= 1$
M2
= 1Ω
RD vo
M3
= 2Ω
W/5 IB
= 200%
&& * = 0.18 $
M1
ii CD RF
&& /
= 10'/
/( = 29.3/$ && ( = 34 $
Operating Point
gm
A v2 := = 0.758 T0 := g m⋅ RD⋅ A v2 = 15.152
1
g m + g mb +
Rf
1 1
C1 := CD + Cgs + 2⋅ Cgd = 281⋅ fF fp1 := ⋅ = 566.388MHz
⋅
2⋅ π Rf ⋅ C1
Relevant
1 1
( )
C2 := 2Cgd + 1 − A v2 ⋅ Cgs + Cdb = 63.879fF
⋅ fp2 := ⋅
2⋅ π RD⋅ C2
= 1.246⋅ GHz
1 g m + g mb Somewhat
C3 := Cgs + Csb + Cdb = 94⋅ fF fp3 := ⋅ = 20.656GHz
⋅
2⋅ π C3 relevant
1 g m + g mb Irrelevant
C4 := Csb = 25⋅ fF fp4 := ⋅ = 77.668GHz
⋅
2⋅ π C4
566MHz =
15.1
1+ 1+
1.24GHz $ $
≅
≅ $ $
$ $
≅ 3.2+
20.6GHz
, = 180° − -. − -. − -.
$ $ $
, = 180° − 80° − 69° − 12 = 19°
Magnitude [dB]
gain_margin(dB) =
0 8.711791
-20
phase_margin(deg) =
-40 20.96923
-60
phase_margin_freq(Hz) =
-80 8 10
2.7774433E+09
10 10
Frequency [Hz] gain_margin_freq(Hz) =
4.6721000E+09
0
Phase [degrees]
loop_gain_at_min_freq(dB) =
-100 23.16820
-200
See appendix for
-300 setup of LSTB
analysis in HSpice
8 10
10 10
Frequency [Hz]
Closed-Loop Response
60
Magnitude [dB]
40
20
0 7 8 9 10 11
10 10 10 10 10
Frequency [Hz]
=
1+ 1+
$ $
1 1
=
1
1+ 1 +
= ∙
+
$ + $ 1
(1 + ) (1 + )
1+
$ $ $ $
=
1+ /+
(1 + )
(1 + ) /=
$ $
= $ $ ≅
$ + $
Q=0.5
5 Q=1/sqrt(2)
Q=1
0 Q=2
Magnitude [dB]
-5
-10
-15
-20 -2 -1 0 1
10 10 10 10
ω /ω 0
H = |H = " |=/
1
1+ /+
10 0 0
Location in the s-plane
/
for Q > 0.5
s1 jω
Q=2
|H(s)|
Imag(s)/ω0
Re(s)/ω0
s-plane
s = σ + jω
|H(s)|
Imag(s)/ω0
Re(s)/ω0
s-plane
s = σ + jω
Q = 0.5
|H(s)|
Imag(s)/ω0
Re(s)/ω0
s-plane
s = σ + jω
ωu/ω p2
ωu/ω p2
Text, p. 219
“underdamped”
Q > 0.5
Q=0.5
Q < 0.5
“overdamped”
Text, p. 164
ω0
45º
σ
Getting the phase margin we need for an MFM response would require
that one of the poles occurs beyond the unity gain frequency of the loop
Leave the poles of a(s) untouched and introduce a left half plane zero in f(s),
close to the unity gain frequency of the loop
Top first order, the phase margin will be 45 degrees (instead of 0 degrees)
when the zero is placed at the unity gain frequency of the loop
Two questions
– How can we realize the zero?
– How exactly should we position it to get a maximally flat response?
= +
= +
= + = +
1 1
= =−
మ భ
= + = +
1 1
=− =
మ భ
RD
M2
C2
M3 C4
= −
1
( + )
C3
vo
M1
" = −
ii C1 RF CF RF CF (" + )
1 +
= = 1 +
&
1+ 1+ &
$ $
1
=
1 1
() 1 + 1 +
= ∙
1+ /+
= (1 + ) $ $ ≅ Unchanged
1 $ + $
/
= + New degree of freedom
&
1 $ + $
/
= + = 2
&
= ≅
2 −
&
$ + $ 2
& =
3.27+
566 + 1.24+
= 3.8+
2 − 3.27+
= = 42%
1
28&
-1.66988g -2.73938g
-10.7624g 0.
40 -20.3840g 0.
-23.3910g 0.
30
zeros ( hertz)
20 real imag
15.5276g 0.
10
-20.1601g -12.6203g
-20.1601g 12.6203g
0 7 8 9 10 11
10 10 10 10 10 -23.3910g 0.
Frequency [Hz]
-8.09147g 0.
40 -20.3512g 0.
-23.3910g 0.
30
zeros (rad/sec)
20 zeros ( hertz)
real imag
10 14.0106g 0.
-18.6215g -11.4893g
0 7 8 9 10 11
-18.6215g 11.4893g
10 10 10 10 10 -23.3910g 0.
Frequency [Hz]
The poles are still not exactly angled at 45 degrees, but this is OK
– A high frequency pole (not analyzed) helps reduce any residual peaking
gain_margin(dB) =
0 24.02825
-20
phase_margin(deg) =
-40 55.81567
-60
phase_margin_freq(Hz) =
-80 8 10
3.1666851E+09
10 10
Frequency [Hz] gain_margin_freq(Hz) =
1.9178802E+10
0
Phase [degrees]
loop_gain_at_min_freq(dB) =
-100 23.16766
References
ir
T ( s) = −
it
[HSpice manual]
vr
T ( jω ) ≅ −
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
available breakpoint
Z1 ⋅ Z2
T(s) = gm
Z1 + Z2
vy Z2
− ≡ Tv = gm ⋅ Z2 +
vx Z1
Solving yields:
1 1 1
True Loop T = g ⋅ Z1Z2 = +
Gain: m
Z1 + Z 2 1 + T 1 + Tv 1 + Ti
Tv Ti − 1
iy T =
Z1 Tv + Ti + 2
≡ Ti = gm ⋅ Z1 +
ix Z2
LSTB Syntax
IB IB IB
M2
RD vo
M3
W/5 IB
M1
ii CD RF vx
[Hurst, 1994]
[Bode 45]:
[Bode 45]:
Boris Murmann
Stanford University
Winter 2013-14
Advantages
• No instability
• Lower gain sensitivity to component and device parameter
variations than amplifiers without local feedback
Disadvantages
• Higher gain sensitivity (less loop gain) than amplifiers with
multi-stage feedback
• Bandwidth obtainable is typically slightly less than that
possible in amplifiers with multi-stage feedback
BW ~ 6 GHz
RF
For gmRE >>1
ܴி
RE ܣ௩ ≅
ܴா
Shunt F/B
Stage “ratiometric”
Series F/B
Stage
RS Cµ io
+
vi ~ v1 rπ Cπ gmv1 ZL
–
+
vE RE
–
Include transistor rb in RS
v1 1
∴ =
v i 1+ (gm + Yπ ) RE + YπRS
Since io = –gmv1
io gm
=−
vi 1+ gmRE + (RS + RE )Yπ
gm 1
= −
1+ gmRE R + RE 1
1+ S + sC π
1+ gmRE rπ
gm 1
= −
1+ gmRE R + RE 1 RS + RE
1+ S 1+ g R + sC π
rπ m E 1+ gmRE
RS + RE 1 gm (RS + RE ) 1
r 1+ g R = β0 << 1
π m E 1+ gmRE
Then
io 1
≅ −gmeq
vi 1− s p1
where
gm
gmeq =
1+ gmRE
1+ gmRE 1
p1 = −
C π RE + RS
If gmRE >> 1
1
gmeq ≅
RE
and
gm RE Cπ + Cµ RE RE
p1 ≅ − = − ωT ≅ − ωT
Cπ RE + RS Cπ RE + RS RE + RS
io
RS
+ Q1
vi ~ Not a large bypass
– capacitor
RE CE
Cπ
Then, defining τE = RECE and assuming τT ≅
gm
io gm
=−
vi 1
1 + gmZE + (RS + ZE )( + sCπ )
rπ
gm
=−
RS RE 1
1+ + sCπRS + gm + + sCπ
rπ 1 + sτE rπ
gm
≅−
R g R C
1 + S + sCπRS + m E 1 + s π
rπ 1 + sτE gm
1
sin ce gm >>
rπ
If CE is chosen so that
τE = τT
Then
io gm 1
≅ −
vi 1+ gmRE RS
1+ sC π
1+ gmRE
1+ gmRE R
p1 = − ≅ −ω T E
RS C π RS
CF
RF
+ +
ii v1 rπ Cπ gmv1 RL CL vo
– –
Include transistor Cµ in CF
Neglect RS or include it in rπ
Neglect rb, rc, re, and rµ
Include ro and Ccs in RL and CL
Analysis (1)
We can analyze this circuit from first principles using KCL and then simplify
to get to a low-entropy result. However, it is much faster to look the circuit
using a two-port feedback approach, and use our understanding from
chapter 8.
= = 1 +
௭
1+ 1+
ଵ ଶ
= = (1 + )ଵ ଶ ≅ ௨
ଶ
1+
+ ଶ
1 ௨
For a maximally flat response (Q=0.707), set: ௭ = ≅
ி ி 2
Model of a(s):
+ +
ii v1 gmv1 vo
– –
rπ||RF Cπ||CF RL||RF CL||CF
1 1
= = −ி ( ||ி ) ଵ = ଶ =
R Cగ ( | ி
1 1
௨ ≅ ଵ ଶ = ∙ ≅ ் ∙
R Cగ R
Analysis (3)
1
ଷௗ = ≅ ௨ ≅ ் ∙
R
Further insight can be gained from this result by considering the input
capacitance of the overall Cherry-Hooper amplifier
గ 1
≅ = = k = “fanout”
ா ் ா
1 ்
ଷௗ ≅ ் ∙ =
௩
R
் ா
Boris Murmann
Stanford University
Winter 2013-14
Root Locus
s-plane
x x σ
(1+T0)p1 p1
A so-called “root locus” plot shows the movement of the poles in the s-
plane as we vary the low-frequency loop gain T0
Root locus plots are most commonly used in control theory, but can
provide valuable intuition for the design of amplifiers and other electronic
circuits
1
H =
ଶ > 0.5 ଵ,ଶ = − 1 ±
4ଶ − 1
1+ + ଶ 2
≤ 0.5 ଵ,ଶ =− 1 ± 1 − 4ଶ
2
= (1 + )ଵ ଶ
jω
(1 + )ଵ ଶ
= s-plane
ଵ + ଶ
p2 p1
σ
= 0.5
ଵ + ଶ
=− =−
2 2
3 3
s s
1 − + T0 = 0 1 − = −T0
p1 p1
(
s1 = p1 1 + 3 T0 )
s2 = p (1 −
1
3 T0 e j60° )
s3 = p (1 −
1
3 T0 e − j60° )
0 = 1 − Re ( 3 T0 e j60° )
0 = 1 − 3 T0 cos(60°)
⇒ T0 = 8
Generalization
In general,
and
1+ c1s + c 2s2 + ⋅⋅⋅ Nf (s)
f(s) = f0 2
= f0
1+ d1s + d2s + ⋅⋅⋅ D f (s)
Thus,
a0 Na (s)D f (s)
A(s) =
Da (s)D f (s) + T0 Na (s)Nf (s)
where T0 = a0f0.
a0 Na (s)D f (s)
T0=0 A(s) =
Da (s)D f (s) + T0 Na (s)Nf (s)
a0 Na (s)D f (s)
T0→∞ A(s) =
Da (s)D f (s) + T0 Na (s)Nf (s)
where
za1,za2 ,... = zeros of a(s)
zeros of T(s)
zf1,zf 2 ,... = zeros of f(s)
pa1,pa2 ,... = poles of a(s)
poles of T(s)
p f1,pf 2 ,... = poles of f(s)
p p ⋅ ⋅⋅ p f1 p f 2 ⋅⋅⋅
T0 ⋅ a1 a2
z a1 z a2 ⋅ ⋅⋅ z f1 z f 2 ⋅⋅⋅
(s − za1)(s − za2 ) ⋅⋅⋅ (s − zf1)(s − zf 2 ) ⋅⋅⋅
× = −1
(s − pa1)(s − pa2 ) ⋅⋅⋅ (s − p f1)(s − pf 2 ) ⋅⋅⋅
Phase Condition
Magnitude Condition
and
The rules for constructing the root locus are based on the phase
condition. The magnitude condition determines where, for a
given T0, the poles of A(s) actually lie on along the locus.
Rule 1: Branches of the root locus start at the poles of T(s), where T0 = 0, and
terminate on the zeros of T(s), where T0 = ∞. If T(s) has more poles
than zeros, some branches terminate at infinity.
Rule 2: If T(s) has all its zeros in the LHP or if T(s) has an even number of
RHP zeros, the locus is situated along the real axis wherever there is
an odd number of poles and zeros of T(s) to the right. If T(s) has an
odd number of RHP zeros, the locus is situated along the real axis
wherever there is an even number of poles and zeros of T(s) to the
right.
Rule 3: All segments of the locus on the real axis between pairs of poles, or
pairs of zeros, must branch out from the real axis.
Rule 4: The locus is symmetric with respect to the real axis because complex
roots occur only in conjugate pairs.
Rule 6: Branches break away from the real axis at points where
the vector sum of reciprocals of distances to the poles of
T(s) equals the vector sum of reciprocals of distances to
the zeros of T(s).
σa =
[ ] [
Σ poles of T(s) − Σ zeros of T(s) ]
Np − Nz
T0
Problem statement
– For the given low-frequency loop gain T0 and high-frequency pole
ωp2, find the proper value of ωp1 that gives a maximally flat magnitude
response for the closed-loop amplifier
ଵ ଶ
1
ଵ |
ଶ |
T0
Problem statement
– For the given low-frequency loop gain and poles of a(s), find the
proper position of the feedback zero that gives a maximally flat
response
From rules 1 and 2, we know that the root locus lies between p1 and p2
and also to the left of zF
Applying rule 6 gives two breakpoints, one between p1 and p2, and one
where the poles return to the real axis after circling around z
In general, the locus tends to bend toward zeros as if attracted and
tends to bend away from poles as if repelled
2
s1 − p1 s2 − p2 ω02 2 zF
T0 = = =
p1 p2 p1 p2 p1 p2
1
ωz = T0ωp1ωp2
2
“Phantom Zero”
s = tf('s');
p1=-1; p2=-2; p3=-4;
T = 1 / [(1-s/p1)*(1-s/p2)*(1-/p3)]
rlocus(T)
Root Locus
10
2
Imaginary Axis
-2
-4
-6
-8
-10
-14 -12 -10 -8 -6 -4 -2 0 2 4
Real Axis
Adding a Zero
s = tf('s');
z=-5; p1=-1; p2=-2; p3=-4;
T = (1-s/z) / [(1-s/p1)*(1-s/p2)*(1-s/p3)]
rlocus(T)
Root Locus
15
10
5 This example
visualizes how
Imaginary Axis
0
introducing a zero in
T(s) can be used to
-5
stabilize a feedback
amplifier with three
poles.
-10
-15
-6 -5 -4 -3 -2 -1 0 1
Real Axis
Transconductor
-gm (e.g. CS or CE stage, differential pair)
gm Current Buffer
(CG or CB stage)
1/gm
1/gm
Voltage Buffer
Av 1
(CD or CC stage)
T0
Idea
– Make one of the loop poles dominant, leave other poles unchanged
T0
Benchmarking
≫ 1
=
1 +
(
)
≅
≅
≪ 1
=
T
=
1− 1− … 1−
s୬
1 + ⋯+ p p …p
ଵ ଶ ଵ ଶ ୬
The product of the low frequency loop gain and all dominant poles is
called the loop gain pole product (LP product)
– Nordholt, Design of High-Performance Negative-Feedback
Amplifiers, 1983
Note that in a first-order system, the LP product is simply the gain-
bandwidth product
T0 T0
ଶ
≅
ଵ
ଶ
௨ ≅ ଵ ଶ ௨ ≅ ଵ ଶ
To first order, since we do not change the poles, the LP product and ωu
are (approximately) unchanged
– Feedback zero compensation is therefore bandwidth efficient, since
we do not need to sacrifice bandwidth to stabilize the circuit
To second order, and LP product will change slightly due to loading from
the capacitance added in the feedback network
T0
௨ ≅
ଵ
ଶ
௨ᇱ ≅
ଵ ≅
ଶ
1
2
-gm gm
Cgs CL
1/gm
Telescopic OTA
(Common mode
feedback circuits not
shown – see EE315A)
Two-Stage OTA
Cc
gm1 -gm2
C1 R1 C2 R2
1
p1 ≅ −
R1 C1 + Cc (1 + gm2R2 ) + R2 (C2 + Cc )
gm2
z=+
CC
R1 C1 + Cc (1 + gmR2 ) + R2 (C2 + Cc )
p2 ≅ − RHP zero
R1 R2 (C1C2 + C1Cc + C2Cc )
c c
a(s):
-gm2
C1+Cc R1 C2+Cc R2
Mag ( jω ) gm2R2R1
a(s) a(s) = −
(1 + sR1 [C1 + Cc ]) (1 + sR2 [C2 + Cc ])
1
f(s) = −sCf
f(s)
ω
Unfortunately, the right half plane zero due to Cc can destroy the PM
RHP LHP
ωz −ωz
ω ω
1− j → − 90 1+ j → + 90
ωz ωz
Phase Phase
ωz ω
0° + 90°
− 45° + 45°
− 90° 0°
ωz ω
Nulling Resistor
Rz Cc
gm1 -gm2
C1 R1 C2 R2
• p1 and p2 unchanged, new pole p3, and a “knob” to tune the zero
1
ωp3 ≅
R z C1
gm2
For Rz = 1/gm2 ωp3 ≅ ≅ ωT
C1
gm2 ωT
For Rz = (1+C2/Cc)/gm2 ωp3 ≅ ≅
C2 C2
1+ C1 1+
Cc Cc
Thus, as we try to cancel the second pole, the third pole moves to a
lower frequency, and may move to a frequency that is comparable to
the original ωp2 before cancellation
My recommendation: Simply push the zero to infinity
The textbook’s recommendation: Spice-monkey the zero into the LHP
and try to squeeze out some phase margin
gm=KVOV
Text, page 260
Ron=1/(KVOV)=1/gm
gma
Cc
gm1 -gm2
C1 R1 C2 R2
1
=−
1 1
+
=−
1 +
Mag ( jω ) Mag ( jω )
a(s) a(s)
1 1
f(s) gma/Cc f(s)
ω ω
gma/Cc
Implementation Options
Mag ( jω ) Cm2
Cm1
gm1, gm2
Feedforward Compensation
ω
Very large achievable bandwidth
|p1| |p2| |z1|
Potential issue
φ(jω)
ω
– The doublet p1, z1 can make it
– π/2 difficult to achieve a fast
–π transient response
• See Kamath, JSSC 12/1974
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.
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.
Summary
All of our analyses have been concerned with the small-signal behavior
of feedback amplifiers at high frequencies
Sometimes the behavior with large input signals (either step inputs or
sinusoidal signals) is also of interest
– See e.g. switched capacitor circuits in EE315A
Consider a step applied to an amplifier in unity gain feedback
Slew Rate
Boris Murmann
Stanford University
Winter 2013-14
Introduction
Io = IOQ + io io
+ gm⋅vi
+
Vi = VIQ + vi vi
-
dIo -
gm =
dVi V = V
i IQ
Io = f(Vi) = f '(VIQ ) io
gm gm
vi
Vi
VIQ
(3)
f '(VIQ ) f ''(VIQ ) 2 f (VIQ )
f(Vi ) = f(VIQ ) + (Vi − VIQ ) + (Vi − VIQ ) + (Vi − VIQ )3 + ...
1! 2! 3!
f2(Vi)
f3(Vi)
f(Vi)
f2(Vi)
Vi
VIQ
f3(Vi)
f (m) (VIQ )
where am =
m!
1 ' 1 ''
Note that a1 ≡ gm a2 ≡ gm a3 ≡ gm
2 6
Graphical Illustration
io
n=2
vi
n→∞
n=3
A model that relates the incremental signal components (vi, io) though a
nonlinear expression is sometimes called “large-signal AC model”
The accuracy of a truncated power series model depends on the signal
range and the curvature of the actual transfer function
– Using a higher order series generally helps, but also makes the
analysis more complex
– As we will see, using a third order series is often sufficient to model
the relevant distortion effects in practical, weakly nonlinear circuits
Apply a sinusoidal signal and collect harmonic terms in the output signal
v i = vˆ i ⋅ cos ( ω t )
2 3
io = a1vˆ i cos ( ω t ) + a2 vˆ i cos ( ω t ) + a3 vˆ i cos ( ω t ) + ...
1 1
cos2 ( α ) = cos ( 2α ) + 1 cos3 ( α ) = cos ( 3α ) + 3 cos ( α )
2 4
1 DC shift
∴ io = a2 vˆ i2
2
3
+ a1vˆ i + a3 vˆ i3 cos ( ω t ) Fundamental
4
1 1
+ a2 vˆ i2 cos ( 2ω t ) + a3 vˆ i3 cos ( 3ω t ) + ... Harmonics
2 4
Observations
The quadratic term (a2) give rises to an undesired second harmonic tone
and a DC shift
The cubic term (a3) give rises to an undesired third harmonic tone and it
also modifies the amplitude of the fundamental
– a3 < 0 “gain compression”
– a3 > 0 “gain expansion”
y x + 0.3x3
x
x - 0.3x3
m
1 m 1 m − j( m − k ) α
cosm ( α ) =
2 m ( e jα + e − jα ) =
2m ∑k
e jkα e
k =0
1
cos4 ( α ) = cos ( 4α ) + 4cos ( 2α ) + 3
8
1
cos5 ( α ) = cos ( 5α ) + 5 cos ( 3α ) + 10cos ( α )
16
1 3 As long as
∴ io = a2 vˆ i2 + a4 vˆ i4
2 8
a 4 vˆ i4 << a2 vˆ i2 and a5 vˆ i5 << a3 vˆ i3
3 5
+ a1vˆ i + a3 vˆ i3 + a5 vˆ i5 cos ( ω t )
4 16
or equivalently
1 1
+ a2 vˆ i2 + a 4 vˆ i4 cos ( 2ω t )
2 2 a2 a3
vˆ i << and vˆ i <<
1 5 a4 a5
+ a3 vˆ i3 + a5 vˆ i5 cos ( 3ω t )
4 16
the 4th and 5th order terms can
1 be neglected
+ a4 vˆ i4 cos ( 4ω t )
8
This condition is usually met in
1
+ a5 vˆ i5 cos ( 5ω t ) practical, weakly nonlinear
16 circuits
2
+a2 vˆ i1 ⋅ cos ( ω1t ) + vˆ i2 ⋅ cos ( ω2 t )
3
+a3 vˆ i1 ⋅ cos ( ω1t ) + vˆ i2 ⋅ cos ( ω2 t ) + ...
Second-Order Intermodulation
2a2 vˆ i1vˆ i2 ⋅ cos ( ω1t ) cos ( ω2 t ) = a2 vˆ i1vˆ i2 cos ({ω1 + ω 2 } t ) + cos ({ω1 − ω 2 } t )
The output will contain tones at the sums and differences of the applied
frequencies
We define the fractional second-order intermodulation as
a2 vˆ i2 a2
= = v̂i
a1 vˆ i a1
= 2HD2
3 3
a3 vˆ i1 ⋅ cos ( ω1t ) + vˆ i2 ⋅ cos ( ω 2 t ) = a3 vˆ i1 ⋅ cos ( ω1t )
Causes HD
3
+a3 vˆ i2 ⋅ cos ( ω 2 t )
+3a3 vˆ i1vˆ i2
2
⋅ cos ( ω1t ) cos2 ( ω2 t )
New
+3a3 vˆ i1
2ˆ
v i2 ⋅ cos2 ( ω1t ) cos ( ω2 t )
3a3 vˆ i1vˆ i2
2
⋅ cos ( ω1t ) cos2 ( ω 2 t )
3 2
=
4
a3 vˆ i1vˆ i2 2cos ( ω1t ) + cos ({2ω2 − ω1} t ) + cos ({2ω 2 + ω1} t )
3 a3 vˆ i3 3 a3 2
= = v̂i = 3HD3
4 a1 vˆ i 4 a1
Vbe Vbe
kT dIc I
Ic = Ise VT VT = a1 = = s e VT
q dVbe VT
Vbe = VBEQ
Vbe = VBEQ
ICQ
= ≡ gm
VT
1 d2Ic 1 ICQ
a2 = 2
=
2 dVbe 2 VT2
Vbe = VBEQ
1 ICQ
am =
m! VTm
2
1 a2 1 vˆ be 1 a3 2 1 vˆ be
HD2 ≅ vˆ be = HD3 ≅ vˆ be =
2 a1 4 VT 4 a1 24 VT
Low distortion in the collector current requires the B-E voltage excursion
to be much smaller than VT ≅ 26mV
Checking for the valid range of a third order model yields
a2 a3
vˆ be << = 12VT and vˆ be << = 20VT
a4 a5
1 W 2 dId W
Id =
2
µCox
L
(
Vgs − Vt ) a1 =
dVgs
= µCox
L
(
Vgs − Vt )
Vgs = VGSQ Vgs = VGSQ
W 2I
= µCox VOV = DQ ≡ gm
L VOV
1 d2Id 1 W IDQ
a2 = 2
= µCox = 2
2 dVgs 2 L VOV
Vgs = VGSQ
a3 = 0
1 a2 1 v̂ gs
HD2 ≅ vˆ gs = HD3 = 0
2 a1 4 VOV
Small second harmonic distortion in the drain current requires the G-S
voltage excursion to be much smaller than the quiescent point gate
overdrive (VOV=VGS-Vt)
An idealized square-law device does not introduce high order distortion
– However, this is not true for a real short-channel MOSFET
Relevant effects
– Velocity saturation, mobility reduction due to vertical field
– Biasing in moderate or weak inversion
– Nonlinearity in the device’s output conductance
– F
Relationship to gm/ID
1 '' VGS
I0
a2 2 ID 1 1 gm ''
ID = e nVT
a1
= ' =
ID
=
2 ( VGS − Vt ) 4 ID (nVT )2
1 ''
In strong inversion, 2/(gm/ID) can a2 2 ID 1 1 gm
be used as an estimate for “VOV” in = ' = =
distortion calculations
a1 ID 2nVT 2 ID
0.7
0.6
Weak Inversion
a2/a1 / (gm/I D) 0.5
Approximation
Square Law
0.4 Approximation
0.3
0.2
0
5 10 15 20 25
gm /I D [S/A]
Simulation Example
.include /usr/class/ee214b/hspice/ee214_hspice.sp
m1 d g s s nmos214 w=34u l=0.18u
= 10 ௦ = 50
vd d 0 1
vg g 0 sin (0.6 50m 1meg)
c1 s 0 1 Expected:
ib s 0 1m
.op 1 ௦
.tran 1n 20u ଶ ≅ = 6.25%
.fourier 1meg i(vd) 4 2/( / )
.options post brief accurate delmax = 10n
.end
V
Icd = Ic1 − Ic2 = αIEE tanh id
2VT
Ic1 Ic2
1 3 2 5
tanh ( x ) = x − x + x − +...
3 15
+
Vid
αIEE αIEE αIEE
- a1 = ≡ Gm a3 = − a5 =
2VT 24VT3 240VT5
IEE
a3
v̂id << = 10VT
a5
2 2
1 a3 2 1 vˆ id 1 vˆ be
HD3,BJTdiff ≅ vˆ id = HD3,BJT ≅
4 a1 48 VT 24 VT
2
V V
Iod = Id1 − Id2 = ISS id 1 − id
VOV 2VOV
2
x 1 1 5
Id1 Id2 x 1 − = x − x3 − x + ...
2 8 128
ISS a3
Third-order model is accurate for: v̂id << = 4VOV
a5
2
1 a3 2 1 vˆ id
HD3 ≅ vˆ id =
4 a1 32 VOV
Vout
Vin
+ Sε
Si – a So
–
Sfb
f
Substitute this expression for So into the result on the previous page and
compare coefficients to find bi
a1
b1 =
1+ a1f
Second-order terms
Third-order terms
Comments
VCC
a1 gm 1 IC 1
b1 = = b2 =
1+ a1f 1+ gmRE 2
2 V (1+ g R )3
T m E
1 IC 1 IC
(1+ gmRE ) − g R
6V 3 2 V3 m E 1
T T gmRE = ⇒ b3 = 0 ⇒ HD3 = 0
b3 =
(1+ gmRE )5 2
Parameters:
RD
= 14.2 ி = 1Ω = 10
M2
̂ = 150
௦ଵ = 150 ∙ 1Ω = 150
M3
Distortion estimate:
(Assuming that distortion from M1 dominates)
vo
M1 1 1 ଶ 1 1
௦ଵ
ଶ =
=
2 (1 + )ଶ ଵ ௦ଵ 4 (1 + )ଶ 2/( / )
ii RF RF
ଶ = 0.08%
Boris Murmann
Stanford University
Winter 2013-14
Continuous Discrete
Time Time
(Our focus)
Example 1
R
ic ic = a1vi + a2 vi2 + a3 vi3 + ...
vo
1 ICQ 1 ICQ
a1 = gm a2 = a3 =
C 2 VT2 6 VT3
vi
VI
vo −R
K(jω) = =
ic 1 + jωRC
vi = vˆ i cos ( ω t )
1 1
ic = a1vˆ i cos ( ω t ) + a2 vˆ i2 cos ( 2ω t ) + a3 vˆ i3 cos ( 3ω t ) + ...
2 4
The output voltage consists of the same tones, with their magnitude and
phase altered by the linear filter K(jω)
v o = K ( jω ) ⋅ a1vˆ i cos ( ω t + φω )
1 φm ω = ∠ K ( m ⋅ j ω )
+ K ( 2jω ) ⋅ a2 vˆ i2 cos ( 2ω t + φ2ω )
2
1
+ K ( 3jω ) ⋅ a3 vˆ i3 cos ( 3ω t + φ3ω ) + ...
4
Two-Tone Input
Substituting this input into the power series, and using the identities
shown below, the complete expression for the collector current is most
elegantly expressed as shown on the next slide
1
cos ( α ) cos ( β ) = cos ( α + β ) + cos ( α − β )
2
1
cos ( α ) cos ( β ) cos ( γ ) = cos ( α + β + γ ) + cos ( α + β − γ )
4
+ cos ( α − β + γ ) + cos ( α − β − γ )
a2 2
+ vˆ 1 cos ([ ω1 ± ω1] t ) + vˆ 22 cos ([ ω2 ± ω 2 ] t ) 0, 2ω1, 2ω2
2
Filtered Output
( )
+ K ( 2jω2 ) ⋅ vˆ 22 cos 2ω 2 t + φ2ω2 + vˆ 22 K ( 0 )
(
+ K ( j [ ω1 − ω2 ]) ⋅ 2vˆ 1 vˆ 2 cos [ ω1 − ω2 ] t + φω1−ω2 )
+ K ( j [ ω1 + ω 2 ]) ⋅ 2vˆ 1 vˆ 2 cos [ ω1 + ω 2 ] t + φω1+ω2
( )
a3
+ [...]
4
v o = a1K ( jωa ) vi + a2K ( jωa + jωb ) v i2 + a3K ( jωa + jωb + jωc ) vi3 + ...
H1( jωa ) H2 ( jωa + jωb ) H3 ( jωa + jωb + jωc )
Hm ( jωa, jωb,...)
and shift phase by
∠Hm ( jωa, jωb,...)
The arguments ωa, ωb, ωc, 6 are auxiliary variables taking on all
permutations of ω1, ±ω2, 6 ±ωm
−a1R
H1(jωa ) =
1 + jωaRC
−a2R
H2 (jωa , jωb ) =
1 + ( jωa + jωb ) RC
−a3R
H3 (jωa , jωb , jωc ) =
1 + ( jωa + jωb + jωc ) RC
1 a2 1 H2 ( jω1, jω1 )
HD2 v̂i v̂i
2 a1 2 H1 ( jω1 )
3 a3 2 3 H3 ( jω1, jω1, − jω 2 ) 2
IM3 v̂i v̂i
4 a1 4 H1 ( jω1 )
1
1 H2 ( jω, jω ) 1 a2 1 + 2jω RC 1 a2 1 + jω RC
HD2 = vˆ i = vˆ i = vˆ i
2 H1 ( jω ) 2 a1 1 2 a1 1 + 2jω RC
1 + jω RC
1 H3 ( jω, jω, jω ) 2 1 a3 1 + jω RC 2
HD3 = vˆ i = vˆ i
4 H1 ( jω ) 4 a1 1 + 3jω RC
ICQ = 1mA
0.5
VT
vˆ i1 = vˆ i2 =
0.4 5
IM3 [%]
0.3
0.2 Taylor
Volterra, ω 2/ω 1=0.9
0.1
0 -2 -1 0 1
10 10 10 10
ω 1RC
2 3
ic = a1v be + a2 vbe + a3 vbe + ...
vo −R
K(jω ) = =
ic 1 + jω RC
vbe 1
Kin (jω) = =
vi 1 + jωRinCin
Two-Tone Input
The two input tones are now processed by a linear filter before being
sent through the nonlinearity
At the base of the BJT, we have
( ) (
vbe = K in ( jω1 ) ⋅ vˆ 1 cos ω1t + ψ ω1 + K in ( jω 2 ) ⋅ vˆ 2 cos ω 2 t + ψ ω2 )
ψmω = ∠K in ( m ⋅ jω )
vbe = K in ( jωa ) v i
2
+ a2K ( jωa + jωb ) K in ( jωa ) v i
3
+ a3K ( jωa + jωb + jωc ) K in ( jωa ) v i
2 2
K in ( jωa ) vi = K in ( jωa ) {vˆ 1 cos ( ω1t ) + vˆ 2 cos ( ω 2 t )}
2
= K in ( jω1 ) ⋅ vˆ 1 cos ω1t + ψ ω1 + K in ( jω 2 ) ⋅ vˆ 2 cos ω2 t + ψ ω2
( ) ( )
2
(
= K in ( jω1 ) ⋅ vˆ 12 cos {ω1 ± ω1} t + ψ ω1 ± ψ ω1 )
2
(
+ K in ( jω 2 ) ⋅ vˆ 22 cos {ω 2 ± ω2 } t + ψ ω2 ± ψ ω2 )
(
+ K in ( jω1 ) K in ( jω 2 ) ⋅ vˆ 1 vˆ 2 cos {ω1 ± ω2 } t + ψ ω1 ± ψ ω2 )
= K in ( jωa ) K in ( jωb ) v i2
1 −a1R
H1(jωa ) =
1 + jωaRC 1 + jωaRC
1 1 −a2R
H2 (jωa , jωb ) =
1 + jωaRC 1 + jωbRC 1 + ( jωa + jωb ) RC
1 1 1 −a3R
H3 (jωa , jωb , jωc ) =
1 + jωaRC 1 + jωbRC 1 + jωcRC 1 + ( jωa + jωb + jωc ) RC
Example 3
ij
C j0 C j0 1
Cj = M
= M M
VOQ + v o ψ0 + VOQ ψ0 + VOQ + v o
1+
ψ0 ψ0 ψ0 + VOQ
Using C j0 1 1
C jQ =
VOQ
M
(1 + x ) M
= 1 − Mx +
2
( )
M + M2 x 2 + ...
1 +
ψ0
M M + M2
VR = VOQ + ψ0 b1 = − b2 =
VR 2VR
we can write C jQ
Cj = = C jQ 1 + b1v o + b2 v o2 + ...
M
vo
1 +
VR
dv o dv dv 1 dv 2 1 dv 3
ij = C j = C jQ 1 + b1v o + b2 v o2 + ... o = C jQ o + b1 o + b3 o
dt dt dt 2 dt 3 dt
2 3
(
i = C jQ + CI ) dvdto + C jQ 21 b1 dvdto + 31 b3 dvdto
dv o 1 dv 2 1 dv 3
(
vi = v o + i ⋅ R = v o + R C jQ + CI ) dt
+ RC jQ b1 o + b3 o
2 dt 3 dt
d d d
vi = v o + RC0 v o + RC1 v o2 + RC2 v 3o
dt dt dt
C0 = C jQ + Ci
b
C1 = 1 C jQ = −
MC jQ b1
C2 = C jQ = −
M + M2 C jQ( )
2 2VR 3 6VR2
Linear Nonlinear
݀ ଶ
݅ଶ = ܥଵ ݒ
݀ ݐ
݀ ଷ
݅ଷ = ܥଶ ݒ
݀ ݐ
ଶ
,ଶ ଵ (2)
ଶ = =
,ଵ ()
d
+ RC0 H1 ( jωa ) v i + H2 ( jωa , jωb ) vi2 + H3 ( jωa , jωb , jωc ) vi3
dt
d 2
+ RC1 H1 ( jωa ) vi + H2 ( jωa , jωb ) vi2 + H3 ( jωa , jωb , jωc ) vi3
dt
d 3
+ RC2 H1 ( jωa ) v i + H2 ( jωa , jωb ) vi2 + H3 ( jωa , jωb , jωc ) vi3
dt
First order
d
1 vi = 1 + RC0 H1 ( jωa ) vi
dt
d
1 = 1 + RC0 H1 ( jωa ) = 1 + RC0 jωa H1 ( jωa )
dt
1
∴ H1 ( jωa ) =
1 + RC0 jωa
Second order
d d 2
0 vi2 = 1 + RC0 H2 ( jωa , jωb ) vi2 + RC1 H1 ( jωa ) vi
dt dt
d
H2 ( jωa , jωb ) = ( jωa + jωb ) H2 ( jωa , jωb )
dt
2
H1 ( jωa ) vi = H1 ( jωa ) H1 ( jωb ) vi2
d
H1 ( jωa ) H1 ( jωb ) = ( jωa + jωb ) H1 ( jωa ) H1 ( jωb )
dt
0 = 1 + RC0 ( jωa + jωb ) H2 ( jωa , jωb ) + RC1 ( jωa + jωb ) H1 ( jωa ) H1 ( jωb )
Third order
d
0 vi3 = 1 + RC0 H3 ( jωa , jωb , jωc ) vi3
dt
d
+ RC1 (
2 H1 ( jωa ) vi ) (H ( jωa, jωb ) vi2 ) Second-order
dt
2 Interaction Term
d 3
+ RC2 H1 ( jωa ) vi
dt
∴ H3 ( jωa , jωb , jωc ) = −RC2 ( jωa + jωb + jωc ) H1 ( jωa ) H1 ( jωb ) H1 ( jωc )
2
RC1 ( 2jω ) H1 ( jω ) H1 ( 2jω )
1 H2 ( jω, jω ) 1
HD2 = vˆ i = vˆ i = ω RC1 H1 ( jω ) H1 ( 2jω ) vˆ i
2 H1 ( jω ) 2 H1 ( jω )
1 v̂
∴ HD2 = Mω RC jQ H1 ( jω ) H1 ( 2jω ) i
2 VR
3
RC2 ( 3jω ) (H1 ( jω ) ) H1 ( 3jω )
1 H3 ( jω, jω, jω ) 2 1
HD3 = vˆ i = vˆ i2
4 H1 ( jω ) 4 H1 ( jω )
3 2
= ω RC2 H1 ( jω ) H1 ( 3jω ) vˆ i2
4
2
1 v̂
( )
∴ HD3 = M + M2 ω RC jQ H1 ( jω ) H1 ( 3jω ) i
8 VR
v̂i = 0.5V
R = 25Ω
Ci = 0
C jQ = 1pF
M = 0.3
VR = VOQ + ψ 0 = 2.2V
1
= 6.4GHz
2πRC jQ
Covered in EE314A
– RF-centric metrics
• Intercept points
• 1-dB gain compression point
Other
– Nonlinearity in MOSFET output conductance
– Nonlinearities in passive components
– Distortion cancelation or pre-/post-compensation techniques
– Cascading nonlinearities
– Series reversion
– Distortion in clipped or amplitude limiting waveforms
• E.g. in oscillators
References (1)
Boris Murmann
Stanford University
Winter 2013-14
Motivation
Device-to-Device Mismatch
C1 C2
C1 − C2 = ∆C
R1 R2 R1 − R2 = ∆R
AP Sometimes referred to as
σ∆P = “Pelgrom’s rule”
WL AP: “Pelgrom coefficient”
where W·L represents the area of the device, and P is the device
parameter under consideration
[M. Pelgrom et al., “Matching Properties of MOS Transistors,” JSSC, Oct. 1989]
Parameter Value
Avt (MOSFET) 5 mV-µm
A∆β/β (MOSFET) 1 %-µm
A∆Is/Is (Bipolar transistor) 2 %-µm
A∆β/β (Bipolar transistor) 4 %-µm
A∆C/C (MIM capacitor) 1 %-µm
A∆R/R (Poly resistor) 3 %-µm
1%
σ∆C/C = = 0.1% 3σ∆C/C = 0.3%
100
http://en.wikipedia.org/wiki/Image:Standard_deviation_diagram.svg
௧ = ி +
௫
௫
௫
(௧ ) = =
௫
௫
௫ ௧
(∆௧ ) = 2 =
∆β
∆I = I1 − I2 ≅ −gm ∆Vt + I1
β
∆I g ∆β
≅ − m ∆Vt +
I1 I1 β
2
S 2 2 2
σ ∆I = 10 ⋅ 2.5mV + ( 0.5% ) = ( 2.5%) + ( 0.5%) = 2.54%
I1 A
If the device area is fixed, choose small gm/ID (large VOV) for improved
matching
ଶ
∆ g ୫ ଶ௧
≅
Text, p. 101
AVt = 4 mVµm
Aβ = 1 %µm
W/L = 2 µm/0.2µm
(β)
Mismatch W
2 ID = 500µA
0
5 10 15 20 25
gm /I D [S/A]
V1 V2=V1
I1 I2 ∆I = I1 − I2 ≅ gm Vwire
M1
∆I gm
M2 ≅ Vwire
I1 I1
- Vwire +
∆I ∆IS
≅
I1 IS
2%
σ ∆I = = 2.4%
I1 0.7
RC1 RC2 RC RC
Ios/2
Vi2 Vi2
IB2
IEE IEE
OFFSETS
VEE VEE
Analysis (1)
If Vod = 0, then
IC1RC1 = IC2RC2
IC1 RC2
∴ =
IC2 RC1
Thus
R I
Vos = VT ln C2 S2
RC1 IS1
For small mismatches ∆RC << RC and ∆IS << IS, it follows that
−1
∆RC ∆IS gm ∆RC ∆IS
Vos ≅ VT − − = − −
RC IS ID RC IS
P. G. Drennan et al., "Implications of Proximity Effects for Analog Design," Proc. CICC, pp.169-176, Sep. 2006.
Transistors near the well edge may see a threshold shift of up to 50mV
[Text, p. 98]
[C.M. Mezzomo, et al., "Characterization and Modeling of Transistor Variability in Advanced CMOS
Technologies," IEEE Transactions on Electron Devices, vol. 58, no. 8, pp. 2235-2248, Aug. 2011]